paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1911.06952
1
1911
2019-11-16T04:02:12
The landscape of Saturn's internal magnetic field from the Cassini Grand Finale
[ "astro-ph.EP" ]
The Cassini mission entered the Grand Finale phase in April 2017 and executed 22.5 highly inclined, close-in orbits around Saturn before diving into the planet on September 15th 2017. Here we present our analysis of the Cassini Grand Finale magnetometer (MAG) dataset, focusing on Saturn's internal magnetic field. These measurements demonstrate that Saturn's internal magnetic field is exceptionally axisymmetric, with a dipole tilt less than 0.007 degrees (25.2 arcsecs). Saturn's magnetic equator was directly measured to be shifted northward by ~ 0.0468 +/- 0.00043 (1-sigma) $R_S$, 2820 +/- 26 km, at cylindrical radial distances between 1.034 and 1.069 $R_S$ from the spin-axis. Although almost perfectly axisymmetric, Saturn's internal magnetic field exhibits features on many characteristic length scales in the latitudinal direction. Examining Br at the a=0.75 $R_S$, c=0.6993 $R_S$ isobaric surface, the degrees 4 to 11 contributions correspond to latitudinally banded magnetic perturbations with characteristic width similar to that of the off-equatorial zonal jets observed in the atmosphere of Saturn. Saturn's internal magnetic field beyond 60 degrees latitude, in particular the small-scale features, are less well constrained by the available measurements, mainly due to incomplete spatial coverage in the polar region. A stably stratified layer thicker than 2500 km likely exists above Saturn's deep dynamo to filter out the non-axisymmetric internal magnetic field. A heat transport mechanism other than pure conduction, e.g. double diffusive convection, must be operating within this layer to be compatible with Saturn's observed luminosity. The latitudinally banded magnetic perturbations likely arise from a shallow secondary dynamo action with latitudinally banded differential rotation in the semi-conducting layer.
astro-ph.EP
astro-ph
Draft version November 19, 2019 Typeset using LATEX twocolumn style in AASTeX63 9 1 0 2 v o N 6 1 . ] P E h p - o r t s a [ 1 v 2 5 9 6 0 . 1 1 9 1 : v i X r a The landscape of Saturn's internal magnetic field from the Cassini Grand Finale Hao Cao,1, 2, 3 Michele K. Dougherty,3 Gregory J. Hunt,3 Gabrielle Provan,4 Stanley W.H. Cowley,4 Emma J. Bunce,4 Stephen Kellock,3 and David J. Stevenson2 1Department of Earth and Planetary Sciences, Harvard University, 20 Oxford Street, Cambridge, MA 02138, USA 2Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA 3Physics Department, The Blackett Laboratory, Imperial College London, London, SW7 2AZ, UK 4Department of Physics and Astronomy, University of Leicester, Leicester, LE1 7RH, UK (Received July 19, 2019; Revised Oct. 31, 2019; Accepted Nov. 7, 2019) ABSTRACT The Cassini mission entered the Grand Finale phase in April 2017 and executed 22.5 highly inclined, close-in orbits around Saturn before diving into the planet on September 15th 2017. Here we present our analysis of the Cassini Grand Finale magnetometer (MAG) dataset, focusing on Saturn's internal magnetic field. These measurements demonstrate that Saturn's internal magnetic field is exceptionally axisymmetric, with a dipole tilt less than 0.007 degrees (25.2 arcsecs). Saturn's magnetic equator was directly measured to be shifted northward by ∼ 0.0468 ± 0.00043 (1σ) RS, 2820 ± 26 km, at cylindrical radial distances between 1.034 and 1.069 RS from the spin-axis. Although almost perfectly axisym- metric, Saturn's internal magnetic field exhibits features on many characteristic length scales in the latitudinal direction. Examining Br at the a = 0.75 RS, c = 0.6993 RS isobaric surface, the degree 4 to 11 contributions correspond to latitudinally banded magnetic perturbations with characteristic width ∼ 15◦, similar to that of the off-equatorial zonal jets observed in the atmosphere of Saturn. Saturn's internal magnetic field beyond 60◦, in particular the small-scale features, are less well constrained by the available measurements, mainly due to incomplete spatial coverage in the polar region. Magnetic fields associated with the ionospheric Hall currents were estimated and found to contribute less than 2.5 nT to Gauss coefficients beyond degree 3. The magneto-disk field features orbit-to-orbit variations between 12 nT and 15.4 nT along the close-in part of Grand Finale orbits, offering an opportunity to measure the electromagnetic induction response from the interior of Saturn. A stably stratified layer thicker than 2500 km likely exists above Saturn's deep dynamo to filter out the non-axisymmetric internal magnetic field. A heat transport mechanism other than pure conduction, e.g. double diffusive convection, must be operating within this layer to be compatible with Saturn's observed luminosity. The latitudinally banded magnetic perturbations likely arise from a shallow secondary dynamo action with latitudinally banded differential rotation in the semi-conducting layer. Keywords: Saturn, interior -- Magnetic fields -- Geophysics 1. INTRODUCTION Intrinsic magnetic field is a fundamental property of a planet. Not only is it a key factor in determining the electromagnetic environment of a planetary body, it also serves as a key diagnostic of the interior struc- ture and dynamics of the host planet (Stevenson 2003, 2010). A strong planetary-scale magnetic field most Corresponding author: Hao Cao [email protected] likely originates from dynamo action within the planet, the operation of which requires a large volume of elec- trically conducting fluid and "fast" and complex fluid motions (Steenbeck et al. 1966; Steenbeck & Krause 1966; Parker 1955; Krause & Radler 1980; Roberts & Stix 1971; Roberts & King 2013). For gas giant dy- namos, metallic hydrogen is the electrically conducting fluid, secular cooling drives "fast" fluid motions, while the rapid background rotation promotes the generation of large-scale magnetic fields (Christensen 2010). The warm interior conditions of the present-day Jupiter and 2 Cao et al. Saturn makes the transition from molecular to metallic hydrogen a gradual process: the electrical conductivity rises rapidly yet continuously from negligible values in the 1-bar atmosphere to significant values in the Mbar region (Weir et al. 1996; Liu et al. 2008). The transi- tion from magnetohydrodynamics (MHD) in the deep dynamo to hydrodynamics in the outer layers inside gas giants is also likely to be gradual (Cao & Stevenson 2017a). It is generally believed that the transition from hydrodynamics to MHD underlies the transition from 100 m/s rapid zonal flows in the non-conducting outer layer to cm/s−mm/s slow deep dynamo flows inside the gas giants (Kaspi et al. 2018; Guillot et al. 2018). How- ever, the physical mechanism of this dynamical transi- tion, in particular that at mid-to-high latitudes, remains unknown. On the other hand, although fluid motions in the semi-conducting layer may not be able to sus- tain dynamo action on their own, they could modify the deep dynamo generated magnetic field and produce ob- servable features outside the planet such as magnetic perturbations spatially correlated with deep zonal flows (Gastine et al. 2014; Cao & Stevenson 2017a) and time variation of the magnetic field (secular variation) (Moore et al. 2019). Saturn's magnetic field has been measured in-situ by four space missions, Pioneer 11 (Smith et al. 1980; Acuna & Ness 1980), Voyager 1 (Ness et al. 1981), Voyager 2 (Ness et al. 1982; Connerney et al. 1982), and Cassini (Dougherty et al. 2005; Burton et al. 2009; Cao et al. 2011, 2012; Dougherty et al. 2018). These measurements revealed an almost perfectly axisymmet- ric, dipole dominant internal magnetic field with non- negligible north-south asymmetry (Dougherty et al. 2018) and a highly dynamic magnetosphere filled with periodic phenomena whose frequencies are close to the rotational frequency of Saturn (Andrews et al. 2012; Provan et al. 2018, 2019b). The periodic magnetic per- turbations in Saturn's magnetosphere are referred to as Planetary Period Oscillations (PPOs). The search for departures from perfect axisymmetry in the internal magnetic field of Saturn is of great interest, since it could yield the true rotation period of the deep interior (see current values derived from different measurements and methods: Anderson & Schubert 2007; Read et al. 2009; Mankovich et al. 2019; Militzer et al. 2019) and provide key constraints on the dynamo process inside Saturn. However, this search is complicated by the existence of ionospheric and field-aligned currents (FACs) at Sat- urn, which feature both PPO and non-PPO components (Hunt et al. 2014, 2015, 2018). Here we would like to stress that the deep dynamo layer of Saturn rotates very much like that of a solid body from the view of observers in an inertial frame, since the expected cm/s − mm/s differential rotation is only about one part in a million when compared to the ∼ 10 km/s background rotation. Among the existing measurements, those from the Grand Finale phase of the Cassini mission (Table 1, Figs. 1 - 4) are the most sensitive to the internal mag- netic field due to their proximity to the planet and the highly inclined orbit. So far, the analysis of Saturn's internal magnetic field has been mostly restricted to the traditional Gauss coefficients representation, in which the internal planetary magnetic field is expressed as a function of the Gauss coefficients (gm n , hm n ) with Br,θ,φ(r, θ, φ) = [gm n f g r,θ,φ(r, θ, φ)+hm n f h r,θ,φ(r, θ, φ)] (cid:88) (cid:88) n m (1) where the functional form of f g,h r,θ,φ can be easily found (e.g., Eqns. 3 - 5 in Dougherty et al. 2018) and repro- duced in Appendix A for convenience. An equivalent and likely more fundamental representation of the in- ternal magnetic field of a planet is the Green's function which maps the internal magnetic field from the dynamo surface (or the planetary surface) to the field outside (e.g. Gubbins & Roberts 1983; Backus et al. 1996; John- son & Constable 1997): (cid:90) 2π (cid:90) π r 0 0 Bobs r,θ,φ(r, θ, φ) = r (θ(cid:48), φ(cid:48))Gr,θ,φ(µ) sin θ(cid:48)dθ(cid:48)dφ(cid:48), BrD (2) here BrD is the radial component of the magnetic field at the dynamo surface (a spherical surface with r = rD in the traditional geophysical formulation, see next para- graph for non-sphericity of isobaric surface inside Sat- urn), Bobs r,θ,φ are three components of the magnetic field measured above the "dynamo surface", and µ is the co- sine of the angle between the position vectors r and r(cid:48) (see Appendix B for more details). The Green's function not only yields an equivalent description of the internal magnetic field, it also admits a simple and straightfor- ward physical interpretation: it describes how sensitive the magnetic field measured outside the planet is to the field at different locations on the dynamo surface. The Green's function has been applied to analyzing the mag- netic field of the Earth (e.g. Johnson & Constable 1997; Jackson et al. 2007), Mars (Purucker et al. 2000; Moore & Bloxham 2017), and Jupiter (Moore et al. 2017). Saturn is the most oblate planet in the solar system, with a measured flattening f = (a − c)/a = 9.8% at the 1-bar surface, where a and c are the equatorial ra- dius and polar radius respectively. The flattening of the interior isobaric surface decreases as the pressure level increases (e.g., see Fig. 2 in Cao & Stevenson 2017b). According to the latest Saturn interior model (Militzer Saturn's internal magnetic field from Cassini Grand Finale 3 et al. 2019) constrained by the Cassini Grand Finale gravity measurements (Iess et al. 2019), the flattening of the isobaric surface decreases to 6.76% at a = 0.75RS and 5.88% at a = 0.65RS. The isobaric surfaces of gi- ant planets are not perfect ellipsoids due to their non- uniform density. The fractional deviation from ellip- soids ∆r/r, however, are on the order of 10−3 or less for Jupiter and Saturn (e.g., see Fig. 2 in Cao & Stevenson 2017b; Militzer et al. 2019), two orders of magnitude smaller than the dominant elliptical flattening. Thus, we treat the "dynamo surface" as ellipsoids when evalu- ating the properties of Saturn's internal magnetic field. Here we report our analysis of the Cassini Grand Fi- nale MAG dataset, focusing on Saturn's internal mag- netic field. It should be noted that the solution of Sat- urn's internal magnetic field were obtained with spheri- cal basis function, such as the spherical harmonics and the Green's functions on a sphere. However, the non- spherical shape of Saturn's "dynamo surface" was ex- plicitly considered when evaluating the properties of the resultant internal magnetic field. We have extended the analysis presented in Dougherty et al. (2018) in several ways: i) MAG data from the last 12.5 Cassini Grand Finale orbits are analyzed here together with those pre- sented in Dougherty et al. (2018), ii) an explicit search for internal non-axisymmetry is carried out, iii) the ef- fect of incomplete spatial coverage is demonstrated with regularized inversion, and iv) Green's functions were em- ployed in addition to the traditional Gauss coefficients in constructing models of Saturn's internal magnetic field, v) ionospheric current and their associated magnetic field are evaluated with a simple axisymmetric model, and vi) search for electromagnetic induction from the interior of Saturn and orbit-to-orbit varying "internal" field is carried out. In section 2, we present the main characteristics of the trajectory of Cassini Grand Finale orbits and the MAG measurements. In section 3, we present the directly measured position of Saturn's mag- netic equator and its spatial variations. In section 4, we present the sensitivity of Cassini Grand Finale MAG measurements to Saturn's axisymmetric internal mag- netic field at depth. In section 5, we present inversion of Saturn's axisymmetric internal magnetic field with different methods. In section 6, we present a search for electromagnetic induction from the interior of Sat- urn. In section 7, we present the orbit-to-orbit vari- ations in Saturn's "internal" quadrupole magnetic mo- ments. In section 8, we present a search for internal non- axisymmetry in Saturn's magnetic field. In section 9, we discuss the constraints and implications on Saturn's in- terior structure and dynamics. Section 10 presents a summary and outlook. 2. CASSINI GRAND FINALE TRAJECTORY AND MAG MEASUREMENTS Figure 1. Trajectory of typical Cassini Grand Finale or- bits. In panel A, the trajectory of Rev 291 from apoapsis to apoapsis is projected onto the meridional plane in which Z is along the spin-axis direction and ρ is in the cylindrical radial direction. Panel B shows the close-in part of the tra- jectory from three Cassini Grand Finale orbits in the same projection. For the blue-red color-coded trajectory, the red part is when the measured magnetic field strength > 10,000 nT . The dashed line shows r = 0.75 RS. Panels C shows the trajectory in latitude local time projection. The Grand Finale phase of the Cassini mission con- sists of 22.5 highly inclined, close-in orbits around Sat- urn between Apr 23rd 2017 (apoapsis time of first Grand Finale orbit Rev 271) and Sep 15th of 2017 (periapsis time of the last orbit Rev 293). Each Grand Finale or- bit took ∼ 6.5 Earth days, with periapsis in the gap between Saturn and the inner edge of the D-ring and apoapsis near the orbit of Titan (Fig. 1). The trajectory and magnetic field measurements from selected Cassini Grand Finale orbits are shown in Figs. 1 - 4. Table 1 lists the periapsis information of all Cassini Grand Fi- nale orbits including time, periapsis distance, altitude, latitude, and local time. Fig. 1 shows the trajectory of a few typical Cassini Grand Finale orbits (the specific orbit shown in panel A is Rev 291, the ones shown in panels BC are Revs 271, 276, 292). The orbits featured 10121416182022 [RS]-6-4-20246Z [RS]Aug 31Sep 01Sep 02Sep 03Sep 04Sep 050246800.511.5-2-1.5-1-0.500.511.521012141618202224Local Time [hr]-80-60-40-20020406080Latitude [deg]BC [RS]Z [RS]02468A 4 Cao et al. Figure 2. Characteristics of the trajectory of Cassini Grand Finale orbits. Panel A shows the periapsis distance from the center of Saturn, panel B shows the periapsis latitude while panel C shows the periapsis local time as a function of the orbit (Rev) number. inclination ∼ 62◦, the periapsis distance from the center of Saturn varied between 1.064 RS and 1.02 RS (1 RS = 60268 km), the periapsis latitudes were -6.2◦ ± 1◦ except that of the dive-in orbit which was ∼ 10◦, the periapsis local times were about ±1 hour around local noon (Fig. 2). Fig. 3 shows the measured magnetic field strength and the azimuthal component along one Cassini Grand Finale orbit, Rev 291, from apoapsis to apoapsis. It can be seen that the measured field strength ranges from < 2 nT to > 20,000 nT . Thus, all four dynamical ranges of the fluxgate magnetometer (Dougherty et al. 2004) were activated during a Cassini Grand Finale orbit. During the Grand Finale phase, the highest dynamical range of the fluxgate magnetometer, range 3, which can measure field above 10,000 nT and up to 44,000 nT with a digi- tization of 5.4 nT were activated for the first time since the Cassini Earth Swing-by (Southwood et al. 2001). The minimum field strength along this orbit, 1.74 nT , Figure 3. Characteristics of the magnetic field measure- ments along a typical Cassini Grand Finale orbit from apoap- sis to apoapsis (shown here is Rev 291). The top panel shows the total amplitude of the magnetic field, and the bottom panel shows the azimuthal component, which exhibits vari- ous magnetospheric features, including Auroral FACs, Intra- D ring FACs, Planetary Period Oscillations (PPOs), and Enceladus fluxtube crossing. was recorded during the crossing of the magnetodisk on the nightside (Fig. 3). To transform the vector magnetic field measurements from the spacecraft coordinate to an astronomical coor- dinate (e.g. the Saturn centered coordinate), the atti- tude of the spacecraft needs to be known to high preci- sion. For example, the spacecraft attitude needs to be known to better than 0.25 milliradian (mrad) for the vector magnetic field to be known to within 5 nT from the true values near the periapsis. The star tracker on- board Cassini was suspended intermittently during the Grand Finale orbits, which we refer to as Star ID sus- pensions. Table 2 lists the timing of the Star ID suspen- sions along each Grand Finale Orbit. The attitude of the spacecraft during the Star ID suspensions were recon- structed using information from the gyroscopes onboard (see Burk 2018, for more information). Spacecraft rolls around two different spacecraft axes were designed and carried out along four Grand Finale orbits: Revs 272, 273, 284, 285. These spacecraft rolls enabled in-flight calibration of range 3 of the fluxgate magnetometer. 11.011.021.031.041.051.061.07Periapsis Distance [RS]-10-50510Periapsis Latitude [deg]271272273274275276277278279280281282283284285286287288289290291292293Rev Number10.51111.51212.51313.5Periapsis Local Time [hr]271272273274275276277278279280281282283284285286287288289290291292293271272273274275276277278279280281282283284285286287288289290291292293Local NoonPlanetary EquatorABC100101102103104B [nT]Aug 30Aug 31Sep 01Sep 02Sep 03Sep 04Sep 05Sep 06UTC Time, Year =2017B [nT]Enceladus fluxtube crossingmagnetodisk crossing40 nT400 nT10,000 nTAuroral FACsPlanetary Period Oscillations (PPOs)Auroral FACsLow-latitude (Intra-D ring) FACs-30-20-1001020304050 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 aAltitude here is defined as the minimum distance to the 1-bar spheroid with a = 60268km, c = 54364km. 2017-116T09:03:34 2017-122T19:42:15 2017-129T06:16:39 2017-135T16:45:20 2017-142T03:14:28 2017-148T14:26:22 2017-155T01:42:28 2017-161T12:53:15 2017-167T23:55:43 2017-174T10:57:42 2017-180T22:14:15 2017-187T09:35:23 2017-193T20:48:00 2017-200T07:54:43 2017-206T18:59:19 2017-213T06:09:10 2017-219T17:23:16 2017-226T04:23:03 2017-232T15:23:00 2017-239T02:18:10 2017-245T13:13:00 2017-252T00:09:45 26 Apr 2017 02 May 2017 09 May 2017 15 May 2017 22 May 2017 28 May 2017 04 Jun 2017 10 Jun 2017 17 Jun 2017 23 Jun 2017 29 Jun 2017 06 Jul 2017 12 Jul 2017 19 Jul 2017 25 Jul 2017 01 Aug 2017 07 Aug 2017 14 Aug 2017 20 Aug 2017 27 Aug 2017 02 Sep 2017 09 Sep 2017 15 Sep 2017 1.048203 1.047782 1.044115 1.043232 1.043970 1.063769 1.063580 1.055669 1.054660 1.055312 1.060773 1.060853 1.046322 1.045308 1.045589 1.047326 1.047682 1.027228 1.026304 1.025832 1.026003 1.026560 1.020827 [hr] 13.135 13.054 12.974 12.894 12.812 12.738 12.659 12.581 12.501 12.422 12.345 12.266 12.185 12.104 12.024 11.945 11.864 11.779 11.696 11.613 11.531 11.448 10.749 2963.16 2939.30 2719.79 2667.90 2713.47 3910.65 3901.04 3427.14 3367.97 3409.05 3740.92 3747.78 2875.56 2816.82 2835.97 2943.12 2967.09 1737.60 1684.73 1659.24 1672.41 1709.06 1443.63 [deg] -5.296 -5.364 -5.429 -5.486 -5.535 -5.717 -5.793 -5.907 -5.974 -6.047 -6.160 -6.239 -6.366 -6.456 -6.539 -6.632 -6.725 -6.826 -6.924 -7.026 -7.126 -7.229 9.559 Saturn's internal magnetic field from Cassini Grand Finale 5 Table 1. Periapsis information of the Cassini Grand Finale orbits Rev Num Periapsis Date UTC Time Radial Altitudea Latitude Local Time Distance [RS] [km] 2017-258T10:31:41.755 The absolute scale of the fluxgate magnetometer was de- termined via comparing the simultaneous measurements carried out by the fluxgate magnetometer (Southwood et al. 2001) and the helium magnetometer (Smith et al. 2001) during the Earth Swing-by. Fig. 3B shows the measured azimuthal component, Bφ, along Rev 291 which remains within ± 50 nT and exhibits various magnetospheric features including the auroral FACs (Hunt et al. 2014, 2015, 2018), low-latitude (intra-D ring) FACs (Dougherty et al. 2018; Khurana et al. 2018; Provan et al. 2019a; Hunt et al. 2019), cross- ing of the Enceladus fluxtube (Sulaiman et al. 2018), and PPOs (Provan et al. 2019b). Fig. 4 shows the total am- plitude and all three components of the measured field in the Saturn centered KRTP coordinate within ± 4 hours of the periapsis along Rev 291. KRTP is a right-handed spherical polar coordinate, with its origin at the cen- ter of mass of Saturn, the polar axis (zenith reference) being the spin axis of Saturn, rotating at the IAU Sys- tem III rotation rate of Saturn, while r, θ, and φ denote radial, meridional, and azimuthal directions. The Ence- ladus fluxtube crossing, auroral FACs, and the intra-D ring FACs are better delineated in this zoomed-in ver- sion. The radial and meridional components exhibit a dipolar geometry, with Br being positive (negative) in the northern (southern) hemisphere while Bθ remains positive. The peak field strength is not encountered at the periapsis but at mid-latitude in the southern hemi- sphere. The overall features of the measured magnetic field are highly repeatable from orbit to orbit, although the magnetospheric features such as auroral FACs and intra-D ring FACs do exhibit orbit to orbit variations (Provan et al. 2019a; Hunt et al. 2019). 3. SATURN'S MAGNETIC EQUATOR POSITION AND ITS SPATIAL VARIATIONS The highly inclined nature of the Cassini Grand Finale orbits enabled direct determination of Saturn's magnetic equator positions, defined as where the cylindrical radial component of the magnetic field, Bρ, vanishes. Fig. 5 displays the measured magnetic equator positions pro- jected onto the ρ − Z plane, where ρ is distance from the spin-axis of Saturn and Z is distance from the plane- tary equator of Saturn defined by the center of mass with northward being positive. Other than the Cassini Grand Finale measurements, the predictions from the Cassini 6 Cao et al. Table 2. Star ID (SID) suspension time along the Cassini Grand Finale Orbits Rev Num 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 First SID suspension Second SID suspension 2017-116T08:35:19.000 to 09:54:57.854 None 2017-122T14:55:09 to 17:59:27 2017-128T18:37:17 to 23:35:19 2017-135T16:23:28 to 19:28:00 2017-142T02:52:31 to 05:57:03 2017-148T13:54:12 to 16:37:24 2017-155T00:28:33 to 02:10:50 2017-161T12:32:51 to 16:10:04 2017-167T23:47:12 to 168T01:09:29 2017-174T10:37:32 to 14:14:45 2017-180T20:04:55 to 23:46:13 2017-187T09:06:11 to 10:03:09 2017-193T19:22:30 to 19:47:22 2017-199T20:13:24 to 200T01:11:02 2017-206T14:12:54 to 17:17:12 2017-213T05:27:28 to 06:57:03 2017-219T15:51:03 to 16:20:43 2017-226T02:51:54 to 03:21:09 2017-232T15:06:47 to 15:52:50 2017-239T00:44:31 to 04:07:28 2017-245T12:44:47 to 14:18:00 2017-251T23:43:37 to 252T02:06:37 2017-258T10:11:19 to End of Mission 2017-122T18:19:40 to 20:53:22 2017-129T03:53:09 to 08:51:11 None None None None None None None None None 2017-193T20:13:51 to 22:21:18 2017-200T05:30:20 to 10:27:58 2017-206T17:38:06 to 20:18:02 None 2017-219T16:48:11 to 18:43:33 2017-226T04:12:29 to 2017-226T06:12:09 None None None None None 11 model (Dougherty et al. 2018) and the Cassini Saturn Orbital Insertion (SOI) measurements are shown in Fig. 5 as well. It can be seen that Saturn's magnetic equator is consistently displaced northward from the planetary equator. The measurements and the model predictions further demonstrate that the northward displacement of the magnetic equator, ZM agEq, is not constant but varies as a function of ρ. Along the Grand Finale or- bits where ρ ∼ 1.05RS, the displacement is ∼ 2820 km (0.0468 RS). Along SOI, the spacecraft crossed the mag- netic equator twice near ρ ∼ 2.5RS, where the displace- ment of the magnetic equator is ∼ 2300 km (0.0382 RS). The data-model comparison strongly suggests the ax- isymmetric part of the internal magnetic field is respon- sible for the majority of the observed spatial variations in ZM agEq. In addition to the axisymmetric variations of ZM agEq with ρ, multiple origins of perturbations in Bρ (e.g. the PPOs and non-axisymmetric internal magnetic moments such as g1 1) could cause additional ZM agEq variations. Near the magnetic equator crossing along the Grand Finale orbits, the relationship between the vertical displacement from the magnetic equator, ∆ZM agEq = Z − ZM agEq, and Bρ can be approximated 1 and h1 as ∆ZM agEq [km] = 1.395 [km/nT ] · Bρ [nT ]. (3) Thus, a magnetic perturbation in Bρ of about 7.2 nT would cause a displacement of the magnetic equator po- sition by about 10 km. It should be noted that if such magnetic perturbations are of internal dipole origin (cor- responding to g1 1), the corresponding Bφ would be about 3.6 nT . 1 and h1 The measured peak-to-peak variations of ZM agEq at similar ρ are less than 18 km along the Grand Finale or- bits. If the observed variations are caused by the inter- nal non-axisymmetric dipole moments, the correspond- ing dipole tilt would be less than 0.01◦. A dipole tilt much larger than 0.01 degrees can be safely ruled out by the data (Fig. 6). Here we carried out an explicit search for m=1 non- axisymmetric patterns in the measured magnetic equa- tor positions in addition to the variations with ρ. We first removed a degree-5 polynomial fit of the measured ZM agEq with 1/ρ: ZM agEq(ρ) =0.215932/ρ5 − 0.600580/ρ4 + 0.651408/ρ3 − 0.331803/ρ2 + 0.084854/ρ + 0.029170, (4) Saturn's internal magnetic field from Cassini Grand Finale 7 Figure 5. Saturn's magnetic equator positions, defined as where the cylindrical radial component of the field vanishes (Bρ = 0), as measured along the Cassini Grand Finale or- bits and the Cassini Saturn Orbital Insertion (SOI). The ex- pected magnetic equator position based on the axisymmetric Cassini 11 model is over-plotted using the grey trace. It can be seen that Saturn's magnetic equator position varies as a function of distance from the spin-axis. The Cassini 11 model under predicts the measured magnetic equator posi- tions by about 20 km near ρ = 1.035, the closest sets of measurements to the spin-axis. Figure 6. Variations of Saturn's magnetic equatorial posi- tions as a function of longitude compared. Prediction from a 0.1◦ dipole tilt is over-plotted using the black trace. A degree-5 polynomial fitting, ZM agEq [RS] =0.215932/ρ5 − 0.600580/ρ4 + 0.651408/ρ3 − 0.331803/ρ2 + 0.084854/ρ + 0.0291700, in which ρ is also in the unit of [RS], has been removed from the measured magnetic equator positions. ual, is at a period of 10h49m30s, close to the dominant northern PPO period - strangely no sign of southern PPO period (Provan et al. 2019b). It should be noted that the peak amplitude of the sin(φ+φ0) pattern is less Figure 4. Characteristics of the magnetic field measure- ments along a typical Cassini Grand Finale orbit within ± 4 hours around periapsis (shown here is Rev 291). The top panel shows the total amplitude of the magnetic field, the radial and meridional component, while the bottom panel shows the azimuthal component, which exhibits various mag- netospheric features. in which both ZM agEq and ρ are in the units of RS. A degree-5 polynomial fit yields an adequate description of the mean position of the magnetic equator without in- troducing additional spatial variations. Then we search for a sin(φ + φ0) pattern in the residual magnetic equa- tor positions ∆ZM agEq (Fig. 6). Here φ is the east longitude in the spherical polar Saturn centered coor- dinate with a certain fixed rotation rate. We searched the possible range of rotation periods from 10h30m00s to 10h55m00s. The results are presented in Figs. 7 & 8. Interestingly, we find that the residual magnetic equa- tor position can be ordered into a sin(φ + φ0) pattern at three different rotation periods, 10h31m32s, 10h34m14s, and 10h49m30s. The period 10h34m14s is almost iden- tical to the internal rotation period of Saturn derived by Read et al. (2009) by considering the Arnol'd sec- ond stability criterion with the observed wind profile on Saturn. The "best" ordering, judged by the amplitude of the pattern and the root-mean-square (RMS) resid- 10:00:0011:00:0012:00:0013:00:0014:00:0015:00:0016:00:0017:00:00-2.5-2-1.5-1-0.500.511.522.5B,Br, [nT]104BBrB10:00:0011:00:0012:00:0013:00:0014:00:0015:00:0016:00:0017:00:00UTC Time, Sep 02 2017-50-40-30-20-1001020304050B [nT]Enceladus fluxtube crossingSouthernAuroral FACsNorthernAuroral FACsLow-latitude (Intra-D ring) FACsPeriapsis1.21.41.61.82.22.42.62.83 [RS]0.0350.040.0450.05Z [RS]Cassini Grand Finale DataCassini 11 ModelCassini SOI Outbound DataCassini SOI Inbound Data220023002400250026002700280029003000Z [km]12120180240300360Saturn East Longitude [deg], assuming P=10h33m38s-100-80-60-40-20020406080100Z (Mag Equator) [km]Prediction from 0.1o dipole tiltCassini Grand Finale Measurements060 8 Cao et al. dr of dynamo surface for Saturn. Local Rm is defined as Rm = UconvHσ/η, here Uconv is the convective velocity, (cid:12)(cid:12) is the conductivity scale-height, η = 1/µ0σ Hσ =(cid:12)(cid:12)σ/ dσ is the local magnetic diffusivity, where µ0 is the magnetic permeability and σ is the local electrical conductivity. According to the Saturn interior electrical conductivity model of Liu et al. (2008), local Rm reaches order 1 (10) at this depth if the convective velocity is on the order of 1 mm/s (cm/s). Thus, downward continua- tion of the potential field to this depth seems appropri- ate. Downward continuation of the potential field from the surface to certain depth inside the planet is only valid when there are no toroidal electrical currents in- between. Thus, downward continuation to depth much deeper than the a = 0.75 RS isobaric surface cannot be guaranteed since local dynamo action is expected to be- come important around this depth. Viewing the down- ward continued internal field around this depth would be most relevant for deciphering internal dynamics. Figure 7. Amplitude and root-mean-square (RMS) residual in searching for a m = 1 pattern in Saturn's magnetic equa- tor positions as a function of rotation rate of Saturn. Three dominant peaks are found at 10h49m30s, close to one of the planetary period oscillations period (Provan et al. 2019b), 10h34m14s, close to the one of the "internal" rotation rate derived from Saturn's 1-bar winds (Read et al. 2009), and 10h31m32s. than 6 km (thus the peak-to-peak variation is less than 12 km), translating into a dipole tilt of 0.0065◦ only. We will return to the search for internal non- axisymmetry with explicit modeling of the non- axisymmetric magnetic moments based on the vector magnetic field measurements in section 8. The analysis so far has established that Saturn's internal magnetic field is exceptionally axisymmetric. 4. THE SENSITIVITY OF CASSINI GRAND FINALE MAG MEASUREMENTS TO SATURN'S INTERNAL MAGNETIC FIELD AT DEPTH Before proceeding to build models of Saturn's inter- nal magnetic field from the Grand Finale MAG mea- surements, we first utilize the Green's function to for- ward calculate the sensitivity of the Grand Finale MAG measurements to Saturn's internal magnetic field at the "dynamo surface", adopted as the a = 0.75 RS, c = 0.6993 RS isobaric ellipsoid here. Estimation of the local magnetic Reynolds number Rm guided the choice Due to the highly axisymmetric nature of Saturn's in- ternal magnetic field, the Green's function can be inte- grated in the azimuthal direction first and the mapping from the field at depth to the measurements along the spacecraft trajectory reduces to Bobs r,θ,φ(r, θ) = r (θ(cid:48)) ¯Gr,θ,φ(µ) sin θ(cid:48)dθ(cid:48) BrD (5) (cid:90) π 0 where the overbar denotes azimuthal integration. It can be easily shown that ¯Gφ = 0: axisymmetric current-free magnetic field has no azimuthal component. Instead of switching to the confocal ellipsoidal coordi- nates to re-derive the Green's function, here we simply compute the Green's function for two different spheri- cal surfaces, r(cid:48) = 0.75 RS and r(cid:48) = 0.6993 RS, which bracket the a = 0.75 RS isobaric surface. Qualita- tively, the Green's function for the a = 0.75 RS iso- baric surface is expected to be close to G0.75RS near the equator and approach G0.6993RS towards the poles. Fig. 9 shows the azimuthally-integrated, area-weighted Green's function, ¯Gr,θ sin θ(cid:48), for three locations along a typical Cassini Grand Finale trajectory (these loca- tions are marked with blue crosses in Fig. 1B), which illustrates the sensitivity of the MAG measurements to Saturn's internal magnetic field at depth. r,θ r,θ is mostly sensitive to B0.75RS Taking the Green's function at the r(cid:48) = 0.75 RS surface for example, at periapsis along the trajectory (Fig. 9A), Bobs around similar latitude (−5◦) with a half-amplitude-half-width (HAHW) of ∼ 20 degrees in latitude. On the other hand, at -22◦ and +12◦ lat- Bobs itude. At mid-latitude (30◦) along the trajectory, Bobs at similar latitude (28.5◦) is mostly sensitive to B0.75RS is mostly sensitive to B0.75RS θ r r r r r 0123456Z(Mag Eq) Amplitude [km]10:30:0010:35:0010:40:0010:45:0010:50:0010:55:00Saturn Rotation Period [HH:MM:SS]012345Z(Mag Eq) RMS misfit [km]10h34m14s10h49m30s10h31m32s10:30:0010:35:0010:40:0010:45:0010:50:0010:55:00 Saturn's internal magnetic field from Cassini Grand Finale 9 Figure 8. Ordering of Saturn's magnetic equator positions as a m = 1 pattern in longitude at three different rotation periods. Figure 9. Area-weighted, azimuthally integrated Green's function for Saturn's axisymmetric internal magnetic field evaluated at three different locations along a typical trajectory of Cassini Grand Finale orbits. The solid traces show the Green's function with rD = 0.75RS, while the dashed traces show the Green's function with rD = 0.6993RS. r r r θ with HAHW of 25 degrees, while Bobs is mostly sen- at 4◦ and 47◦ latitude (Fig. 9B). At sitive to B0.75RS high latitude (−60◦) along the trajectory, Bobs is mostly at somewhat lower latitude (−50◦) sensitive to B0.75RS with good sensitivity until −80◦ latitude, while Bobs is around −67◦ with good sen- most sensitive to B0.75RS sitivity until −80◦ and even higher latitude (Fig. 9C). It should be noted that ¯Gr,θ sin θ(cid:48) is always zero at the poles due to the area factor sin θ(cid:48). θ r This forward calculation illustrates that MAG mea- surements along the Cassini Grand Finale trajectory are sensitive to Saturn's magnetic field at depth to very high latitudes (±80◦). However, the Green's function is fairly wide in latitude near the polar region. This in- dicates that although the large-scale magnetic field at high-latitude should be well determined, the small-scale magnetic field beyond 60◦ may not be uniquely deter- mined. 5. SATURN'S INTERNAL MAGNETIC FIELD FROM THE CASSINI GRAND FINALE MAG MEASUREMENTS Now we move on to retrieve Saturn's internal mag- netic field from the Grand Finale MAG measurements. Although the Gauss coefficients are convenient mathe- matical tools to describe the magnetic field outside their source region, the physical quantity is the profile of Sat- urn's internal magnetic field at the dynamo surface and at the planetary surface. If there exist spatially local- ized features in the magnetic field near the spacecraft trajectory (e.g. a magnetic spot or a latitudinal flux band near the equator), the physical magnetic features could be well resolved by the MAG measurements yet the Gauss coefficients needed to represent the features might be uncertain and non-unique. This is because the Gauss coefficients are defined with respect to global functions which also depend on the field elsewhere on the globe. Thus, uncertainties and uniqueness of the so- 120180240300360Saturn East LongtitudeZ(Mag Eq) [km]P=10h31m32sP=10h34m14s-10-5051015P=10h49m30s-10-5051015-10-5051015060120180240300360Saturn East Longtitude060120180240300360Saturn East Longtitude060306090Latitude [deg]-0.3-0.2-0.100.10.20.30.40.50.60.7Magnetic Green's functionrSC=1.055 RS, latSC=-5.5°-0.2-0.100.10.20.30.4rSC=1.18 RS, latSC=+30°-0.1-0.0500.050.10.150.2rSC=1.56 RS, latSC=-60°-90-60-300306090Latitude [deg]-90-60-300306090Latitude [deg]-90-60-300 10 Cao et al. lution should be evaluated in real space (e.g. evaluating the uncertainties and uniqueness in retrieved Br at the dynamo surface) rather than in the Gauss coefficients space, in particular when there is incomplete or uneven spatial coverage. In addition to the internal magnetic field generated by the MHD dynamo process in the deep interior, three categories of physical sources contribute to the MAG measurements along the spacecraft trajectory: magne- tospheric currents (e.g. magnetodisk, magnetopause, magnetotail currents, and FACs), ionospheric currents, and electromagnetic induction response from the interior of Saturn. Along the close-in part of the trajectory (e.g. r < 2.5RS), magnetospheric contributions other than those from the adjacent FACs would appear as an ex- ternal field and can easily be separated from the internal field given their different radial dependence. Moreover, existing analytical formulas for the magnetodisk field (Connerney et al. 1983; Giampieri & Dougherty 2004) allow a physics-based modeling. The magnetodisk field can be well approximated by a uniform BZ field around 12 nT (Bunce et al. 2007) along the closest part of the Grand Finale orbits. The ionospheric contributions, however, will appear as "internal" field in the MAG measurements since the main conducting layer of the ionosphere, estimated to be ∼ 1100 km above the 1-bar level (Muller-Wodarg et al. 2006), lies below the trajectory of the Cassini Grand Finale orbits. Given the highly variable nature of Saturn's ionosphere from radio occultation and in-situ measurements (Kliore et al. 2014; Wahlund et al. 2018; Persoon et al. 2019), we do not expect the ionospheric contributed magnetic field to be stable with time, which provides one way of separating ionospheric contributions from deep dynamo contributions. In addition, we have made explicit estimations of the amplitude and profile of ionospheric contributed magnetic field at the top of the ionosphere and along the Cassini trajectory (see Ap- pendix C). We found that their biggest contribution is to the axial dipole, which could amount to 6 nT . Their contributions to Gauss coefficients beyond degree-3 are expected to be less than 2.5 nT (see Table 7 in Ap- pendix C). Their impact on determining the deep dy- namo magnetic field of Saturn can thus be explicitly assessed. The magnetospheric and ionospheric field, in particular their time variations, will induce additional internal magnetic field by setting up eddy currents in the conducting layer inside Saturn. For a time-varying signal with frequency close to the rotational frequency of Saturn or the orbital frequency of the Cassini Grand Finale orbits, the induction response will occur around 0.86 RS given our current understanding of Saturn's in- terior electrical conductivity profile (Liu et al. 2008; Cao & Stevenson 2017a; Dougherty et al. 2018). We will present our search for the induced internal field from the time-varying magnetodisk field in section 6. We first average the original 32 Hz MAG measure- ments using a 10-sec window, keeping in mind that the raw attitude information from Star Trackers or gyro- scopes were obtained once every 4 seconds. The contri- butions from the magnetodisk current are then deter- mined orbit-by-orbit with the analytical formula given in Giampieri & Dougherty (2004) as the basis func- tion. The determination of the magnetodisk field uti- lizes only MAG measurements with total field strength between 400 nT and 10000 nT , corresponding approx- imately to radial distance between 1.5 RS and 3.8 RS. These measurements are less affected by the determina- tion of the small-scale internal magnetic field, thus offer- ing better separation of internal and external magnetic field. Furthermore, only field amplitude were employed to derive the magnetodisk field, reducing the effects of high-latitude field aligned currents. Table 3 lists the pa- rameters of the magnetodisk field for each Grand Finale orbit, from a non-linear least-square fitting procedure based on the Levenberg-Marquardt method (Levenberg 1944; Marquardt 1963). The value of magnetodisk field at the equator of Saturn, BZ, along each orbit is listed in Table 3 as well. It can be seen that the magnetodisk BZ field varied between 12 nT and 15.4 nT along the Grand Finale orbits. 5.1. Inversion of Saturn's axisymmetric internal magnetic field with Gauss coefficients representation After removal of the magnetodisk field, we solve for Saturn's axisymmetric internal magnetic field with the traditional Gauss coefficients representation first. Since we are only seeking an axisymmetric internal field solu- tion at this step, which has zero contribution to the az- imuthal field Bφ, only (Br, Bθ) from the measurements were adopted. Excluding Bφ has no effect on the model solutions but does affect the values of the reported RMS residual. We tested two different data selection (DS) criteria: 1) only selecting measurements with B > 10000nT , which approximately corresponds to r < 1.5RS along the Grand Finale orbits; 2) selecting all measurements with r < 3RS, which approximately corresponds to B > 1274nT . Criterion 1 avoids measurements dur- ing the crossing of the high latitude FACs (Dougherty et al. 2018) whilst criterion 2 extends the data to the maximum latitude coverage. 5.1.1. Un-regularized inversion Saturn's internal magnetic field from Cassini Grand Finale 11 Table 3. Parameters of the magnetodisk field and the corre- sponding surface BZ along the Cassini Grand Finale orbits. Here a and b are the radial distance of the inner and outer edge of the magnetodisk from the center of Saturn respec- tively, D is the vertical half thickness of the magnetodisk, and µ0I is the surface current amplitude, see Connerney et al. (1983); Giampieri & Dougherty (2004); Bunce et al. (2007) for more details. In our analysis, only µ0I were var- ied while a, b, and D were fixed, due to the insensitivity of the MAG measurements inside 3 RS to the later three parameters. Rev Num a 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 [RS] 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 6.5 b [RS] 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 µ0I [nT ] 48.1 47.8 57.4 49.2 60.9 53.8 48.2 54.8 51.2 47.7 57.0 51.3 52.7 51.0 56.5 56.9 55.3 57.5 60.5 59.3 56.4 57.2 47.6 D Surface BZ [RS] 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 [nT ] 12.2 12.1 14.5 12.4 15.4 13.6 12.2 13.9 12.9 12.1 14.4 13.0 13.3 12.9 14.3 14.4 14.0 14.5 15.3 15.0 14.2 14.5 12.0 The forward linear problem can be formulated as data = G model, (6) in which data represents MAG measurements with the magnetodisk field removed, model represents the Gauss coefficients, and G represents the matrix expression of equation (1). In un-regularized inversion, we seek to minimize the data-model difference data − G model2 , (7) without placing explicit constraints on the behavior of the model. We monotonically increase the maximum spherical harmonic (SH) degree, nmax, of the axisymmetric inter- nal field model and examine the behavior of the data- model fit. Both the RMS residual and the vector resid- ual at each data point are evaluated. This exercise aims at revealing the minimum spectral content required by the measurements. Table 4 lists the Gauss coefficients from the un- regularized inversion with the two different data selec- tion criteria, while Fig. 10 shows the RMS residual. It can be seen that although the RMS residuals cor- responding to the two different data selection criteria behave slightly differently, the resulted model solutions from the two data selection criteria are almost identi- cal. This indicates the FACs do not have a significant impact on the internal field modeling given the Grand Finale trajectory. Table 4 also shows that the Gauss coefficients beyond degree 3 are on the order of 100 nT or less, significantly smaller than those of degrees 1 - 3. Table 4. Gauss coefficients of the un-regularized inversion of Saturn's axisymmetric internal magnetic field with two different data selection (DS) criteria. DS 1 21120 1522 2218 DS 2 21127 1527 2223 DS 1 21156 1591 2300 116 77 49 nmax = 3 nmax = 3 nmax = 6 nmax = 6 nmax = 9 nmax = 9 DS 2 21139 1576 2255 77 −9 −8 −100 −39 −54 DS 1 21139 1578 2255 82 −9 −3 −100 −36 −55 DS 2 21150 1586 2291 108 71 45 [nT ] g0 1 g0 2 g0 3 g0 4 g0 5 g0 6 g0 7 g0 8 g0 9 Figure 10. Root-mean-square (RMS) residual from the un-regularized axisymmetric inversion. Only (Br, Bθ) were adopted in this analysis. The two different traces represent two different data selection criteria. 1234567891011121314151617181920Maximum Spherical Harmonic Degree100101102103RMS Residual [nT]B>10000 nTr<3 RS, B>1274 nT 12 Cao et al. The RMS residual in the un-regularized inversion de- creases monotonically with the maximum SH degree, with a few distinct features: 1) the RMS residual drops by more than an order of magnitude from nmax = 2 to nmax = 3, 2) the RMS residual remains roughly con- stant (∼ 10 nT ) between nmax = 6 and nmax = 8, 3) the RMS residual decreases by more than a factor of two from nmax = 8 to nmax = 9. Fig. 11 shows the vector residuals as a function of time from periapsis along the S/C trajectory for Rev 283 to Rev 292, with the contribution from the mean magnetodisk field being over-plotted (thick black dashed lines). The behavior along all other orbits are quite sim- ilar. It can be seen that the vector residuals from the un-regularized degree-3 model feature larger amplitude and larger spatial-scale in the northern hemisphere while the vector residuals from the un-regularized degree-6 model features mostly north-south symmetric oscilla- tions. The residuals from the un-regularized degree-9 model are broadly consistent with the average magne- todisk field, except within [-20, +10] minutes around the periapsis. Given that the un-regularized degree-9 model fits the measurements reasonably well except very close to the periapsis, why not take it as a new basis solution of Sat- urn's internal magnetic field? To answer this question, we examine the magnetic perturbations associated with Gauss coefficients above degree 3 at the a = 0.75 RS, c = 0.6993 RS isobaric ellipsoidal surface. As shown in Fig. 12, when evaluated at the a = 0.75 RS iso- baric surface, ∆Br associated with the degree 4 - 9 co- efficients of the un-regularized degree-9 model features 3.75 times higher values above 60◦ latitude compared to those within ± 60◦ latitude. Moreover, the frac- tional amplitude of the small-scale field perturbations ∆Br(n > 3)/B(n ≤ 3) above 60◦ are about 2.5 times larger than that within ±60◦. Given that the Cassini spacecraft did not go much beyond ±60◦ latitude dur- ing the Grand Finale phase, the model field behavior beyond ±60◦ latitude is likely to be neither justified nor uniquely determined by the measurements. Thus, we turn to the regularized inversion technique (Holme & Bloxham 1996; Gubbins 2004) to construct internal field models for Saturn that not only fit the Cassini mea- surements but are also well-behaved. Here, we define "well-behaved" in the sense that the fractional ampli- tude of the small-scale field perturbations beyond 60◦ are similar to that within ± 60◦. This definition of "well-behaved" is a subjective choice, but it is a rea- sonable one given the available measurements. 5.1.2. Regularized inversion In regularized inversion, in addition to seeking models that fit the data, constraints are placed on the behavior and properties of the model. This can be formulated as minimizing data − G model2 + γ2 L model2 , (8) face integrated power in the radial flux,(cid:82) B2 here γ is a tunable damping parameter controlling the relative importance of model constraints and data-model misfit, while L represents the particular form of con- straint on the model. Here we seek to minimize the sur- r (n > 3)dΩ, at r = 0.6993RS. Since we expect the regularization to mainly constrain the behavior of the magnetic field above ±60◦ latitude, we set the regularization radius to 0.6993 RS, the polar radius of the a = 0.75 RS isobaric surface. Thus, the model constraint is (cid:18) Rp (cid:19)n+2 (9) L = n + 1√ 2n + 1 rdamp for n > 3 and L = 0 for n ≤ 3, in which Rp is the radius of the planet and rdamp is the damping radius at which the constraints are placed. Here, Rp = RS, and rdamp = 0.6993 RS. Fig. 13 displays the Gauss coefficients and ∆Br(n > 3)/B(n ≤ 3) at the a = 0.75 RS, c = 0.6993 RS el- lipsoidal surface from a survey of regularized inversion with different damping parameters. The preferred so- lution is highlighted using thick red traces in both pan- els. Compared to the un-regularized degree-9 model, our preferred solution features ∆Br/B with similar ampli- tude beyond ±60◦ and within ± 60◦. Moreover, Fig. 13 shows that the model Br are broadly similar within ± 60◦. This preferred solution constructed from the entire Grand Finale dataset is very similar to the Cassini 11 model (Dougherty et al. 2018) derived from 9 of the first 10 Grand Finale orbits in the profile of Br and in the Gauss coefficients (see Table 5 for the Gauss coeffi- cients). We refer to this newly constructed model as the Cassini 11+ model. 5.2. Inversion of Saturn's internal magnetic field with Green's function 5.2.1. The eigenvectors of the inverse problem formulated with Green's function In addition to the traditional Gauss coefficients repre- sentation, the inverse problem for the internal magnetic field can be formulated with Green's function represen- tation. In this formulation, the model in data = G model (10) Saturn's internal magnetic field from Cassini Grand Finale 13 Figure 11. Component residuals, (∆Br, ∆Bθ), from the un-regularized degree 3, degree 6, and degree 9 models along Rev 283 to Rev 292 within ± 4 hours of the periapsis. In each panel, thick black dashed line represents contribution from the mean magnetodisk field. Table 5. Gauss Coefficients of newly derived Cassini 11+ model compared to that of the Cassini 11 model (Dougherty et al. 2018) 21140 1581 2260 91 12.6 17.2 −59.6 −10.5 −12.9 15 18 [nT ] Cassini 11 Cassini 11+ 21141 1583 2262 95 10.3 17.4 −68.8 −15.5 −24.2 9.0 11.3 −2.8 −2.4 −0.8 g0 1 g0 2 g0 3 g0 4 g0 5 g0 6 g0 7 g0 8 g0 9 g0 10 g0 11 g0 12 g0 13 g0 14 is the profile of Br at the dynamo surface, and G is the matrix expression of equation (2). For simplicity, we choose Br at rd = 0.6993 RS, same as the damping radius in our regularized inversion, as the model here. Each eigenvector of the inverse problem is a profile of axisymmetric Brd r as a function of latitude, which we de- note as Brd , here i is the order of the eigenvector. Here i we emphasize that the eigenvectors here are not stan- dard predetermined functions but depend on the specific trajectory of the measurements. The final solution is a weighted sum of the eigenvectors of different order Brd r = (cid:88) βiBrd i , i = 1, 2, ... (11) i here βi are the weights of the eigenvector. Both βi and Bi can be computed with the singular-value- decomposition (SVD) (e.g. Jackson 1972; Connerney 1981; Aster et al. 2013, also see Appendix B). We choose the Gauss-Legendre quadrature points with 180 grids in the latitudinal direction to ensure high- precision integration for smooth functions. In Fig. 14, we show the first six eigenvectors in parameter space de- rived along the trajectory of the Cassini Grand Finale orbits. It can be seen that all eigenvectors feature zero Brd r at the poles, in contrast to the m = 0 associated Legendre functions (the basis functions for axisymmet- ric Gauss coefficients) which all peak at the poles. It becomes immediately clear that with the given trajec- tory, the Green's function method seeks solutions with 283284285286287288289290291292-100-50050100 BrCassini Measurements - Un-regularized Degree-3 Model B-30-20-100102030-4-3-2-1Hours from Periapsis-40-20020406012340-4-3-2-1Hours from Periapsis12340Cassini Measurements - Un-regularized Degree-6 ModelCassini Measurements - Un-regularized Degree-9 Model-100-50050100-60-40-200204060-60-40-200204060 B B Br Br 14 Cao et al. Figure 12. Profile of the small-scale (n > 3) axisymmetric magnetic field ∆Br and ∆Br(n > 3)/B(n ≤ 3) at the a = 0.75 RS, c = 0.6993 RS isobaric surface according to the un-regularized degree-9 model. It can be seen that in this un-regularized model, ∆Br above ±60◦ latitude are about 3.75 times larger than that within ±60◦, and ∆Br/B above ±60◦ are about 2.5 times larger than that within ±60◦. zero Br at the poles, which is an intriguing mathemat- ical property of this method. Given this property and the fact that Saturn's internal magnetic field is predom- inantly dipolar, we employ the Green's function method to seek small-scale internal magnetic field solutions be- yond spherical harmonic degree 3. 5.2.2. Small-scale features in Saturn's internal magnetic field from Green's function inversion We adopt the degree 1 to 3 Gauss coefficients from the Cassini 11 model as the basis model, and seek the internal magnetic field beyond this basis model using the Green's function. To obtain a smooth solution, one needs to either truncate the solution at a certain order Figure 13. Gauss coefficients and ∆Br(n > 3)/B(n ≤ 3) at the a = 0.75 RS, c = 0.6993 RS isobaric surface from a survey of regularized inversion based on Cassini Grand Finale MAG measurements. The thick red traces represent our preferred solution, the Cassini 11+ model. imax (see Appendix B for more details) or apply certain form of regularization. Here we choose to truncate the solution at imax as a first step. The truncation order of the eigenfunction, imax, is determined by the RMS residual and the model-data misfit. Fig. 15 shows the small scale magnetic field beyond spherical harmonic degree 3, ∆Br, constructed from the Green's function with rd = 0.6993 RS and imax = 12, which we refer to as CG12 model, in which C stands for Cassini, G stands for Green's function, and 12 in- dicates the truncation order of the eigenfunction. This truncation order is chosen to yield a similar RMS resid- ual to that of the Cassini 11+ model. The perturbation field from the Cassini 11+ model and the Cassini 11 model are shown in Fig. 15 for comparison (the same -6-4-20246 Br @ a=0.75 RS [nT]−80−60−40−20020406080Latitude [deg]Un−regularized degree−9 model104-0.3-0.2-0.100.10.20.3 Br/B @ a=0.75 RS−80−60−40−20020406080Latitude [deg]AB1011121314Spherical Harmonic Degree10-1100101102103104105=310-3,RMS=4.58 nT=110-3,RMS=4.66 nT=310-2,RMS=4.73 nT=0.1,RMS=5.32 nT=0.2,RMS=7.64 nTgn0 [nT]−80−60−40−20020406080Latitude [deg]B-0.3-0.2-0.100.10.20.3 Br/B @ a=0.75 RSA567891234 Saturn's internal magnetic field from Cassini Grand Finale 15 this nonetheless highlights the non-uniqueness in the so- lution beyond ±60◦ latitude. This non-uniqueness in the polar region should be kept in mind when interpreting the resultant ∆Br. (cid:18) rd RP g0 n = 2n + 1 2(n + 1) (cid:19)n+2(cid:90) π 0 Once we obtain Br at r = rd, the corresponding Gauss coefficients can be easily computed via a surface integra- tion given the orthogonality of the spherical harmonics on a sphere. BrP 0 n(cos θ) sin θdθ, results the pre-factor (12) where from the Schmidt- normalization. Supplementary Table 1 compares the Gauss coefficients of the Green's function model (the CG12 model) to that of the Cassini 11 model (Dougherty et al. 2018) and the Cassini 11+ model. For the CG12 model, the degree 1-3 Gauss coefficients are the sum of the basis model and those computed from Eq. (12). It can be seen that the Gauss coefficients of these models are also broadly similar: beyond degree 3, all models fea- ture a strong and positive g0 4 and a strong and negative g0 7. 6. ELECTROMAGNETIC INDUCTION RESPONSE FROM SATURN'S INTERIOR Electromagnetic (EM) induction can be employed to probe the interiors of planetary bodies. Examples of planetary applications of this technique include the dis- covery of the subsurface ocean inside Europa and Cal- listo from Galileo magnetometer measurements (Khu- rana et al. 1998), constraints on lunar core size from Apollo 12 and Explorer 35 magnetometer measurements (Hood et al. 1982), and constraints on water content variations in the mantle transition zone inside the Earth (Kelbert et al. 2009). depth, d = (cid:112)2/ωindµ0σ, which depends on the fre- The key parameter in the EM induction is the skin- quency of the inducing field ωind and the local electrical conductivity σ. µ0 is the magnetic permeability. Since the electrical conductivity is expected to rise continu- ously yet rapidly as a function of depth inside Saturn (Weir et al. 1996; Liu et al. 2008; Cao & Stevenson 2017a), the EM induction response is expected to oc- cur at different depths for inducing fields with different frequencies. The depth at which the EM induction oc- curs is where the frequency dependent skin-depth dind becomes comparable to or smaller than the local scale- height of the electrical conductivity Hσ =(cid:12)(cid:12)σ/ dσ (cid:12)(cid:12). Given our current understanding of the electrical conductiv- ity profile inside Saturn based on a band-closure model (Liu et al. 2008), EM induction is expected to occur at rind around 0.87RS and 0.86RS for sounding frequencies dr Figure 14. First six eigenvectors of the magnetic Green's function at r =0.6993 RS (the polar radius of the a = 0.75 RS, c = 0.6993 RS ellipsoidal surface). It can been seen that the eigenfunctions constructed from the Green's func- tion feature zero values at the poles, in contrast to the m = 0 Legendre functions which peak at the poles. degree-3 model has been removed for a fair comparison). It can be seen from Fig. 15 that the field structures constructed from two different methods are very similar within ±60 degrees: there are four latitudinal magnetic field bands between the equator and 60◦ latitude in each hemisphere. Above ±60◦, the solution from the Green's function features zero Br at the poles (an intrinsic prop- erty of the method) while the Cassini 11+ model fea- tures comparable ∆Br/B to that within ± 60◦ (which results from the chosen regularization). Although the difference between the two models beyond ±60◦ latitude originates from the intrinsic properties of the methods, -80-60-40-20Latitude at r=0.6993 RS [deg]-0.2-0.15-0.1-0.0500.050.10.150.2Bi(')i=1i=2i=3204060800i=4i=5i=6-80-60-40-20Latitude at r=0.6993 RS [deg]204060800-0.2-0.15-0.1-0.0500.050.10.150.2Bi(')AB 16 Cao et al. Figure 15. Small-scale (n > 3) magnetic field of Saturn viewed at the a = 0.75 RS, c = 0.6993 RS isobaric surface constructed from regularized Gauss coefficients inversion (Cassini 11+ model) and from the Green's function inversion (CG 12 model). equal to the rotational frequency of Saturn (∼ 10.5 hr) and the orbital frequency of Cassini Grand Finale orbits (6.5 Earth days) respectively (Fig. 16A). The electrical conductivity at these depths are about 0.1 S/m and 1 S/m respectively. The depth from the 1-bar atmosphere is about 8000 km. θ r r This corresponds to an induction response in which the induced radial field Bind perfectly cancels the radial component of the external inducing field Bext at rind. Note that the induced tangential component Bind acts to increase the external tangential component by 50% instead of canceling it at rind. The factor 1/2 in Eq. 13 originates from the normalization of the associated Leg- endre polynomials which is part of the definition of g0 1. Thus, the slope of ∆g0 1 versus ∆BZ reveals the depth at which the induction response occurs. For an induc- tion depth at 0.87RS (0.86RS), the expected slope is −0.4773 (−0.4755). We solve for ∆g0 1 orbit by orbit after removing the Cassini 11+ model and the magnetodisk field. Figure 16B shows ∆g0 1 as a function of the time-varying mag- netodisk ∆BZ field orbit-by-orbit. With the available data an induction signal seems present. If one performs a formal inversion analysis on this dataset, the expected slope is within 1σ of that from the formal inversion anal- ysis. However, the large scatter in the data precludes any definitive constraint on the induction depth. The magnetodisk BZ field (Table 3) is expect to in- duce an internal axial dipole g0 1(ind) inside Saturn. This induction response consists of two parts, a time- stationary part and a time-varying part. The magne- todisk field has a well defined mean component of order 10 nT , which seems to be stable over at least decadal time-scales with available in-situ observations. Given the very high electrical conductivity in Saturn's deep interior, an induction response to the stable part of the magnetodisk BZ field is expected. However, this in- duction response cannot be effectively separated from a stable internal axial dipole. Thus, in searching for an induction response from the interior of Saturn, we focus on the expected time-varying part. The expected time-varying induction response ∆g0 1 to the time-varying part of the magnetodisk field ∆BZ is that (cid:18) rind (cid:19)1/3 ∆g0 1 = − 1 2 RS ∆BZ. (13) 7. ORBIT-TO-ORBIT VARIATIONS IN SATURN'S "INTERNAL" QUADRUPOLE MAGNETIC MOMENTS Cassini 11CG 12Cassini 11+ Br @ a=0.75 RS [nT]204060-40-200-60Latitude [deg]607590-90-75-60-1-0.8-0.6-0.4-0.200.20.40.60.81104104104-2-1.5-1-0.500.511.52-2-1.5-1-0.500.511.52 Saturn's internal magnetic field from Cassini Grand Finale 17 2 stay within ± 4.6 It can be seen that the variations in g0 nT , except along Rev 288 where a factor of 1.5 larger variation in g0 2 were observed. Near simultaneous Hub- ble Space Telescope (HST) observations of the northern far-ultraviolet aurorae of Saturn recorded a strong inten- sification of total auroral power in the H2 bands close to the periapsis time of Rev 288 (Lamy et al. 2018). Figure 16. Electromagnetic induction response from the interior of Saturn. Panel A shows the skin depth versus the electrical conductivity scale-height. It can be seen that for inducing field with frequencies between the spin frequency of Saturn and the orbital frequency of the Cassini Grand Finale orbits, the skin depth becomes comparable to or smaller than the local conductivity scale height around 0.86 RS. Panel B shows the orbit-to-orbit varying internal dipole ∆g0 1 as a function of the orbit-to-orbit varying magnetodisk field ∆BZ derived from the Cassini Grand Finale MAG measurements. The expected induction response from an induction depth at 0.86 RS is overplotted. In addition to solving for ∆g0 1 orbit by orbit, we also attempted to solve for ∆g0 2 orbit by orbit and found some non-negligible variations. Solving for ∆g0 2 does im- prove the data-model misfits, while solving for ∆g0 n with n > 2 does not reduce the data-model misfit much fur- ther. We attempted to solve for ∆g0 2 separately and simultaneously, and observed negligible differences in the resulting values. Table 6 lists the resultant ∆g0 2, which are also plotted against Rev Number in Fig. 17. 1 and ∆g0 Figure 17. Orbit-to-orbit variations in Saturn's exter- nal magnetodisk field, "internal" dipole, and "internal" quadrupole coefficients. 1 and ∆g0 Moreover, ∆g0 1. The standard deviation of ∆g0 2 do not exhibit strong corre- lation: the coefficients of correlation between the two is only 34%. The variability in ∆g0 2 is larger than that in ∆g0 2 is 2.8 nT (2.4 nT if Rev 288 is excluded), while the standard devia- tion of ∆g0 1 is 2.0 nT . We speculate that the observed variations in g0 2 mostly reflect variations in the east-west (zonal) currents in the ionosphere. The quadrupole mo- ment g0 2 corresponds to north-south antisymmetric zonal currents: e.g. a positive g0 2 is consistent with eastward current in the north and westward current in the south. The order 5 nT amplitude is consistent with our order- of-magnitude estimations of the ionospheric Hall current contributions (see Appendix C), while the pattern indi- cates stronger north-south asymmetry compared to the 0.650.70.750.80.850.910−1100101102103104105Radial Distance [RS]Skin depth versus conductivity scale height [km] Hσd(T=10.5 hours)d(T=6.5 days) BZ [nT] g10 [nT]AB-2-1.5-1-0.5200.511.5-5-4-3-2-10123451213141516Magnetodisk BZ [nT]-5-2.502.55 g10 [nT]-8-6-4-202468 g20 [nT]271272273274275276277278279280281282283284285286287288289290291292Rev Num 18 Cao et al. Table 6. Orbit-to-orbit varying Internal Dipole and Quadrupole Coefficients Measured along the Cassini Grand Finale Orbits Rev Num ∆g0 271 272 273 274 275 276 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 1[nT ] ∆g0 1.2 3.2 1.4 -0.5 2.2 -0.7 1.1 -2.7 4.2 -3.7 1.6 -1.3 -0.3 1.1 -1.7 -1.2 2.1 -2.1 -0.5 0.6 -1.8 2[nT ] 1.1 1.9 -1.3 -0.8 -0.5 3.5 -1.2 2.1 0.1 -4.0 1.1 -4.6 2.3 -0.4 -2.1 -3.4 7.0 0.4 -4.0 2.0 1.8 expectation of continuing the 1-bar wind pattern up to the 1100 km altitude ionospheric layer. 8. SEARCH FOR NON-AXISYMMETRY IN SATURN'S INTERNAL MAGNETIC FIELD As demonstrated in the analysis of Saturn's mag- netic equator positions (section 3), the level of depar- ture from perfect axisymmetry is likely only on the or- der of 3 × 10−4. Nonetheless, we performed a search for the non-axisymmetric internal magnetic moments of Saturn based on the Cassini Grand Finale MAG mea- surements. The traditional Gauss coefficients represen- tation is adopted, and the maximum SH degree and or- der for the non-axisymmetric moments are both set to be 3. Since the deep interior rotation rate of Saturn re- mains uncertain (Anderson & Schubert 2007; Read et al. 2009; Mankovich et al. 2019; Militzer et al. 2019), we surveyed a wide range of possible rotation periods from 10h30m00s to 10h55m00s. Fig. 18 shows the dipole tilt, the relative non- axisymmetry in SH degree 2 and 3 (defined as the ratio of the amplitude of the non-axisymmetric magnetic mo- ments to that of the axisymmetric magnetic moment of the same degree), and the RMS residual from the search. No dominant peak in the amplitude of the in- Figure 18. Results from the search for non-axisymmetry in Saturn's internal magnetic field based on the Cassini Grand Finale MAG measurements. Panel A shows the dipole tilt, panel B and C show the relative non-axisymmetry in degree 2 and degree 3 moments respectively, and Panel D shows the RMS residual. All quantities are shown as a function of the assumed rotation period of Saturn's deep interior. No dom- inant peak in internal non-axisymmetry can be identified, and the peak dipole tilt is less than 0.007◦ (25.2 arcsecs). ternal non-axisymmetric can be identified, and the peak dipole tilt is less than 0.007 degrees (25.2 arcsecs). The relative non-axisymmetry in degree 2 and 3 are less than 1.5×10−3. Thus, Saturn's internal magnetic field is 1000 times more axisymmetric compared to those of Earth and Jupiter. What makes Saturn's internal magnetic field so drastically different? We discuss this in the next section. 9. IMPLICATION FOR SATURN'S INTERIOR 9.1. Magnetic axisymmetry and deep stable stratification inside Saturn The exceptional level of axisymmetry in Saturn's in- ternal magnetic field revealed by the Cassini Grand Fi- nale MAG measurements presents a challenge and an opportunity. The challenge is to our understanding of natural dynamos while the opportunity is to decode Sat- urn's interior structure and dynamics. Cowling's the- orem (Cowling 1933; Backus & Chandrasekhar 1956; 10:30:0010:35:0010:40:0010:45:0010:50:0010:55:0000.0020.0040.0060.008Dipole tilt [deg]10:30:0010:35:0010:40:0010:45:0010:50:0010:55:0000.511.5 SH degree 210-310:30:0010:35:0010:40:0010:45:0010:50:0010:55:0000.511.5Relative non-axisym SH degree 310-310:30:0010:35:0010:40:0010:45:0010:50:0010:55:00Saturn Rotation Period [HH:MM:SS]8910RMS Residual [nT] Saturn's internal magnetic field from Cassini Grand Finale 19 Hide & Palmer 1982) precludes a perfectly axisymmet- ric magnetic field to be maintained by natural dynamos, although no lower bound on the departure from axisym- metry has been placed by this theorem. Furthermore, Cowling's theorem is a statement about the entire mag- netic field in the dynamo region, much of which we cannot observe (e.g., the toroidal field). Setting Cowl- ing's theorem aside for now, Saturn's axisymmetric in- ternal magnetic field appears special from the perspec- tives of both observations and modern understanding of the planetary dynamo process. From observations, highly axisymmetric magnetic fields are rare among planets. Both Earth and Jupiter feature ∼ 10◦ dipole tilt, while Uranus and Neptune feature ∼ 50◦ dipole tilt and strong non-axisymmetric quadrupole and octopole fields. The case of Mercury and Ganymede are less clear at this stage. Mercury's magnetic equator positions do feature ∼ 100 km peak- to-peak variations (see Fig. 4 in Anderson et al. 2012), which are much bigger variations compared to that of Saturn given the relative small size of Mercury (RM ercury = 2439.7km). However, whether such vari- ations are due to internal non-axisymmetry or mag- netospheric processes (Jia et al. 2015) remains to be clarified. The ESA-JAXA BepiColombo mission is ex- pected to help resolve this issue. The non-axisymmetry of Ganymede's internal magnetic field is less clear due to the ambiguity in separation of the dynamo-generated in- ternal field and the EM induced field given the limited spatial-temporal coverage of Galileo Ganymede flybys (Kivelson et al. 2002). The ESA JUpiter ICy moons Explorer (JUICE) mission is expected to resolve this ambiguity with low-altitude Ganymede orbits. From modern understanding of the planetary dynamo process, highly axisymmetric magnetic fields are rare in convective dynamo simulations. Highly supercritical ro- tating convection is strongly non-axisymmetric. Due to inverse cascade (Guervilly et al. 2014; Rubio et al. 2014), the non-axisymmetry in the convective flows tends to have strong large-scale components. These large-scale non-axisymmetric convective flows are expected to gen- erate large-scale non-axisymmetric magnetic fields as observed in the majority of convective numerical dy- namo simulations. In numerical dynamo surveys, the magnetic field in the dipolar branch tends to feature a modest amount of non-axisymmetry, e.g. with dipole tilt between 5 to 10 degrees, while the magnetic field in the multi-polar branch tends to be dominated by non- axisymmetry (Christensen & Aubert 2006; Soderlund et al. 2012; Duarte et al. 2013). The most appealing mechanism to axisymmetrize Sat- urn's internal magnetic field is via the combination of strong differential rotation and suppression of large-scale non-axisymmetric convective motion on top of the dy- namo region (Stevenson 1980, 1982). It should be em- phasized that the differential rotation here refers to the shear between the flow in the convective dynamo re- gion and the flow in an electrically conducting layer above the convective dynamo region. In principle, only differential rotation in the spherical radial direction is needed. Such differential rotation tends to destroy non- axisymmetric magnetic features via advectively shearing them, then diffusively dissolving them. Under the case of angular velocity as a function of radial distance only and ignoring the dynamic feedback from the Lorentz force induced, this process can be thought of as elec- tromagnetic filtering. In addition to strong differen- tial rotation on top of the deep dynamo, suppression of large-scale non-axisymmetric convective motion out- side the deep dynamo is a necessary ingredient to main- tain an axisymmetric magnetic field, since any large- scale non-axisymmetric convective motion in an electri- cally conducting region would lead to large-scale non- axisymmetric magnetic field. The most likely way these two conditions are satisfied inside Saturn is via the for- mation of a stably stratified (Stevenson 1980) or double diffusively convecting (Leconte & Chabrier 2012, 2013) layer on top of the deep fully convective dynamo. He- lium rain (Stevenson 1975; Stevenson & Salpeter 1977; Morales et al. 2009; Lorenzen et al. 2009) could lead to the formation of such a layer. However, the picture of helium rain inside Saturn is in doubt since we lack a direct measurement of significant helium depletion in the atmosphere of Saturn. The established helium de- pletion in Jupiter from Galileo results and the expected lower entropy in Saturn suggests helium rain should oc- cur in Saturn to a greater extent than in Jupiter but this is contingent on the standard assumption of isentropy down to the pressure level of helium insolubility in both planets. Other processes inside Saturn could lead to the formation of such a layer on top of the dynamo. For example, if dissolved core material (heavy elements) is convectively mixed upward to around 0.6 RS, this would create a stable compositional gradient near this depth since the layer above would feature less heavy elements. The thickness of this layer and the format of radial motion in this layer, e.g. oscillatory motion or small- scale double diffusive convective motion, is determined by the competition between the thermal gradient and the compositional gradient (Leconte & Chabrier 2012). The measured extreme level of axisymmetry in Saturn's magnetic field can help us constrain these properties. We loosely refer to this layer as a "stable layer" even 20 Cao et al. though it should be understood that this layer could be double diffusively convecting. An important non-dimensional parameter to quantify the stable layer's ability to axisymmetrize the dynamo generated magnetic field is αRm = mLStable RDynamo ∆uφLStable ηStable , (14) here m is the azimuthal wave number (spherical har- monic order m), LStable is the thickness of the stable layer, RDynamo is the radius of the deep dynamo, ∆uφ is the differential rotation between the stable layer and the deep dynamo, and ηStable is the magnetic diffusivity of the stable layer. Fig. 19 shows the maximum attenu- ation factor of the dipole tilt (m = 1), which is the ratio of the dipole tilt above the stable layer to that below the stable layer, as a function of αRm according to the plane layer kinematic model of Stevenson (1982): ∆max = 1.59 (αRm)1/12 exp (cid:104)− √ 2/3 (αRm)1/2(cid:105) . (15) Figure 19. The attenuation factor of the internal dipole tilt as a function of αRm according to the kinematic plane-layer model by Stevenson (1982). To reach a 0.007◦ dipole tilt, αRm needs to be larger than 238. The stable layer needs to be thicker than 2500 km (5600 km) if the differential rotation between the deep dynamo and the stable layer is about 5 mm/s (1 mm/s). Assuming a 10◦ dipole tilt in the deep dynamo re- gion, to achieve the observed upper limit of dipole tilt, 0.007◦, outside the stable layer, αRm needs to be larger than 238. If we assume 1 mm/s (5 mm/s) differential rotation between the stable layer and the deep dynamo and a magnetic diffusivity of 4 m2/s (equivalent to an electrical conductivity of 2 × 105 S/m) and a deep dy- namo radius around 0.55 RS, this requires a stable layer thicker than 5600 km (2500 km). It should be immedi- ately realized that a "stable" layer over 2500 km thick cannot be a purely diffusive layer. Assuming a thermal conductivity of 100 W/K/m (French et al. 2012), to dif- fusively transport the observed luminosity 2 W/m2 of Saturn through a purely conducting layer over 2500 km thick around 0.55 RS would require a thermal gradient as large as 66 K/km or a temperature jump over 165000 K across the stable layer. Thus, double diffusive con- vection and/or fluid waves must be present to transport the heat out. Moreover, αRm and the "stable" layer thickness de- rived here is likely a lower limit. In this kinematic model (Stevenson 1982), the dynamical feedback from the mag- netic field to the flow via the Lorentz force was ignored. Such dynamical feedback likely would reduce the effi- ciency of axisymmetrization. Whether a very large αRm can be achieved in a fully dynamic situation is unclear, since the differential rotation between the stable layer and the deep dynamo ∆uφ would be dynamically con- strained. In published Saturn dynamo simulations with a stable layer (Christensen & Wicht 2008; Stanley 2010), αRm is on the order of 15 or less, consistent with the ∼ 1◦ dipole tilt achieved. Whether there is a dynami- cal limit on αRm and the axisymmetrization efficiency of this mechanism remains an open question for future investigations. 9.2. Banded magnetic perturbations and deep zonal flows in the semi-conducting layer of Saturn It is intriguing that although Saturn's internal mag- netic field appears to be perfectly axisymmetric, it does feature a rich axisymmetric magnetic spectrum extend- ing to spherical harmonic degree 9 and beyond. The degrees 1 to 3 magnetic moments likely originate from the deep dynamo given their order-of-magnitude power dominance over that of the higher degree moments when viewed at 0.75 RS. The magnetic moments beyond de- gree 3 and the associated latitudinally banded magnetic perturbations likely originate from a shallow secondary dynamo with alternating bands of deep zonal flows in the semi-conducting layer of Saturn. As shown in Cao & Stevenson (2017a), banded differential rotation and local helical motion in the semi-conducting region could generate a rich axisymmetric magnetic spectrum even if the deep dynamo field is simply an axial dipole. The Cassini MAG data suggests that there are eight alter- nating bands of magnetic perturbations between ± 60◦ at the a = 0.75RS elliptical surface (Fig. 15 & 20B). The typical latitudinal width of each magnetic band is 100101102103Rm10-410-310-210-1100Attenuation factor of dipole tilt dip1o tilt0.01o tilt0.001o tilt Saturn's internal magnetic field from Cassini Grand Finale 21 Figure 20. Saturn's large and small scale radial magnetic field at the a = 0.75, c = 0.6993 RS isobaric surface according to the Cassini 11+ model. Saturn's large scale radial magnetic field at this depth features a relatively weak equatorial region, Br remains less than 50,000 nT (<1/3 of its peak value) between ±40◦. Saturn's small-scale magnetic field at this depth features eight alternating bands between ±60◦, with typical amplitude of ∼ 5% - 10% of the background field. ∼ 15◦. If we project the observed 1-bar surface zonal winds along the direction of the spin-axis towards the a = 0.75RS elliptical surface, there are eight alternat- ing bands of zonal jets between ± 60◦ with the off- equatorial jets feature typical latitudinal width ∼ 15◦ at this depth. Thus, the characteristic width of the latitudinally banded magnetic perturbations is similar to that of the Z-projection of the surface off-equatorial zonal jets. layer which produces toroidal magnetic field BT from B0 through the dynamo ω-effect, and 3) local helical mo- tion which produces observable poloidal magnetic field perturbations ∆BP from BT through the dynamo α- effect (Parker 1955; Steenbeck et al. 1966; Steenbeck & Krause 1966). Heat transport requirements and back- ground rotation naturally lead to helical motion and lo- cal dynamo α-effect in the semi-conducting layer. The spatial profile of the resultant BT and ∆BP are ex- pected to be spatially correlated with that of the differ- ential rotation. The fact that the characteristic width of the latitudinally banded magnetic perturbations is sim- ilar to that of the Z-projected surface zonal jets lends Three necessary ingredients for a secondary dynamo in the semi-conducting layer are 1) the existence of a deep dynamo which provides the background magnetic field B0, 2) differential rotation in the semi-conducting 22 Cao et al. further support to the idea that the profile of deep zonal flows in Saturn's semi-conducting layer strongly resem- ble that of the observed surface zonal jets (Iess et al. 2019; Galanti et al. 2019; Militzer et al. 2019). In addi- tion to the idealized mean-field model (Cao & Stevenson 2017a), secondary dynamo action has also been observed in some global numerical dynamo simulations for giant planets featuring a radially varying electrical conductiv- ity and deep zonal flows in the outer layers (e.g. Gastine et al. 2014; Duarte et al. 2018). The peak toroidal magnetic field production could oc- cur anywhere between the top of the semi-conducting layer (e.g. ∼ 0.87RS where σ ∼ 0.1 S/m) and the base of the semiconducting layer (to be defined later), since it is determined by the competition between the decay- ing wind velocity and the increasing electrical conduc- tivity as a function of depth. Regardless of the peak production depth, the toroidal magnetic field will dif- fuse downward to the base of the semi-conducting layer (e.g., see Figs. 2 & 10 in Cao & Stevenson 2017a). The poloidal magnetic field perturbations ∆BP, how- ever, are expected to be generated mainly near the base of the semi-conducting layer, due to its dependence on σ2. The "base of the semi-conducting layer" is defined by either 1) the transition to the main dynamo, which likely occurs before the saturation of the electrical con- ductivity, or 2) the upper end of the "stable layer" which provides a well-defined separation of the shallow dynamo from the deep dynamo. Since the secondary dynamo lies above the "sta- ble layer", will it generate secondary non-axisymmetric magnetic field that violate the observational con- straints? The answer to this question is two-fold. First, in the spirit of mean field electrodynamics, the α-effect is not dependent on longitude and hence does not in- troduce large scale non-axisymmetric field, though at the scale of the convective eddies it necessarily involves motions and small scale fields that have longitudinal de- pendence. However, the longitudinal dependent fields are expected to be much smaller than the axisymmet- ric field arising from the α-effect. Second, a 5% non- axisymmetry associated with the high-degree (n > 3) magnetic moments will produce peak non-axisymmetric magnetic fields on the order of 5 nT along the S/C tra- jectory. This likely is still compatible with the Cassini MAG measurements. As discussed in Dougherty et al. (2018) and in Cao & Stevenson (2017a), the separation of the magnetic field of shallow origin from that of deep origin is not clear- cut. Taking a step-back to examine the large-scale field which most likely originates from the deep dynamo field, the fact that g0 3 take the same sign implies that 1 and g0 the radial magnetic flux is expelled from the equatorial region and pushed towards mid-to-high latitude (see Fig. 20A). This could originate from a deep "equatorial" jet either in the stable layer or in the deep dynamo region itself, which would tend to clear-out the radial flux so that the steady-state magnetic field approaches that of a Ferraro-corotation state: B · ∇ω = 0, here ω is the local angular velocity. Also as discussed in Dougherty et al. (2018), if a significant part of the magnetic field with n ≤ 9 has a deep origin, the poles deep inside the planet (e.g. at 0.5 RS) could feature almost zero radial magnetic field. Almost zero radial magnetic field at the poles at the deep dynamo surface could originate from flux expulsion and/or time-varying process inside a tangent cylinder (Sreenivasan & Jones 2005; Landeau et al. 2017; Schaeffer et al. 2017; Cao et al. 2018) defined by a central core (mostly likely a stably stratified fluid core instead of a solid core inside Saturn), which does not participate in the large-scale convection in the deep dynamo. 10. SUMMARY AND OUTLOOK We have analyzed the full Cassini Grand Finale MAG dataset with the goal to characterize and understand the internal magnetic field and interior of Saturn. Saturn's internal magnetic field turns out to be axisymmetric with respect to the spin-axis to an exceptional level; the dipole tilt which is a good proxy for the large-scale non- axisymmetry, must be smaller than 0.007◦ (25.2 arc- secs). This extreme level of axisymmetry sets key con- straints on the form of convection in the highly conduct- ing layer of Saturn. A stably stratified electrically con- ducting layer thicker than 2500 km above Saturn's deep dynamo could axisymmetrize Saturn's internal magnetic field to the observed level, if the dynamical feedback from the magnetic field does not enter the leading order force/vorticity balance. Furthermore, a heat transport mechanism other than pure conduction, e.g. double dif- fusive convection or waves, must exist within this layer to be compatible with the observed luminosity of Sat- urn. Although almost perfectly axisymmetric, there is a modest amount of north-south asymmetry in Saturn's internal magnetic field, directly demonstrated by the ∼ 5% northward offsets of Saturn's magnetic equator from the planetary equator. In addition to the well-resolved axisymmetric low spherical harmonic degree (n ≤ 3) magnetic moments, Saturn's magnetic field features an axisymmetric yet rich magnetic energy spectrum, which corresponds to latitudinally banded magnetic perturba- tions when viewed at the a = 0.75 RS, c = 0.6993 RS isobaric surface. Such latitudinally banded magnetic Saturn's internal magnetic field from Cassini Grand Finale 23 perturbations likely arise from a "shallow" secondary dynamo action within the semi-conducting layer of Sat- urn, enabled by differential rotation, small-scale helical motion, and the background magnetic field provided by the deep dynamo. Regularized inversion with spherical harmonic solutions as basis functions as well as trun- cated Green's function solutions demonstrated that the small-scale axisymmetric magnetic field between ±60◦ latitude at the a = 0.75 RS non-spherical "dynamo sur- face" can be well determined, while the details of the small-scale field above ±60◦ latitude are less certain. It should be noted that the area above ±60◦ latitude is less than 14% of the surface area. To fully resolve the small-scale magnetic field of Saturn above ±60◦ lati- tude, including both the axisymmetric field and the non- axisymmetric field, low altitude magnetic field measure- ments directly above the polar region are needed. This task is left to future missions to the Saturn system. APPENDIX A. GAUSS COEFFICIENTS REPRESENTATION OF THE INTERNAL PLANETARY MAGNETIC FIELD The traditional Gauss coefficients representation of the internal planetary magnetic field outside of the source region are shown here for convenience. V = (cid:88) n(cid:88) n=1 m=0 (cid:18) Rp (cid:19)n+1 Rp r [gm n cosmφ + hm n sinmφ] P m n (cosθ) , B = −∇V, Br = (n + 1) (cid:18) Rp (cid:19)n+2 n(cid:88) (cid:88) (cid:19)n+2 (cid:18) Rp n(cid:88) Bθ = −(cid:88) (cid:19)n+2 m (cid:18) Rp n(cid:88) (cid:88) m=0 m=0 n=1 n=1 r r n=1 m=0 r sinθ Bφ = [gm n cosmφ + hm n sinmφ] P m n (cosθ) , [gm n cosmφ + hm n sinmφ] dP m n (cosθ) dθ , n sinmφ − hm [gm n cosmφ] P m n (cosθ) , (A1) (A2) (A3) (A4) (A5) where Rp is the reference radius here taken to be the 1-bar equatorial radius of Saturn, (gm n ) are the Gauss coefficients, n and m are the spherical harmonic degree and order respectively, r is the spherical radial distance from the center of the planet, θ and φ are the co-latitude and east longitude respectively, and P m n (cosθ) are the Schmidt semi-normalized associated Legendre functions. n , hm B. GREEN'S FUNCTION FOR THE INTERNAL PLANETARY MAGNETIC FIELD AND THE EIGENVECTORS OF THE INVERSE PROBLEM As shown in Gubbins and Roberts (1983) and Johnson and Constable (1997), the mapping between the magnetic field at a spherical dynamo surface to anywhere above is Bobs r,θ,φ(r, θ, φ) = r (θ(cid:48), φ(cid:48))Gr,θ,φ(µ) sin θ(cid:48)dθ(cid:48)dφ(cid:48), BrD (B6) (cid:90) 2π (cid:90) π 0 0 r is the radial component of the magnetic field at the r = rD spherical dynamo surface, Bobs where BrD r,θ,φ are three components of the internal magnetic field measured above the dynamo surface, θ is colatitude, φ is longitude, and µ is the consine of the angle between the position vectors r and r(cid:48). The Green's function for each component are Gr(µ) = Gθ(µ) = − b3 4π , 1 − b2 f 3 b2 4π 1 + 2f − b2 f 3T dµ dθ , (B7) (B8) 24 and Cao et al. Gφ(µ) = − b3 4π sin θ(cid:48) 1 + 2f − b2 f 3T dµ dφ , µ = r · r(cid:48), b = rD r , f = (1 − 2bµ + b2)1/2, T = 1 + f − µb. The surface integration can be discretized, the forward problem can then be expressed as data = G model, (B9) (B10) (B11) (B12) (B13) (B14) in which data is the three component internal magnetic field at the measurement location Bobs profile of BrD here that G is a function of the position of the measurements only. r,θ,φ(r, θ, φ), model is the r , and G is the matrix expression of the integration of the Green's functions (B6). It should be emphasize The inverse problem can then be computed using the generalized inversion analysis (e.g. Jackson 1972; Connerney 1981; Aster et al. 2013). Here we briefly explain this analysis, aiming at clarifying the meaning of the eigenvector of parameter space here. Assuming there are n number of measurements and m number of parameters which means discretizing the surface integration (eq. B6) into m points on the spherical surface r = rD, data is a n× 1 vector, G is a n × m matrix, and model is a m × 1 vector. The matrix G can be factored using the singular-value-decomposition into the product (B15) in which U is a n × p matrix, Λ is a diagonal matrix of p number of non-zero eigenvalues (λ1,λ2,λ3,...,λp), and V is a m × p matrix. Each column of the V matrix, Vi, is one eigenvector in the parameter space. In our formulation, each Vi is a profile of BrD r . The solution model can then be computed as a weighted sum of the different eigenvectors in the parameter space G = U ΛV T , (B16) (B17) which for this particular problem can be expressed as (cid:88) (cid:88) i i model = βiVi, i = 1, 2, ... BrD r = βiBrD i , i = 1, 2, ... (cid:0)U T data(cid:1) here βi is a weight whose value is the ith element of the vector U T data divided by the ith eigenvalue λi: βi = i /λi. In constructing the final model solution, truncation at order imax here simply means truncating the summation in equation (B16) at order imax. C. IONOSPHERIC HALL CURRENTS AND THEIR ASSOCIATED MAGNETIC FIELD Zonal flows likely exist in the ionosphere of Saturn. The intra-D ring field-aligned current as measured along the Cassini Grand Finale orbits could arise from the ionospheric Pedersen currents driven by the zonal flows. Such zonal flows would also drive ionospheric Hall currents, which would be in the zonal ( φ) direction. Modeling of the measured Bφ combined with a global ionospheric conductivity profile (Muller-Wodarg et al. 2006; Galand et al. 2011; Muller- Wodarg et al. 2012) indicates that amplitude of the zonal flow at the ionospheric peak conductivity layer likely is 50% of that at 1 bar. Taking this value, we can make an order of magnitude estimation of the zonal ionospheric Hall current as (C18) in which ΣH is the height-integrated ionospheric Hall conductivity (∼10 S near local noon at the equator), B is the magnetic field strength, and uφ is the zonal velocity in the ionospheric peak conductivity layer. Iφ = ΣHBuφ, Saturn's internal magnetic field from Cassini Grand Finale 25 Table 7. Gauss Coefficients associated with zonal Hall currents in Saturn's Ionosphere [nT] g0 1(Hall) g0 2(Hall) g0 3(Hall) g0 4(Hall) g0 5(Hall) g0 6(Hall) g0 7(Hall) g0 8(Hall) g0 9(Hall) g0 10(Hall) 6 0.06 -4.15 -0.24 2.55 0.22 -1.26 -0.42 0.20 0.20 Since we aim at an order-of-magnitude estimation of the magnetic field associated with the ionospheric Hall current, we assume axisymmetry as a first step. In this first step, we further assume the ionospheric Hall conductivity takes the noon values at all local times, which should yield an upper bound on the current density and the associated magnetic fields. The axisymmetric assumption is a reasonable one as long as the zonal extent of the current is much wider than the spatial coverage of the measurements. One can then obtain the (Br, Bθ) associated with the zonal Hall currents via solving a boundary value problem: treating the ionospheric Hall currents as boundary currents. The boundary conditions are Br,above = Br,below, Bθ,above − Bθ,below = µ0Iφ, (C19) (C20) here above and below refers to above and below the ionosphere respectively. It can be shown that above the ionosphere, the magnetic field associated with the Hall currents can be expressed as (cid:88) VH = BH = −∇VH , (cid:18) RI (cid:19)n+1 RI r A0 nP 0 n (cos θ) , µ0I n φ , 2n + 1 n = − n A0 (cid:88) n/dθ, Iφ = n I n φ dP 0 n(cos θ) dθ . (C21) (C22) (C23) (C24) (C25) here RI is the radial distance of the ionospheric peak conductivity layer from the center of the planet and I n degree coefficients of the decomposition of Iφ onto dP 0 φ is n-th The corresponding Gauss coefficients, re-normalized with respect to the 1-bar radius, are then simply (cid:18) RI RP (cid:19)n+2 . g0 n(Hall) = A0 n ACKNOWLEDGMENTS We acknowledge support from the Cassini Project. Work at Imperial College London was funded by Science and Technology Facilities Council (STFC) consolidated grant ST/N000692/1. Work at the University of Leicester was funded by STFC consolidated grant ST/N000749/1. M.K.D. is funded by Royal Society Research Professorship RP140004. H.C. is funded by NASA Jet Propulsion Laboratory (JPL) contract 1579625. H.C.'s visit to Imperial 26 Cao et al. College London was funded by the Royal Society grant RP 180014. E.J.B. was supported by a Royal Society Wolfson Research Merit Award. The derived model parameters are given in Tables 3 - 6 and Supplementary Table 1. We thank Burkhard Militzer for providing the interior shape of Saturn and helpful discussions. Fully calibrated Cassini magnetometer data are available at the NASA Planetary Data System at https://pds.nasa.gov. Acuna, M. H., & Ness, N. F. 1980, Science, 207, 444, Christensen, U. R., & Aubert, J. 2006, Geophysical Journal REFERENCES doi: 10.1126/science.207.4429.444 Anderson, B. J., Johnson, C. L., Korth, H., et al. 2012, Journal of Geophysical Research (Planets), 117, E00L12, doi: 10.1029/2012JE004159 Anderson, J. D., & Schubert, G. 2007, Science, 317, 1384, doi: 10.1126/science.1144835 Andrews, D. J., Cowley, S. W. H., Dougherty, M. K., et al. 2012, Journal of Geophysical Research (Space Physics), 117, A04224, doi: 10.1029/2011JA017444 Aster, R. C., Borchers, B., & Thurber, C. H. 2013, in Parameter Estimation and Inverse Problems, second edition edn., ed. R. C. Aster, B. Borchers, & C. H. Thurber (Boston: Academic Press), 55 -- 91, doi: 10.1016/B978-0-12-385048-5.00003-3 Backus, G., Parker, R., & Constable, C. 1996, Foundations of Geomagnetism, 370 Backus, G. E., & Chandrasekhar, S. 1956, Proceedings of the National Academy of Science, 42, 105, doi: 10.1073/pnas.42.3.105 Bunce, E. J., Cowley, S. W. H., Alexeev, I. I., et al. 2007, Journal of Geophysical Research (Space Physics), 112, A10202, doi: 10.1029/2007JA012275 Burk, T. A. 2018, 2018 AIAA Guidance, Navigation, and Control Conference, doi: 10.2514/6.2018-2113 Burton, M. E., Dougherty, M. K., & Russell, C. T. 2009, Planet. Space Sci., 57, 1706, doi: 10.1016/j.pss.2009.04.008 Cao, H., Russell, C. T., Christensen, U. R., Dougherty, M. K., & Burton, M. E. 2011, Earth and Planetary Science Letters, 304, 22, doi: 10.1016/j.epsl.2011.02.035 Cao, H., Russell, C. T., Wicht, J., Christensen, U. R., & Dougherty, M. K. 2012, Icarus, 221, 388, doi: 10.1016/j.icarus.2012.08.007 Cao, H., & Stevenson, D. J. 2017a, Icarus, 296, 59, doi: 10.1016/j.icarus.2017.05.015 -- . 2017b, Journal of Geophysical Research: Planets, 122, 686, doi: 10.1002/2017JE005272 International, 166, 97, doi: 10.1111/j.1365-246X.2006.03009.x Christensen, U. R., & Wicht, J. 2008, Icarus, 196, 16, doi: 10.1016/j.icarus.2008.02.013 Connerney, J. E. P. 1981, Journal of Geophysical Research: Space Physics, 86, 7679, doi: 10.1029/JA086iA09p07679 Connerney, J. E. P., Acuna, M. H., & Ness, N. F. 1983, Journal of Geophysical Research, 88, 8779, doi: 10.1029/JA088iA11p08779 Connerney, J. E. P., Ness, N. F., & Acuna, M. H. 1982, Nature, 298, 44, doi: 10.1038/298044a0 Cowling, T. G. 1933, MNRAS, 94, 39, doi: 10.1093/mnras/94.1.39 Dougherty, M. K., Kellock, S., Southwood, D. J., et al. 2004, Space Sci. Rev., 114, 331, doi: 10.1007/s11214-004-1432-2 Dougherty, M. K., Achilleos, N., Andre, N., et al. 2005, Science, 307, 1266, doi: 10.1126/science.1106098 Dougherty, M. K., Cao, H., Khurana, K. K., et al. 2018, Science, 362, aat5434, doi: 10.1126/science.aat5434 Duarte, L. D. V., Gastine, T., & Wicht, J. 2013, Physics of the Earth and Planetary Interiors, 222, 22, doi: 10.1016/j.pepi.2013.06.010 Duarte, L. D. V., Wicht, J., & Gastine, T. 2018, Icarus, 299, 206, doi: 10.1016/j.icarus.2017.07.016 French, M., Becker, A., Lorenzen, W., et al. 2012, The Astrophysical Journal Supplement, 202, 5, doi: 10.1088/0067-0049/202/1/5 Galand, M., Moore, L., Mueller-Wodarg, I., Mendillo, M., & Miller, S. 2011, Journal of Geophysical Research (Space Physics), 116, A09306, doi: 10.1029/2010JA016412 Galanti, E., Kaspi, Y., Miguel, Y., et al. 2019, Geophysical Research Letters, 46, 616, doi: 10.1029/2018GL078087 Gastine, T., Wicht, J., Duarte, L. D. V., Heimpel, M., & Becker, A. 2014, Geophys. Res. Lett., 41, 5410, doi: 10.1002/2014GL060814 Cao, H., Yadav, R. K., & Aurnou, J. M. 2018, Proceedings Giampieri, G., & Dougherty, M. 2004, Annales of the National Academy of Sciences, 115, 11186, doi: 10.1073/pnas.1717454115 Geophysicae, 22, 653, doi: 10.5194/angeo-22-653-2004 Gubbins, D. 2004, Time Series Analysis and Inverse Theory Christensen, U. R. 2010, Space Science Reviews, 152, 565, doi: 10.1007/s11214-009-9553-2 for Geophysicists (Cambridge University Press), doi: 10.1017/CBO9780511840302 Saturn's internal magnetic field from Cassini Grand Finale 27 Gubbins, D., & Roberts, N. 1983, Geophysical Journal, 73, Kliore, A. J., Nagy, A., Asmar, S., et al. 2014, Geophys. 675, doi: 10.1111/j.1365-246X.1983.tb03339.x Res. Lett., 41, 5778, doi: 10.1002/2014GL060512 Guervilly, C., Hughes, D. W., & Jones, C. A. 2014, Journal Krause, F., & Radler, K. H. 1980, Mean-field of Fluid Mechanics, 758, 407?435, doi: 10.1017/jfm.2014.542 magnetohydrodynamics and dynamo theory (Pergamon) Lamy, L., Prang´e, R., Tao, C., et al. 2018, Geophysical Guillot, T., Miguel, Y., Militzer, B., et al. 2018, Nature, Research Letters, 45, 9353, doi: 10.1029/2018GL078211 555, 227, doi: 10.1038/nature25775 Hide, R., & Palmer, T. N. 1982, Geophysical and Astrophysical Fluid Dynamics, 19, 301, doi: 10.1080/03091928208208961 Holme, R., & Bloxham, J. 1996, Journal of Geophysical Research: Planets, 101, 2177, doi: 10.1029/95JE03437 Hood, L. L., Herbert, F., & Sonett, C. P. 1982, Journal of Geophysical Research: Solid Earth, 87, 5311, doi: 10.1029/JB087iB07p05311 Hunt, G., Cowley, S., Provan, G., et al. 2019, Journal of Geophysical Research (Space Physics), 124, 5675, doi: 10.1029/2019JA026588 Hunt, G. J., Provan, G., Bunce, E. J., et al. 2018, Journal of Geophysical Research (Space Physics), 123, 3806, doi: 10.1029/2017JA025067 Hunt, G. J., Cowley, S. W. H., Provan, G., et al. 2014, Journal of Geophysical Research (Space Physics), 119, 9847, doi: 10.1002/2014JA020506 -- . 2015, Journal of Geophysical Research (Space Physics), 120, 7552, doi: 10.1002/2015JA021454 Iess, L., Militzer, B., Kaspi, Y., et al. 2019, Science, 364, aat2965, doi: 10.1126/science.aat2965 Jackson, A., Constable, C., & Gillet, N. 2007, Geophysical Journal International, 171, 995, doi: 10.1111/j.1365-246X.2007.03530.x Jackson, D. D. 1972, Geophysical Journal, 28, 97, doi: 10.1111/j.1365-246X.1972.tb06115.x Jia, X., Slavin, J. A., Gombosi, T. I., et al. 2015, Journal of Geophysical Research: Space Physics, 120, 4763, doi: 10.1002/2015JA021143 Johnson, C. L., & Constable, C. G. 1997, Geophysical Journal International, 131, 643, doi: 10.1111/j.1365-246X.1997.tb06604.x Kaspi, Y., Galanti, E., Hubbard, W. B., et al. 2018, Nature, 555, 223, doi: 10.1038/nature25793 Kelbert, A., Schultz, A., & Egbert, G. 2009, Nature, 460, 1003, doi: 10.1038/nature08257 Khurana, K. K., Dougherty, M. K., Provan, G., et al. 2018, Geophys. Res. Letts, 45, 10,068, doi: 10.1029/2018GL078256 Landeau, M., Aubert, J., & Olson, P. 2017, Earth and Planetary Science Letters, 465, 193, doi: 10.1016/j.epsl.2017.02.004 Leconte, J., & Chabrier, G. 2012, Astronomy and Astrophysics, 540, A20, doi: 10.1051/0004-6361/201117595 -- . 2013, Nature Geoscience, 6, 347, doi: 10.1038/ngeo1791 Levenberg, K. 1944, Quarterly of Applied Mathematics, 2, 164 Liu, J., Goldreich, P. M., & Stevenson, D. J. 2008, Icarus, 196, 653, doi: 10.1016/j.icarus.2007.11.036 Lorenzen, W., Holst, B., & Redmer, R. 2009, Phys. Rev. Lett., 102, 115701, doi: 10.1103/PhysRevLett.102.115701 Mankovich, C., Marley, M. S., Fortney, J. J., & Movshovitz, N. 2019, The Astrophysical Journal, 871, 1, doi: 10.3847/1538-4357/aaf798 Marquardt, D. W. 1963, Journal of the Society for Industrial and Applied Mathematics, 11, 431 Militzer, B., Wahl, S., & Hubbard, W. B. 2019, The Astrophysical Journal, 879, 78, doi: 10.3847/1538-4357/ab23f0 Moore, K. M., & Bloxham, J. 2017, Journal of Geophysical Research (Planets), 122, 1443, doi: 10.1002/2016JE005238 Moore, K. M., Bloxham, J., Connerney, J. E. P., Jørgensen, J. L., & Merayo, J. M. G. 2017, Geophysical Research Letters, 44, 4687, doi: 10.1002/2017GL073133 Moore, K. M., Cao, H., Bloxham, J., et al. 2019, Nature Astronomy, 3, 730, doi: 10.1038/s41550-019-0772-5 Morales, M. A., Schwegler, E., Ceperley, D., et al. 2009, Proceedings of the National Academy of Science, 106, 1324, doi: 10.1073/pnas.0812581106 Muller-Wodarg, I., Moore, L., Galand, M., Miller, S., & Mendillo, M. 2012, Icarus, 221, 481, doi: 10.1016/j.icarus.2012.08.034 Muller-Wodarg, I. C. F., Mendillo, M., Yelle, R. V., & Aylward, A. D. 2006, Icarus, 180, 147, doi: 10.1016/j.icarus.2005.09.002 Ness, N. F., Acuna, M. H., Behannon, K. W., et al. 1982, Khurana, K. K., Kivelson, M. G., Stevenson, D. J., et al. Science, 215, 558, doi: 10.1126/science.215.4532.558 1998, Nature, 395, 777, doi: 10.1038/27394 Kivelson, M., Khurana, K., & Volwerk, M. 2002, Icarus, 157, 507 , doi: https://doi.org/10.1006/icar.2002.6834 Ness, N. F., Acuna, M. H., Lepping, R. P., et al. 1981, Science, 212, 211, doi: 10.1126/science.212.4491.211 Parker, E. N. 1955, ApJ, 122, 293, doi: 10.1086/146087 28 Cao et al. Persoon, A. M., Kurth, W. S., Gurnett, D. A., et al. 2019, Soderlund, K. M., King, E. M., & Aurnou, J. M. 2012, Geophys. Res. Lett., 46, 3061, doi: 10.1029/2018GL078020 Provan, G., Cowley, S. W. H., Bradley, T. J., et al. 2018, Journal of Geophysical Research (Space Physics), 123, 3859, doi: 10.1029/2018JA025237 Provan, G., Cowley, S. W. H., Bunce, E. J., et al. 2019a, Journal of Geophysical Research (Space Physics), 124, 379, doi: 10.1029/2018JA026121 Provan, G., Lamy, L., Cowley, S. W. H., & Bunce, E. J. 2019b, Journal of Geophysical Research (Space Physics), Earth and Planetary Science Letters, 333, 9, doi: 10.1016/j.epsl.2012.03.038 Southwood, D. J., Dougherty, M. K., Balogh, A., et al. 2001, Journal of Geophysical Research (Space Physics), 106, 30109, doi: 10.1029/2001JA900110 Sreenivasan, B., & Jones, C. A. 2005, Geophysical Research Letters, 32, L20301, doi: 10.1029/2005GL023841 Stanley, S. 2010, Geophys. Res. Lett., 37, L05201, doi: 10.1029/2009GL041752 Steenbeck, M., & Krause, F. 1966, Z. Naturforsch, 21a, 1285 Steenbeck, M., Krause, F., & Radler, K.-H. 1966, Z. 124, 1157, doi: 10.1029/2018JA026079 Naturforsch, 21a, 369 Purucker, M., Ravat, D., Frey, H., et al. 2000, Geophysical Stevenson, D. J. 1975, Phys. Rev. B, 12, 3999, Research Letters, 27, 2449, doi: 10.1029/2000GL000072 doi: 10.1103/PhysRevB.12.3999 Read, P. L., Dowling, T. E., & Schubert, G. 2009, Nature, 460, 608, doi: 10.1038/nature08194 Roberts, P. H., & King, E. M. 2013, Reports on Progress in Physics, 76, 096801, doi: 10.1088/0034-4885/76/9/096801 Roberts, P. H., & Stix, M. 1971, NCAR Technical Note, NCAR/TN-60+IA, doi: 10.5065/D6DJ5CK7 Rubio, A. M., Julien, K., Knobloch, E., & Weiss, J. B. 2014, Phys. Rev. Lett., 112, 144501, doi: 10.1103/PhysRevLett.112.144501 Schaeffer, N., Jault, D., Nataf, H.-C., & Fournier, A. 2017, Stevenson, D. J. 1980, Science, 208, 746, doi: 10.1126/science.208.4445.746 -- . 1982, Geophysical and Astrophysical Fluid Dynamics, 21, 113, doi: 10.1080/03091928208209008 -- . 2003, Earth and Planetary Science Letters, 208, 1, doi: 10.1016/S0012-821X(02)01126-3 -- . 2010, Space Science Reviews, 152, 651, doi: 10.1007/s11214-009-9572-z Stevenson, D. J., & Salpeter, E. E. 1977, The Astrophysical Journal Supplement, 35, 239, doi: 10.1086/190479 Sulaiman, A. H., Kurth, W. S., Hospodarsky, G. B., et al. Geophysical Journal International, 211, 1, doi: 10.1093/gji/ggx265 2018, Geophys. Res. Lett., 45, 7347, doi: 10.1029/2018GL078130 Smith, E. J., Davis, L., Jones, D. E., et al. 1980, Science, Wahlund, J. E., Morooka, M. W., Hadid, L. Z., et al. 2018, 207, 407, doi: 10.1126/science.207.4429.407 Smith, E. J., Dougherty, M. K., Russell, C. T., & Southwood, D. J. 2001, Journal of Geophysical Research (Space Physics), 106, 30129, doi: 10.1029/2001JA900115 Science, 359, 66, doi: 10.1126/science.aao4134 Weir, S. T., Mitchell, A. C., & Nellis, W. J. 1996, Physical Review Letters, 76, 1860, doi: 10.1103/PhysRevLett.76.1860
1212.6379
1
1212
2012-12-27T14:41:29
Variability of Water and Oxygen Absorption Bands in the Disk-Integrated Spectra of the Earth
[ "astro-ph.EP" ]
We study the variability of major atmospheric absorption features in the disk-integrated spectra of the Earth with future application to Earth-analogs in mind, concentrating on the diurnal timescale. We first analyze observations of the Earth provided by the EPOXI mission, and find 5-20% fractional variation of the absorption depths of H2O and O2 bands, two molecules that have major signatures in the observed range. From a correlation analysis with the cloud map data from the Earth Observing Satellite (EOS), we find that their variation pattern is primarily due to the uneven cloud cover distribution. In order to account for the observed variation quantitatively, we consider a simple opaque cloud model, which assumes that the clouds totally block the spectral influence of the atmosphere below the cloud layer, equivalent to assuming that the incident light is completely scattered at the cloud top level. The model is reasonably successful, and reproduces the EPOXI data from the pixel-level EOS cloud/water vapor data. A difference in the diurnal variability patterns of H2O and O2 bands is ascribed to the differing vertical and horizontal distribution of those molecular species in the atmosphere. On the Earth, the inhomogeneous distribution of atmospheric water vapor is due to the existence of its exchange with liquid and solid phases of H2O on the planet's surface on a timescale short compared to atmospheric mixing times. If such differences in variability patterns were detected in spectra of Earth-analogs, it would provide the information on the inhomogeneous composition of their atmospheres.
astro-ph.EP
astro-ph
Draft version November 7, 2018 Preprint typeset using LATEX style emulateapj v. 10/09/06 VARIABILITY OF WATER AND OXYGEN ABSORPTION BANDS IN THE DISK-INTEGRATED SPECTRA OF THE EARTH Yuka Fujii1, Edwin L. Turner2,3, and Yasushi Suto1,2,4. Draft version November 7, 2018 ABSTRACT We study the variability of major atmospheric absorption features in the disk-integrated spectra of the Earth with future application to Earth-analogs in mind, concentrating on the diurnal timescale. We first analyze observations of the Earth provided by the EPOXI mission, and find 5-20% fractional variation of the absorption depths of H2O and O2 bands, two molecules that have major signatures in the observed range. From a correlation analysis with the cloud map data from the Earth Observing Satellite (EOS), we find that their variation pattern is primarily due to the uneven cloud cover distribution. In order to account for the observed variation quantitatively, we consider a simple opaque cloud model, which assumes that the clouds totally block the spectral influence of the atmosphere below the cloud layer, equivalent to assuming that the incident light is completely scattered at the cloud top level. The model is reasonably successful, and reproduces the EPOXI data from the pixel- level EOS cloud/water vapor data. A difference in the diurnal variability patterns of H2O and O2 bands is ascribed to the differing vertical and horizontal distribution of those molecular species in the atmosphere. On the Earth, the inhomogeneous distribution of atmospheric water vapor is due to the existence of its exchange with liquid and solid phases of H2O on the planet's surface on a timescale short compared to atmospheric mixing times. If such differences in variability patterns were detected in spectra of Earth-analogs, it would provide the information on the inhomogeneous composition of their atmospheres. Subject headings: Earth -- scattering -- techniques: spectroscopic 1. INTRODUCTION Determining the nature of the atmospheres and sur- faces of exoplanets is of primary importance in prob- ing not only their formation history but also in identi- fying possible signatures of life. While it is very chal- lenging, and perhaps only feasible through direct photo- metric/spectroscopic observations of the planetary light resolved from that of the host star, there are several pro- posals for eventual astrobiological investigations of po- tentially habitable, rocky exoplanets (e.g., Levine et al. 2009; Savransky et al. 2010; Matsuo & Tamura 2010). The available exoplanetary light would be disk- integrated, i.e., that of a point source without any spatial resolution. Deciphering the exoplanetary light properly, therefore, is inevitably a highly difficult task, in partic- ular for those planets with diverse surface types and at- mospheres like our own Earth. Indeed, habitable planets are likely to exhibit a variety of complex patterns of their surfaces and atmospheres, intrinsically dependent on their climatology and geol- ogy. According to the global water cycle, liquid water on the planetary surface vaporizes, forms clouds, is car- ried by atmospheric circulation, and precipitates as rain- fall/snowfall. Depending on the total amount of water and the atmospheric circulation pattern, the surface of Electronic address: [email protected] 1 Department of Physics, The University of Tokyo, Tokyo 113- 0033, Japan 2 Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544 3 Kavli Institute for the Physics and Mathematics of the Uni- verse, The University of Tokyo, Kashiwa 277-8568, Japan 4 Research Center for the Early Universe, Graduate School of Science, The University of Tokyo, Tokyo 113-0033, Japan the habitable planets may be partially covered by ocean (e.g. Abe et al. 2011), and be observed only through at- mospheres with highly inhomogeneous and variable cloud cover patterns. Towards that goal, Given these complexities, techniques to properly de- cipher the disk-integrated light of exoplanets need to be developed. the time variation of planetary light due to spin rotation and orbital revolution is a powerful tool, and sev- eral authors have computed the expected variation patterns and proposed the reconstruction methods (Ford et al. 2001; Tinetti et al. 2006a,b; Cowan et al. 2009; Oakley & Cash 2009; Kawahara & Fujii 2010; Robinson et al. 2010; Cowan et al. 2011; Robinson et al. 2011; Kawahara & Fujii 2011; Fujii et al. 2010, 2011; Fujii & Kawahara 2012; Sanrom´a & Pall´e 2012). The variation of the continuum level in the visible to near-infrared (NIR) range mainly reflects the distribu- tion of landmass, ocean, cloud cover, and possibly vege- tation. The peak-to-trough diurnal variability for 0.1µm- wide photometry of the Earth in the visible/NIR range is found to be 10-30% (Livengood et al. 2011). Thermal emission of the Earth also shows a few percent of diurnal variation in the mid-infrared (G´omez-Leal et al. 2012), which primarily originates from the uneven cloud cover and humidity. In addition to the light-curve in broad-band photome- try mentioned above, molecular absorption depths ex- hibit diurnal variation. Since atmospheric absorption depths of molecules are determined by the column den- sity of the corresponding molecules along the optical path, they are sensitive to the presence of highly reflec- tive cloud cover in the atmosphere that effectively blocks the spectral influence of molecules in the lower atmo- 2 sphere. For idealized uniformly mixed atmospheres, the diurnal variation of molecular absorption features should be strongly linked to the spatial distribution of clouds. In reality, the distribution of molecules itself is not entirely uniform and may be time-dependent. Moreover, the spatial and time variations are not the same for differ- ent molecular species comprising the atmosphere. In par- ticular, the local column density of water vapor is known to vary widely on a timescale of ∼ 1 hr (Blake & Shaw 2011), while other molecules, such as N2 and O2, are well-mixed in the troposphere and thus show no signifi- cant variations on short timescales. This paper extends our previous work on photomet- ric light-curves in visible/NIR bands (Fujii et al. 2010, 2011), and examines the diurnal variation of molecu- lar absorption signatures in the disk-integrated reflection spectra of the Earth with future application to Earth- analogs in mind5. We focus on the absorption bands of H2O at 1.13µm and O2 at 1.27µm, which are among the most prominent absorption bands in NIR and are an indicator of habitability and a biosignature molecule, re- spectively. We first analyze data from space-based NIR spectroscopy of the Earth by NASA's EPOXI mission. Then we consider a simple model (referred to as opaque cloud model) that reasonably reproduces the observed variation pattern, if the cloud pattern data are provided separately. Our model implies that the different behav- ior of water vapor and oxygen in their absorption band variations can be ascribed to their intrinsically different spatial distribution patterns, rather than to the common cloud coverage. We discuss how this difference can be used to probe signatures of surface and atmospheric in- homogeneity of exoplanets with next-generation direct imaging. The organization of this paper is as follows. Section 2 introduces the EPOXI data used in this paper and analy- ses the correlation between absorption depths and pixel- to-pixel climatological data obtained with Earth Observ- ing Satellites. Section 3 describes our model to reproduce the variation pattern of absorption depths and compares the simulation results with observation. Section 4 fur- ther discusses the different behavior between H2O varia- tion and O2 variation. Finally, Section 5 draws our main conclusions and discusses the implication for future ob- servation of exoplanets. An analysis of CO2, a molecule with properties intermediate between those of H2O and O2 in some ways, is presented in Appendix A. 2. ANALYSIS OF EPOXI DATA 2.1. EPOXI observation The EPOXI6 mission (Livengood et al. 2011) per- formed photometric and spectroscopic monitoring of the Earth and the Moon from space as a benchmark for fu- ture characterization of terrestrial exoplanets. A part of the observations were devoted to spectroscopy of the 5 The term "Earth-like" has been used extensively in the litera- ture, often without any precise definition. In this paper, the term "Earth-analogs" is used to refer to rocky planets that resemble the Earth in having surface temperatures, obliquities, continents, oceans and atmospheres sufficiently like the Earth's to give them similar global hydrologies. 6 The Deep Impact flyby spacecraft for the Extrasolar Planetary Observation and Characterization investigation (EPOCh) and the Deep Impact eXtended Investigation (DIXI) Equatorial 1, (March 2008) O2 H2O CO2 H2O y t i v i t c e l f e R 0.15 0.1 0.05 0 t=0 t=4 t=8 t=12 t=16 t=20 t=24 CO2 H2O 1.2 1.4 1.6 1.8 2 Wavelength [µm] Fig. 1. -- Examples of NIR reflection spectra of the Earth observed by EPOXI in March of 2008 (Earth 1: equinox) (Livengood et al. 2011; Robinson et al. 2011) . Different lines rep- resent different short exposures obtained at 2-hour intervals and labeled by the time t[hr] from the start of the observations on that date (see Figure 2). Three vertical dashed lines indicate the wavelength ranges that we adopt to compute the equivalent widths (H2O: 1.07-1.24µm, O2: 1.24µm-1.283µm, CO2: 1.59-1.62µm). disk-integrated scattered light of the Earth over the wavelength of 1.10-4.54µm. These observations were car- ried out on 2008 March 18-19, 2008 June 4-5, 2009 March 27-28, and 2009 October 4-5, with 12 exposures every two hours of the day (integration time for each exposure is less than 2 sec). The sub-observer latitudes (the latitude of the intersection between the Earth's surface and the line connecting the center of the Earth and the detector) are 1◦.7 N, 0◦.3 N, 61◦.7 N, and 73◦.8 S, respectively. Following Cowan et al. (2011), we henceforth refer to these 4 observations as Earth1:equinox, Earth5:solstice, Polar1:north, and Polar2:south. Figure 1 displays the 1-2µm portion of the observed reflection spectra of the Earth in Earth1. The broad absorption features at 1.08-1.18, 1.30-1.53, and 1.75- 1.99µm are mostly due to H2O, and the narrower fea- ture around 1.27µm is due to O2 plus oxygen collision complexes O2·O2 and O2·N2 (Pall´e et al. 2009, and ref- erences therein). Absorptions at 1.6µm and 2.0µm are signatures of CO2 (e.g. Robinson et al. 2011). We measure the equivalent widths of H2O (wH2O) and O2 (wO2 ) for each exposure. For wH2O, we focus on the spectral features centered at ∼ 1.13µm and consider the absorption from 1.07µm to 1.24µm. For wO2 , we use the spectral features centered at ∼ 1.27µm and consider the absorption from 1.24µm to 1.283µm. In each case, the continuum line is assumed to connect the data points at both boundaries linearly. Additionally, we consider the variation of reflectivity at 1.24µm as a measure of the continuum level. Figure 2 shows the diurnal fluctuations of wH2O and wO2 , as well as the variation of the reference continuum level (1.24 µm). Table 1 summarizes the average and the fractional variation amplitudes of continuum level, wH2O and wO2 . The variation amplitude defined by (maximum-minimum)/average is typically 5-20%. Fig- ure 2 clearly indicates that the diurnal patterns of wH2O and wO2 significantly differ from that of the continuum level which primarily traces the continental distribution on the Earth (see e.g. Cowan et al. 2009, 2011; Fujii et al. 2011). Although not shown here in detail, we also con- Earth1:equinox Earth5:solstice Polar1:north Polar2:south 3 0.14 0.12 0.1 0.08 550 ) m 4 2 . 1 ( f e R ] Å [ O 2 H w ] Å [ 2 O w 500 450 400 45 40 35 0 6 12 Time [hr] 18 24 0.1 0.08 0.06 560 ) m 4 2 . 1 ( f e R ] Å [ O 2 H w ] Å [ 2 O w 520 480 440 48 44 40 36 0 6 12 Time [hr] 18 24 0.12 0.09 0.06 450 ) m 4 2 . 1 ( f e R ] Å [ O 2 H w ] Å [ 2 O w 400 350 51 48 45 42 0.1 0.09 0.08 0.07 400 ) m 4 2 . 1 ( f e R ] Å [ O 2 H w ] Å [ 2 O w 0 6 12 Time [hr] 18 24 380 360 340 48 46 44 0 6 12 Time [hr] 18 24 Fig. 2. -- Diurnal variations of continuum level (top), equivalent width of H2O absorption centered at 1.13µm (middle), and that of O2 centered at 1.27µm (bottom). Symbols indicate observed data points and the connecting lines are drawn only to guide the eyes. Snapshots at corresponding times were generated at http://www.fourmilab.ch/cgi-bin/uncgi/Earth are attached. Diurnal Variation of Continuum Level, Equivalent Width of H2O at 1.13µm and that of O2 at 1.27µm. TABLE 1 Earth1:equinox (Mar.2008) Earth5:solstice (Jun.2008) Polar1:north (Mar.2009) Polar2:south (Oct.2009) Reflectivity at 1.24µm ave. (max-min)/ave 0.111 0.084 0.084 0.084 32.1% 30.6% 29.5% 18.2% H2O at 1.13µm O2 at 1.27µm ave.[A] (max-min)/ave ave.[A] (max-min)/ave 457 499 392 372 16.5% 18.2% 16.1% 9.4% 39 42 46 47 16.8% 18.7% 7.2% 5.5% firmed that absorption features of H2O at other wave- lengths (centered at ∼ 1.4µm and ∼ 1.85µm) exhibit variation patterns matching those at the 1.13µm band. For equatorial observations, the general trend that the depth is weaker in the first half of the day and becomes stronger in the second half is evident, and is shared by both H2O and O2. Referring to the snapshots shown in the bottom panels in Figure 2, the weaker absorption corresponds to the time when the Indonesia, a persis- tently cloudy region, dominates the field-of-view (see also G´omez-Leal et al. 2012). This already demonstrates the strong relation between the absorption bands and cloud cover. In the next subsection, we investigate this con- nection in more detail. 2.2. Correlation between diurnal variabilities and ocean/cloud parameters As mentioned in Section 2.1, the variation pattern in Figure 2 is likely correlated with the extent and nature of cloud cover. In order to confirm the correlation be- tween the absorption features and clouds, we collect daily global maps of atmospheric parameters for the corre- sponding days from Terra/MODIS Atmosphere Level 3 Product, which is available online7. Each global map is derived on a pixel-to-pixel basis from the Remote Sensing data obtained with MOderate Resolution Imag- ing Spectroradiometer (MODIS) onboard the Earth Ob- serving Satellites Terra and Aqua (Dorothy et al. 2006). Among the various parameters reported, we choose four which we suspect will influence the diurnal absorption variations strongly, including Atmospheric Water Vapor 7 http://ladsweb.nascom.nasa.gov/ Mean, Cloud Top Pressure Mean (Pctp), Cloud Frac- tion Mean (fcld), and Cloud Optical Thickness Combined Mean (τcld). We further add the ocean fraction as a fifth parameter. We compute the weighted average of each parameter for each EPOXI exposure. We adopt the geometric weight as Max{µ0µ1, 0}, where µ0/µ1 denote the cosine between the normal direction of the surface and the direction to- ward the Sun/observer. We adopt this factor because each surface pixel would contribute to the total reflec- tivity with that weight if the scattering were isotropic (Lambert's Law) (e.g. Lester et al. 1979); while this is not strictly the case, it is a reasonable approximation for these purposes. Figure 3 displays Spearman's rank correlation coeffi- cients between absorption depths of H2O (left) and O2 (right) measured by EPOXI observation and parameters taken from MODIS. Spearman's rank coefficient rs is cal- culated by converting each value xi (i:index for exposures in one series of observation) to the rank Xi. If xj is the n-th largest value of all possible values of xi, the rank is defined as Xj = n. Then we calculate the correlation coefficient of the ranks of two different parameters xi and yi as rs = Xi sXi ¯X =Xi (Xi − ¯X)(Yi − ¯Y ) (Xi − ¯X)2sXi ¯Y =Xi Xi, Yi. (Yi − ¯Y )2 , (1) (2) Since the chosen parameters do not necessarily follow a m m m m 4 O 2 H w h t i w t n e i c i f f e o c k n a r s ' n a m r a e p S 2 O w h t i w t n e i c i f f e o c k n a r s ' n a m r a e p S H2O Earth1:equinox Earth5:solstice Polar1:north Polar2:south WV Pctp t cld fcld Ocean O2 Earth1:equinox Earth5:solstice Polar1:north Polar2:south 1 0.5 0 -0.5 -1 1 0.5 0 -0.5 -1 WV Pctp t cld fcld Ocean Fig. 3. -- Spearman's rank correlation coefficient between ab- sorption depth (left:H2O, right:O2) and Atmospheric Water Vapor (WV), Cloud Top Pressure (Pctp), Cloud Optical Thickness (τftp), Cloud Cover Fraction (fcld), and Ocean Fraction (Ocean), respec- tively. Horizontal dotted lines show the significance level of 0.05 for 13 samples (# of exposures per observation). gaussian distribution, we adopt Spearman's rank corre- lation coefficients instead of Pearson's correlation coeffi- cients. In most cases, the absorption depths are positively cor- related with the cloud top pressure (or equivalently, anti- correlated with the cloud altitude), while anti-correlated with both cloud optical thickness and the cloud cover fraction. These trends are plausibly understood as fol- lows: For cloud top pressure, the incident light is scat- tered at the altitude of the upper cloud layer without absorption by molecules in the atmosphere below the cloud. Since molecules are abundant in the lower atmo- sphere, the higher cloud altitude reduces the absorption depth more effectively. As for cloud optical thickness, thicker clouds scatter the incident light more completely and thus reduce absorptions below that layer. Addition- ally, the higher reflectivity of the thicker cloud increases the contribution of the region to the disk-integrated spec- tra. Similarly, the absorption depths become weaker as the cloud cover fraction increases. Although more weakly, the absorption depths appear to be slightly anti-correlated with the ocean fraction. The lower reflectivity of ocean reduces the absorption depths of the disk-integrated spectra8. We should emphasize here, however, that the above 8 The light reflected from ocean in principle contains the sig- nature of liquid water (Palmer & Williams 1974), which is shifted toward longer wavelength and less prominent compared to the sig- nature of water vapor we consider in this paper. interpretation requires some caution because the five pa- rameters we adopted are not necessarily independent. In the equatorial region, for instance, there exists a positive correlation between cloud optical thickness and cloud top altitude. There is also generally positive correlation be- tween cloud cover and atmospheric water vapor, likely leading to the counter-intuitive anti-correlation between the atmospheric water vapor and the absorption depths. Figure 3 also shows that the correlation of the H2O band with clouds varies more significantly than that of O2, depending on the location of the sub-observer's lat- itude. In particular, the correlation of H2O for Polar2 observation even changes its sign compared to the other three cases. This is not the case for O2. We will discuss this behavior in detail in Section 4. 3. A SIMPLE OPAQUE CLOUD MODEL In the previous section, we discussed the influence of clouds on absorption features by considering the cor- relation between the absorption depths and individual cloud/surface parameter. In this section, we try to in- terpret the observed variation in absorption depths more quantitatively. For that purpose, we consider a simple opaque cloud model (see also e.g. Harrison & Coombes 1988), which assumes that the cloud cover totally blocks the influence of the lower atmosphere and that the inci- dent light is completely scattered at the cloud top. 3.1. Model Description The equivalent width of an atmospheric absorption band in the disk-integrated scattered light of the planet can be expressed as wmodel =Z λ0+∆λ λ0−∆λ dλ RSIV dΩ ≡ sin θdθdφ, cλ(θ, φ)g(θ, φ) dΩ cλ(θ, φ) [1 − e−τλ] g(θ, φ) dΩ RSIV g(θ, φ) ≡ µ0µ1, , (3) (4) where (θ, φ) represents the polar-coordinate of the sur- face point on the planet, µ0 = µ0(θ, φ) and µ1 = µ1(θ, φ) are the directional cosines between the normal direction of the surface point and the direction toward the Sun and the observer, respectively. The wavelength-dependent continuum level (i.e., reflectivity) and optical depth are denoted by cλ(θ, φ) and τλ. The interval of integration over λ is centered at the center of the absorption band, λ0. The integral over solid angle of planetary surface is performed over the illuminated and visible portion, SIV. The above expression is based on the approxima- tions that 1) the lower boundary scatters the light ac- cording to the Lambert Law, i.e. the scattered intensity is independent of the emergent direction, and 2) scatter- ing by the atmosphere is negligible. We further neglect the wavelength-dependence of the continuum level, and assume that the wavelength depen- dence of cλ(θ, φ) can be factored out, i.e., cλ(θ, φ) = h(λ)c(θ, φ) with h(λ) being a function of wavelength alone. Then, equation (3) may be approximated as wmodel =RSIV c(θ, φ)W (N ; θ, φ)g(θ, φ)dΩ , c(θ, φ)g(θ, φ)dΩ W (N ; θ, φ)≡Z dλ (1 − exp{−τλ(N ; θ, φ)}) , τλ(N ; θ, φ) =Z ∞ ∼(cid:18) 1 RSIV n(z)σλ(cid:18) 1 µ1(cid:19)R n(z)dz ·R n0(z)σλdz µ1(cid:19) dz R n0(z)dz µ0 + + 1 1 z0 = τλ,0 , µ0 Neff N0 Neff ≡(cid:18) 1 µ0 + 1 µ1(cid:19)Z ∞ z0 n(z)dz , (5) (6) (7) (8) where W (N ; θ, φ) is the equivalent width at each sur- face patch, σλ is the absorption coefficient, N0 is the total column density of molecules under the canonical atmospheric model, Neff is the effective column density obtained by integrating n(z) along the optical path from the top of atmosphere down to the boundary at altitude z = z0, and τλ,0 is the absorption depth of the canonical model. Equation (7) is only approximately valid because in reality the absorption coefficient σλ depends on the T-P profile of atmosphere. Nevertheless, we employ this approximation for the sake of simplicity. In practice, we divide the planetary surface into 2◦×2◦ pixels, and treat the cloudless and cloudy portions in each pixel separately for convenience, as described below. Then, equation (5) is discretized as wmodel = Xi (cid:8)(1 − f i cld) W i nocldci nocld + f i cld W i cldci cld(cid:9) giδΩi cld(cid:9) giδΩi cld) ci nocld + f i cld ci Xi (cid:8)(1 − f i (9) where i is the index for the surface pixels, fcld is the cloud cover fraction, and the suffix nocld/cld indicates the value at the cloudless/cloudy portion. The position-dependent parameters in equation (9) are determined on the basis of daily global maps which are extracted from the same MODIS data product used in Section 2.2. For instance, the value for fcld is adopted from Cloud Fraction Mean. Input data for other parameters are described below. The effective column density Neff of H2O at each patch is determined by Atmospheric Water Vapor Mean (w) and Cloud Top Pressure Mean (Pctp) as well as the ge- ometric factors µ0 and µ1. For the cloudless portion, Neff is identical to (1/µ0 + 1/µ1)w because we can safely set z0 = 0. For cloudy portions of a pixel, however, we need to define the boundary height z0 in equation (7) that is essentially the layer below which the the molecule in question does not influence the emergent spectra. In the present opaque cloud model, we assume that the syn- thetic absorption depth is completely unaffected by the atmosphere below the cloud top altitude zctp, and thus set z0 = zctp. According to this opaque cloud assump- tion, N is determined once the vertical profile of water vapor and the altitude of the cloud top layer are given. In what follows, we simply assume that the vertical profile of water vapor is proportional to that of the US Standard 5 Model, and determine the proportionality coefficient so that the total column density matches the value of w. The altitude of the cloud top layer is derived from Pctp assuming the US Standard Temperature-Pressure (T-P) profile9. The total column density Neff for O2 is esti- mated in the same way except that we neglect the hor- izontal inhomogeneity, and determine it only from Pctp and the geometric factor. ] m o r t s g n A [ h t d i W t n e l a v i u q E H2O O2 CO2 10000 1000 100 10 1 0.1 0.01 0.01 0.1 1 10 100 N/NUSstandard , Fig. 4. -- Curve of growth for the equivalent width of the wa- ter vapor absorption band at 1.13µm-1.24µm, oxygen at 1.24µm- 1.283µm, and carbon dioxide at 1588µm-1.622µm calculated by lbl2od with GEISA 2011 database under the no-cloud condition with surface albedo 0.3. The remaining task is to compute the equivalent width W as a function of the effective column density. For this purpose, we compute line-by-line optical depths τλ,0 under the US Standard Atmosphere with the GEISA 2011 molecule spectroscopic database10 and the pub- lic code lbl2od11. Then we calculate R λ0+∆λ λ0−∆λ e−κτλ,0dλ with varying κ. Figure 4 shows the resultant curve of growth of the equivalent widths of H2O measured at 1.06 − 1.24µm, O2 at 1.235 − 1.310µm as well as CO2 at 1.560−1.630µm used in Appendix A. Based on these cal- culations, the normalized effective column density of each patch, N i eff /NUSstandard, is translated into the equivalent width W i by equating N i eff/NUSstandard to κ in Figure 4. Finally, the continuum level c in equation (9) is es- timated using the 2-stream approximation (e.g. Liou 1980). We consider a non-absorbing atmosphere with optical thickness τcld and asymmetry factor β. Denoting the reflectivity at the surface by rg, the net reflectivity at the top of the atmosphere is: c(τcld) = α"1 − 1 + (√3/2)(1 − rg)(1 − β)τcld# , (10) 1 − rg where we adopt the optical thickness of the cloud (Cloud Optical Thickness Combined Mean) for τcld. The typical value for β of Earth's clouds at the relevant wavelengths is β = 0.85 (e.g. Liou 1980). The overall scaling fac- tor α is introduced here to empirically incorporate the anisotropic scattering due to the bulk cloud cover. We 9 T-P profiles are not significantly different in different atmo- spheric models. 10 http://ether.ipsl.jussieu.fr/etherTypo/?id=1293 11 http://www.libradtran.org/doku.php?id=lbl2od 6 determine the value of α by hand so as to match the disk-averaged continuum level to the observed data. According to the procedures described above, we com- pute W i and ci at each pixel, and then obtain the pre- dicted absorption depth in the disk-averaged spectra of the Earth from equation (9). The model prediction is compared with the EPOXI data in Section 3.2. 3.2. Comparison of the opaque cloud model and the EPOXI Data Figure 5 compares the EPOXI data (symbols) to our model predictions (dotted lines). The left, middle and right panels correspond to the disk-integrated continuum level, H2O absorption, and O2 absorption, respectively. The anisotropic parameter for bulk cloud scattering α in equation (10) is determined so that the disk-averaged continuum level matches the observed value (Table 2). TABLE 2 Normalization Parameters and Comparison in Daily Average and Standard Deviation between Simulations and Observations. Earth1 Earth5 Polar1 Polar2 α ¯wH2O,model/ ¯wH2O,obs ¯wO2,model/ ¯wO2,obs σH2O,model/σH2O,obs σO2,model/σO2,obs 0.61 0.85 0.65 1.32 0.57 0.64 0.87 0.66 1.73 0.77 0.73 0.88 0.64 0.70 0.61 0.79 0.88 0.63 0.92 0.88 We first consider the comparison between the simula- tion and the observation in terms of the daily average, ¯w, and the standard deviation, ¯σ. Table 2 summarizes the ratio of the daily average of the simulated ¯wmodel, to the observed one, ¯wobs, ¯wmodel/ ¯wobs, and that of stan- dard deviation ¯σmodel/¯σobs. The daily simulation average typically results in a 10-15% underestimation for H2O and 35% underestimation for O2. The underestimation is likely due to our "opaque cloud" assumption which tends to diminish the absorption depth; in reality cloud cover does not completely prevent the light from going through the lower atmosphere. Uncertainty can also come from the difficulty in determining the continuum level because the absorption bands we are considering are surrounded by other absorption features. The variation pattern, which is of our primary interest in this paper, is less sensitive to those uncertainties. The absorption depths normalized so that the daily average is 0 and the standard deviation is unity are exhibited in the middle and right panels of Figure 5. In most cases, the overall variation patterns of both H2O and O2 are well re- produced by our opaque cloud model. In particular, the equatorial data (Earth1, Earth5) exhibit striking agree- ment with the model predictions. For polar observations, the agreement is somewhat degraded, especially for H2O variations in the Polar2 data. This may be partly as- cribed to the fact that there is a slight mismatch between the time of the EPOXI observation and that of the in- put atmospheric data of our simulation; the local time of MODIS observation is 10:30am for Terra and 12:10pm for Aqua, while the disk-integrated spectra observed by EPOXI reflects the information of slices with different local times. In addition, the assumed vertical profile for H2O (the US standard atmosphere) is not expected to be as accurate for the equatorial/polar regions. While our model is admittedly rough in this regard, it is encour- aging that such a simple model reproduces the general trends and features observed by EPOXI fairly well. 4. DIFFERENT VARIATION PATTERNS OF H2O AND O2 ABSORPTION DEPTHS We now focus on the intriguing differences in behavior between H2O and O2. In the Earth1 and Earth5 data, for instance, note the small deviation of the H2O pat- tern from the O2 pattern, including the sharper bump of wO2 at t = 18[hr] than wH2O. Polar1 observations show the variation of wH2O rising at 11 ≤ t[hr] ≤ 19, which is not present in the variation of wO2. In addi- tion, only wO2 has a local peak at t = 6. Significantly and reassuringly, these divergences are reproduced by our simulations, which allows us to isolate the origin of this difference. Let us discuss the possible causes of the different di- urnal pattern of H2O and O2 bands. Given that the ab- sorption depth in our model is directly linked to the total number of molecules above the cloud layers, the different behaviors of H2O and O2 exhibited in Figure 5 can only be ascribed to the differences in their column number density distribution in the planetary atmosphere. In the case of the Earth, the distribution of water vapor differs from that of O2 in at least two ways. Firstly, the total column density of atmospheric water vapor is, un- like O2, a strong function of latitude (highly concentrated in the equatorial region) with additional local/temporal fluctuations. Secondly, the vertical profile of water vapor mixing ratio typically decreases as a function of altitude due to condensation, in contrast to the constant mix- ing ratio (volume fraction of the molecule) of O2 in the troposphere (Figure 6). In order to evaluate these effects quantitatively, we run simulations of H2O band variation with different assump- tions for its atmospheric distribution. We consider two model distributions in addition to our fiducial model de- scribed in Section 3.1: Model A adopts the water vapor vertical profile of the US Standard Model everywhere, and thus the mixing ratio decreases with altitude (red solid line in Figure 6). Model B instead assumes a con- stant mixing ratio of H2O in the troposphere as is the case for O2 (blue dashed line in Figure 6). Both Models A and B neglect the inhomogeneity of water vapor over the surface of the planet. These three simulations and observations in the Po- lar1:north case are plotted in Figure 7. The observed variation of H2O agrees well with fiducial model, while the other two models that neglect the horizontal or sur- face inhomogeneity and/or peculiar vertical profiles fail to match the observed data. We also note that the Model B (blue line in Figure 7) prediction of wH2O is very close to that of the simulation for O2 in the Polar1 case (solid line in the third panel of O2 in Figure 5). This indeed implies that that the differences in variation patterns ba- sically reflect their differing spatial distributions in the atmosphere. Both of the above-mentioned properties of water vapor distribution are consequences of the coexistence of mul- tiple phases of water on Earth. The horizontal inhomo- geneity of water vapor originates from the temperature gradient over the surface (since the vaporization rate is ) m 4m 2 . 1 ( f e R 0.35 0.3 0.25 0.2 0.15 0.4 0.35 0.3 0.25 0.2 0.45 0.4 0.35 0.3 0.25 0.45 0.4 0.35 0.3 Reflectivity at 1.24m m Earth1:equinox Obs. Model Earth5:solstice Polar1:north Polar2:south Water Vapor Oxygen 7 Earth1:equinox Earth5:solstice Obs. Model Polar1:north Polar2:south O 2 H w d e z i l a m r o N 2 0 -2 2 0 -2 2 0 -2 2 0 -2 Earth1:equinox Earth5:solstice Obs. Model Polar1:north Polar2:south 2 O w d e z i l a m r o N 2 0 -2 2 0 -2 2 0 -2 2 0 -2 0 6 12 18 24 0 6 12 18 24 0 6 12 18 24 Time [hr] Time [hr] Time [hr] Fig. 5. -- Comparison between the EPOXI data (symbols) and simulations (dashed lines) in terms of continuum level measured at 1.24µm (left), water vapor absorption (middle), and oxygen absorption (right). ] m k [ e d u t i t l A 100 80 60 40 20 0 H2O CO2 O2 Polar1:north Obs. Fiducial Model A Model B O 2 H w d e z i l a m r o N 2 1 0 -1 -2 1e+00 1e+02 1e+04 1e+06 Mixing Ratio [vppm] 0 6 12 18 24 Time [hr] Fig. 6. -- Vertical profiles of H2O (red solid) and O2 (blue dashed) taken from US standard model. basically determined by the surface temperature) and/or the inhomogeneous distribution of liquid/solid reservoirs on the surface. The water vapor distribution is not mixed well since the mean residence time (MRT) of atmospheric H2O (∼10 days) is much shorter than the time scale of the vertical and global mixing times of the atmosphere (both ∼ 1 year). In contrast, the MRT of O2 is ∼4000 years, thus insuring that O2 is well mixed and only very slowly changing, if at all. The short MRT of water vapor implies that there is a non-negligible flux of H2O into/out of atmosphere and the existence of a substantial water reservoir (compared to the abundance in the atmosphere), such as the oceans on the Earth. Figure 7 also indicates that the difference in the vertical profile alone may lead to differences in the variation of absorption bands, but this is also related to the condensation of water in the atmosphere. In either Fig. 7. -- Comparison in H2O band variation between obser- vation and simulations with different assumption for water vapor distribution. Model A adopts the water vapor vertical profile of US standard model but neglect the horizontal inhomogeneity. Model B assumes constant mixing ratio in the troposphere as is the case for O2. case, the deviation of absorption features of H2O from the well-mixed gas like O2 may serve as an indicator of phases transitions happening in the atmosphere or on the surface of a planet. This signature of the Earth's "wa- ter cycle" could be an observable indicator of a similar climate on a terrestrial exoplanet. 5. SUMMARY AND DISCUSSION In this paper we examined the diurnal variability of the absorption depths of the two important molecules, H2O and O2, in the disk-integrated spectra of the Earth. We analyzed the H2O band centered at ∼ 1.13µm and O2 band centered at ∼ 1.27µm observed by EPOXI and found that these absorption bands show diurnal vari- ations correlated with uneven cloud cover. A simple 8 opaque cloud model, which assumes that the cloud com- pletely blocks the atmospheric signatures below the cloud top layer, is able to reproduce the basic variation patterns of the H2O and O2 bands using pixel-level cloud data ob- tained with Earth Observing Satellite. Thus we conclude that the non-uniform cloud cover distribution dominates the observed diurnal variations of those molecular ab- sorption depths. However, we also found that the diurnal variability pat- terns of H2O and O2 bands are not identical, and the dif- ferences originate from the inhomogeneous distribution of water vapor in the atmosphere. The variability pat- tern of O2 is basically explained by the the cloud cover distribution because it is well mixed in the troposphere of the Earth. In contrast, the variability pattern of H2O is well reproduced only with additional information on the vertical profile and spatial inhomogeneity of atmo- spheric water vapor as a model input. The nature of the water vapor distribution in the atmosphere is linked to the fact that H2O circulates in the planetary surface layer changing its phase among water vapor, liquid wa- ter (ocean/lake/pond/river) and/or water ice. There- fore, different behavior in the variability patterns of O2 and H2O absorptions may carry information on the in- homogeneous phase distribution of H2O in the surface layer. Our study of the Earth demonstrates the possible role of the variability of absorption bands in characterizing the surface environments of Earth-analogs. If future in- struments eventually succeed in detecting scattered light from such exoplanets, their time-resolved measurements of the absorption bands will reveal the uneven cloud cover and/or inhomogeneous spatial distribution of at- mospheric constituents. Our current results suggest that the comparison in variation pattern between presumably well-mixed gases (such as O2) and H2O may indicate frequent phase transitions of water in the surface layer and thus serve as a probe of habitability, in a comple- mentary fashion to the direct detection of liquid wa- ter (e.g. Williams & Gaidos 2008; Oakley & Cash 2009; Robinson et al. 2010; Zugger et al. 2010) and the search for other potential biosignatures (e.g. Kaltenegger et al. 2010; Kawahara et al. 2012). While we focused on absorption bands in the NIR range where the EPOXI data are available, it is natural to expect that other bands also exhibit similar variation patterns at wavelengths dominated by scattering of the primary star's light, rather than planetary thermal emis- sion. In the visible range, there are O2 bands at 0.69µm (equivalent width ∼13.5A) and at 0.76µm (equivalent width ∼47.8A) (e.g. Pall´e et al. 2009) as well as several H2O bands. Inhomogeneity of an atmospheric constituent becomes appreciable when its MRT is short compared to the time- scale of the atmosphere mixing. This is equivalent to there being a substantial flux of the molecular species in question between the atmosphere and some surface reservoir, i.e., substantial compared to the total amount residing in the atmosphere. Besides vaporization, photo- chemical production (e.g. O3) or biotic production (e.g. N2O, CH4) could lead to such a situation. If sufficiently accurate data become available in the future, we may be able to interpret implied inhomogeneous distributions in terms of those or other specific mechanisms. We thank an anonymous referee for constructive com- ments. We are grateful to Timothy A. Livengood for his kind assistance in obtaining EPOXI data. We ac- knowledge support from the Global Collaborative Re- search Fund (GCRF) "A World-wide Investigation of Other Worlds" grant and the Global Scholars Program of Princeton University. Y.F. is supported by JSPS (Japan Society for the Promotion of Science) Fellowship for Re- search, DC:23-6070. The work of Y.S. is supported in part from the Grant-in-Aid Nos. 20340041 and 24340035 by JSPS. E.L.T. was supported in part by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. REFERENCES Abe, Y., Abe-Ouchi, A., Sleep, N. H., & Zahnle, K. J. 2011, Astrobiology, 11, 443 Blake, C. H., & Shaw, M. M. 2011, PASP, 123, 1302 Cowan, N. B., et al. 2009, ApJ, 700, 915 Cowan, N. B., et al. 2011, ApJ, 731, 76 Dorothy, K. H., George A. Riggs, A. G., & Salomonson, V. V. 2006, updated daily, MODIS/Terra Aerosol Cloud Water Vapor Ozone Daily L3 Global 1Deg CMG, Boulder, Colorado USA: National Snow and Ice Data Center. Digital media. Ford, E. B., Seager, S., & Turner, E. L. 2001, Nature, 412, 885 Fujii, Y., & Kawahara, H. 2012, arXiv:1204.3504 Fujii, Y., Kawahara, H., Suto, Y., Fukuda, S., Nakajima, T., Livengood, T. A., & Turner, E. L. 2011, ApJ, 738, 184 Fujii, Y., Kawahara, H., Suto, Y., Taruya, A., Fukuda, S., Nakajima, T., & Turner, E. L. 2010, ApJ, 715, 866 G´omez-Leal, I., Pall´e, E., & Selsis, F. 2012, ApJ, 752, 28 Harrison, A. W., & Coombes, C. A. 1988, Solar Energy, 41, 387 Kaltenegger, L., et al. 2010, Astrobiology, 10, 89 Kawahara, H., & Fujii, Y. 2010, ApJ, 720, 1333 Kawahara, H., & Fujii, Y. 2011, ApJ, 739, L62 Kawahara, H., Matsuo, T., Takami, M., Fujii, Y., Kotani, T., Murakami, N., Tamura, M., & Guyon, O. 2012, arXiv:1206.0558 Lester, T. P., McCall, M. L., & Tatum, J. B. 1979, JRASC, 73, 233 Levine, M., et al. 2009, ArXiv e-prints Liou, K. N. 1980, New York, NY (USA): Academic Press Livengood, T. A., et al. 2011, Astrobiology, 11, 907 Matsuo, T., & Tamura, M. 2010, in the Society of Photo- Optical Instrumentation Engineers (SPIE) Conference, Vol. 7735, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Mayer, B., & Kylling, A. 2005, Atmospheric Chemistry & Physics, 5, 1855 Oakley, P. H. H., & Cash, W. 2009, ApJ, 700, 1428 Pall´e, E., Zapatero Osorio, M. R., Barrena, R., Montan´es- Rodr´ıguez, P., & Mart´ın, E. L. 2009, Nature, 459, 814 Palmer, K. F., & Williams, D. 1974, Journal of the Optical Society of America (1917-1983), 64, 1107 Robinson, T. D., Meadows, V. S., & Crisp, D. 2010, ApJ, 721, L67 Robinson, T. D., et al. 2011, Astrobiology, 11, 393 Sanrom´a, E., & Pall´e, E. 2012, ApJ, 744, 188 Savransky, D., Spergel, D. N., Kasdin, N. J., Cady, E. J., Lisman, P. D., Pravdo, S. H., Shaklan, S. B., & Fujii, Y. 2010, in the Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol. 7731, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Tinetti, G., Meadows, V. S., Crisp, D., Fong, W., Fishbein, E., Turnbull, M., & Bibring, J. 2006a, Astrobiology, 6, 34 Tinetti, G., Meadows, V. S., Crisp, D., Kiang, N. Y., Kahn, B. H., Fishbein, E., Velusamy, T., & Turnbull, M. 2006b, Astrobiology, 6, 881 Williams, D. M., & Gaidos, E. 2008, Icarus, 195, 927 9 Zugger, M. E., Kasting, J. F., Williams, D. M., Kane, T. J., & Philbrick, C. R. 2010, ApJ, 723, 1168 APPENDIX VARIATION OF CO2 We also examine the variability of CO2, another absorption feature imprinted in the observed spectra. Absorption bands of CO2 exist around 1.6µm and 2µm (Figure 1). These CO2 bands override the wings of H2O absorption, making it harder to determine the continuum level. We pick up the absorption band at 1.59-1.62µm, which is likely to be least affected by H2O absorption. The peak-to-throat variation amplitude of four observations are 5-20%12(Table 3), on the same level of O2 and H2O. Figure 8 shows the diurnal variation patterns of CO2 band at 1.59-1.62µm extracted from EPOXI data as well as those of our simulation following the same procedure as H2O/O2 and assuming the uniform distribution (the assumed growth curve of this band is displayed in Figure 4). The normalization factors are summarized in Table 4. We see the clear similarity in variation patterns to O2 rather than H2O. For instance, the clearer bumps at t = 18 of Earth1 and Earth5 observations and the gradient in Polar2 observation. This result is consistent with our discussion in Section 4, since the MRT of CO2 is 3-5 years and slightly longer than the tropospheric mixing timescale. The behavior of Polar1 is less conclusive and the observation does not match the simulation so well . Although we do not identify the cause, it might be related to some level of inhomogeneity of CO2 due to the intermediate MRT. Diurnal Variation of Continuum Level, Equivalent Width of CO2 at 1.6µm. TABLE 3 CO2 at 1.6µm ave.[A] (max-min)/ave Earth1:equinox (Mar.2008) Earth5:solstice (Jun.2008) Polar1:north (Mar.2009) Polar2:south (Oct.2009) 12.1 17.8 17.4 17.9 80.1% 17.4% 11.3% 5.4% Carbon Dioxide (1.59-1.62m m) Earth1:equinox Obs Model Earth5:solstice Polar1:north Polar2:south 2 O C w d e z i l a m r o N 2 0 -2 2 0 -2 2 0 -2 2 0 -2 0 6 12 18 24 Time [hr] Fig. 8. -- Diurnal Variations of CO2 absorption bands around 1.59-1.62 µm. EPOXI data (symbols) and simulations (dashed lines) are compared. 12 The large peak-to-throat amplitude of Earth1:equinox is at least partly due to the failure in continuum determination. In Earth1:equinox only, the observed wavelength grid changes for each exposure and the sharpness of the very narrow feature at 1.58µm, which is critical to determine the continuum level, also changes. This results in the uncertainty in estimation of equivalent width. 10 Normalization Parameters and Comparison in Daily Average and Standard Deviation between Simulations and TABLE 4 Observations. Earth1 Earth5 Polar1 Polar2 ¯wCO2,model/ ¯wCO2,obs σCO2,model/σCO2,obs 1.14 0.23 0.85 1.23 0.95 0.74 0.90 1.75
1302.4799
2
1302
2013-04-08T22:12:39
Two Beyond-Primitive Extrasolar Planetesimals
[ "astro-ph.EP" ]
Using the Cosmic Origins Spectrograph onboard the Hubble Space Telescope, we have obtained high-resolution ultraviolet observations of GD 362 and PG 1225-079, two helium-dominated, externally-polluted white dwarfs. We determined or placed useful upper limits on the abundances of two key volatile elements, carbon and sulfur, in both stars; we also constrained the zinc abundance in PG 1225-079. In combination with previous optical data, we find strong evidence that each of these two white dwarfs has accreted a parent body that has evolved beyond primitive nebular condensation. The planetesimal accreted onto GD 362 had a bulk composition roughly similar to that of a mesosiderite meteorite based on a reduced chi-squared comparison with solar system objects; however, additional material is required to fully reproduce the observed mid-infrared spectrum for GD 362. No single meteorite can reproduce the unique abundance pattern observed in PG 1225-079; the best fit model requires a blend of ureilite and mesosiderite material. From a compiled sample of 9 well-studied polluted white dwarfs, we find evidence for both primitive planetesimals, which are a direct product from nebular condensation, as well as beyond-primitive planetesimals, whose final compositions were mainly determined by post-nebular processing.
astro-ph.EP
astro-ph
Two Beyond-Primitive Extrasolar Planetesimals S. Xu(许偲艺)a, M. Juraa, B. Kleina, D. Koesterb, B. Zuckermana ABSTRACT Using the Cosmic Origins Spectrograph onboard the Hubble Space Telescope, we have obtained high-resolution ultraviolet observations of GD 362 and PG 1225- 079, two helium-dominated, externally-polluted white dwarfs. We determined or placed useful upper limits on the abundances of two key volatile elements, car- bon and sulfur, in both stars; we also constrained the zinc abundance in PG 1225-079. In combination with previous optical data, we find strong evidence that each of these two white dwarfs has accreted a parent body that has evolved beyond primitive nebular condensation. The planetesimal accreted onto GD 362 had a bulk composition roughly similar to that of a mesosiderite meteorite based on a reduced chi-squared comparison with solar system objects; however, addi- tional material is required to fully reproduce the observed mid-infrared spectrum for GD 362. No single meteorite can reproduce the unique abundance pattern observed in PG 1225-079; the best fit model requires a blend of ureilite and mesosiderite material. From a compiled sample of 9 well-studied polluted white dwarfs, we find evidence for both primitive planetesimals, which are a direct product from nebular condensation, as well as beyond-primitive planetesimals, whose final compositions were mainly determined by post-nebular processing. Subject headings: planetary systems stars: abundances white dwarfs 1. INTRODUCTION Planetesimals are building blocks of planets and their formation is a key step towards planet formation. How do planetesimals form? What determines their bulk composition? To answer these questions, we start by examining our own solar system. aDepartment of Physics and Astronomy, University of California, Los Angeles CA 90095-1562; [email protected], [email protected], [email protected], [email protected] bInstitut fur Theoretische Physik und Astrophysik, University of Kiel, 24098 Kiel, Germany; [email protected] -- 2 -- The overall configuration of the solar system is that volatile-depleted, dry rocky objects are ubiquitous relatively close to the Sun while volatile-rich, icy objects are found beyond the snow line. This correlation between the volatile fraction and heliocentric distance can be explained by primitive nebular condensation: refractory elements condensed closer to the Sun while volatile elements can only be incorporated into the planetesimals where the temperature is low enough. Many solar system objects have experienced some additional processing that changed their initial compositions. For example, it has been argued that a collision between a large asteroid and proto-Mercury stripped off most of Mercury's silicate mantle, leaving it ∼70% iron by mass (Benz et al. 1988). Also, the "late veneer" has delivered a large amount of water and volatiles onto Earth (Chyba 1990). Post-nebular processing, such as collisions, melting and differentiation, is important in redistributing the elements among solar system objects. Currently, the best way to measure the elemental compositions of planetesimals in the solar system is from meteorites, which are fragments from collisions among asteroids. Follow- ing O'Neill & Palme (2008), we classify all meteorites into two categories in this paper. (i) "Chondritic" is used to refer to chondrites, which are a direct product of nebular processing. Objects in this category are described as "primitive" planetesimals. (ii) "Non-chondritic" objects consist of achondrites, stoney-iron meteorites and iron meteorites. Examples of their parent bodies include the Moon, Mars or asteroids that have experienced various amounts of post-nebular processing. Planetesimals in this category are considered to be "beyond- primitive". What about planetesimal formation in extrasolar planetary systems? High-resolution, high-sensitivity spectroscopic observations of externally-polluted white dwarfs are a powerful tool for determining the bulk elemental compositions of extrasolar planetesimals (Jura 2013). Calculations show that minor planets can survive the red giant stage of a star and persist into the white dwarf phase with most of their internal water and volatiles intact (Jura 2008; Jura & Xu 2010). Orbital perturbations from one or multiple planets can cause these planetesimals to stray into the tidal radius of the white dwarf and get tidally disrupted (Debes & Sigurdsson 2002; Bonsor et al. 2011; Debes et al. 2012), sometimes producing a dust disk that emits mostly in the infrared (Jura 2003; Kilic et al. 2006; von Hippel et al. 2007; Farihi et al. 2009; Xu & Jura 2012). Eventually, all this planetary debris is accreted onto the central white dwarf and pollutes its otherwise pure hydrogen or helium atmosphere. The first comprehensive abundance measurement of an externally-polluted white dwarf was performed by Zuckerman et al. (2007), who identified 15 elements heavier than helium in the atmosphere of GD 362, including Mg, Si and Fe, which are often called the "com- mon elements" (Larimer 1988). The disrupted object had a minimum mass ∼1022 g, which -- 3 -- is comparable to that of a massive solar system asteroid. Three years later, the abun- dances of eight heavy elements were determined in the atmosphere of GD 40, including all the major rock-forming elements -- O, Mg, Si and Fe (Klein et al. 2010). Now there are many more high-resolution optical spectroscopic studies of externally-polluted white dwarfs [e.g., Klein et al. (2011); Melis et al. (2011); Zuckerman et al. (2011); Farihi et al. (2011); Dufour et al. (2012); Vennes et al. (2010, 2011)]. However, optical spectroscopy of externally-polluted white dwarfs typically does not enable sensitive detection of highly-volatile elements, such as carbon, nitrogen and sulfur, which are key to understanding the thermal history of the system. Ultraviolet spectroscopy is complimentary to optical observations in determination of volatile abundances. To-date, there are four white dwarfs with both published high-resolution optical and ultraviolet measurements1; we are beginning to accumulate an atlas of the compositions of extrasolar planetesimals. To zeroth order, we find that they are strikingly similar to meteorites in the solar system: (i) O, Mg, Si and Fe are always dominant and their sum is more than 85% of the accreted mass; (2) volatile elements, especially C, are typically depleted by more than a factor of 10 compared to solar abundances2. In this paper, we report ultraviolet spectroscopic observations of GD 362 and PG 1225- 079 with the Cosmic Origins Spectrograph (COS) onboard the Hubble Space Telescope (HST), complimentary to previous optical studies from the Keck High Resolution Echelle Spectrom- eter (HIRES) (Zuckerman et al. 2007; Klein et al. 2011). PG 1225-079 has been observed with the low-resolution International Ultraviolet Explorer (IUE) (Wolff et al. 2002); there is no previous ultraviolet spectroscopy for GD 362. The rest of the paper is organized as follows. Data reduction is summarized in section 2 and atmospheric abundance determi- nations are reported in section 3. In section 4, we used a reduced chi-squared analysis to look for solar system analogs to the accreted parent bodies. The formation mechanisms of extrasolar planetesimals are assessed in section 5 and conclusions are given in section 6. In Appendix A, we report the Herschel Photodetecting Array Camera and Spectrometer (PACS) observation of GD 362. In Appendix B, we extend the reduced chi-squared analysis to two additional externally-polluted helium white dwarfs with both high-resolution optical and ultraviolet observations. 1The four white dwarfs are: GD 61 (Desharnais et al. 2008; Farihi et al. 2011); GD 40, G241-6 (Klein et al. 2010, 2011; Zuckerman et al. 2010; Jura et al. 2012) and WD 1929+012 (Vennes et al. 2010, 2011; Melis et al. 2011; Gansicke et al. 2012). 2Very recently, Koester et al. (2012) reported several white dwarfs with solar carbon-to-silicon ratio. However, the source of this pollution is unclear and more analysis is forthcoming. -- 4 -- 2. OBSERVATIONS AND DATA REDUCTION GD 362 and PG 1225-079 were observed during HST/COS Cycle 18 under program 12290. These two white dwarfs are too cool to be observed effectively with the G130M grating centering around 1300 A, as was employed by Jura et al. (2012) and Gansicke et al. (2012) for other hotter white dwarfs. Instead, the G185M grating was used with a central wavelength of 1921 A and wavelength coverage of 1800 -- 1840 A, 1903 -- 1940 A and 2008 -- 2044 A. The spectral resolution was ∼18,000. Total exposure times were 7411 and 1805 sec for GD 362 and PG 1225-079, respectively. The raw data were processed using the standard pipeline CALCOS 2.13.6. The fluxes at 2030 A are 2.9 × 10−15 erg s−1 cm−2 A−1 and 1.5 × 10−14 erg s−1 cm−2 A−1 for GD 362 and PG 1225-079, respectively, in approximate agreement with broadband NUV fluxes from the GALEX satellite. The signal-to-noise ratio (SNR) in the original un-smoothed spectrum was 6 for PG 1225-079 and 4 for GD 362. Following previous data reduction procedures (Klein et al. 2010, 2011; Jura et al. 2012), for PG 1225-079, equivalent widths (EWs) of each spectral line were measured in the un- smoothed spectra by fitting a Voigt profile with three different nearby continuum intervals in IRAF. The EW uncertainty is calculated by adding the standard deviation of the three EWs and the average uncertainty from the profile fitting in quadrature. The EW upper limit is obtained by artificially inserting a spectral line with different abundance into the model and comparing with the data. We adopt a different method to measure the EW for C I 1930.9 A in GD 362, as described in section 3.1. The measured values are listed in Tables 1 and 2 for GD 362 and PG 1225-079, respectively. The average Doppler shift relative to the Sun for PG 1225-079 is 42 ± 13 km s−1, in essential agreement with the value 49 ± 3 km s−1 derived from optical studies (Klein et al. 2011). The large velocity dispersion in the ultraviolet is due to the low SNR of the spectrum and the ∼ 15 km s−1 uncertainty of COS (COS Instrument Handbook). For GD 362, we marginally detected C I 1930.9 A and it has a Doppler shift of 48 km s−1, in agreement with 49.3 ± 1.0 km s−1 from the optical study (Zuckerman et al. 2007). 3. ATMOSPHERIC ABUNDANCE DETERMINATIONS Because we are most interested in the abundance of an element relative to other heavy elements and these ratios are not strongly dependent upon the stellar temperature and surface gravity (Klein et al. 2011), we only adopt one set of stellar parameters as listed in Table 3 and compute the model spectra following Koester (2010). Atomic data are mostly -- 5 -- Table 1: Measured Equivalent Widths and Abundance Determinations for GD 362 Ion λ (A) C I 1930.905 Elow (eV) 1.26 EW (mA) 560 +230 −158 log n(Z)/n(He) a -6.70 ± 0.30 1807.311 1820.341 0 0.049 . 900 . 710 S I S I S . -6.70 . -6.40 . -6.70 a This is measured from the model spectra, as described in section 3.1. taken from the Vienna Atomic Line Database (Kupka et al. 1999). The computed model atmosphere spectra were convolved with the COS NUV line spread function3. The abundance of each element was derived by comparing the EW of each spectral line with the value derived from the model atmosphere, as shown in Figures 1-5 and Tables 1 and 2. The final abundances, combining ultraviolet with optical observations, are given in Tables 4 and 5 for GD 362 and PG 1225-079, respectively. Our results mostly agree with previous reports but have a higher accuracy. For PG 1225-079, we newly derive the abundances of carbon and silicon and have tentative detections of sulfur and zinc. The magnesium abundance is updated while the iron abundance agrees with previous optical results. Because the data are noisier for GD 362, we are only able to crudely constrain the abundance of carbon and sulfur. 3.1. Carbon There is only one useful carbon line in the observed wavelength interval, C I 1930.9 A, as shown in Figures 1 and 2. Because it arises from an excited level, it cannot be contaminated by interstellar absorption. However, this line can be blended with Mn II 1931.4 A. Fortunately, accurate Mn abundances have been determined for both stars from optical data (Zuckerman et al. 2007; Klein et al. 2011) and the predicted EW for Mn II 1931.4 A is less than 50 mA in the model spectrum. Considering the measured EW of this feature is more than 500 mA for both stars (see Tables 1 and 2), we conclude that the line 3http://www.stsci.edu/hst/cos/performance/spectral resolution/nuv model lsf -- 6 -- Table 2: Measured Equivalent Widths and Abundance Determinations for PG 1225-079 Ion λ (A) C I 1930.905 Elow (eV) 1.26 EW (mA) log n(Z)/n(He) 1600 ± 200 -7.80 ± 0.10 1807.311 1820.341 0 0.049 . 170 . 150 S I S I S . -9.50 . -9.30 . -9.50 Mg I 2026.477a 1808.013 1816.928 1925.987 2011.347 2019.429 2021.402 2033.061 2041.345 Si II Si II Si Fe II Fe II Fe II Fe II Fe II Fe II Fe 0 0 0.04 2.52 2.58 1.96 1.67 2.03 1.964 288 ± 100b . -7.60 936 ± 109 1232 ± 145 192 ± 72 309 ± 97 211 ± 67 181 ± 68 311 ± 67 215 ± 50 -7.44 ± 0.10 -7.46 ± 0.10 -7.45 ± 0.10 -7.62 ± 0.28 -7.35 ± 0.24 -7.56 ± 0.24 -7.71 ± 0.27 -7.24 ± 0.17 -7.24 ± 0.18 -7.45 ± 0.23 Zn II 2026.136 0 288 ± 47b . -11.30 a The atomic parameters for this line are taken from Kelleher & Podobedova (2008). b Mg I 2026.5 A and Zn II 2026.1 A are blended and the reported EW is for the entire feature. -- 7 -- Table 3: Adopted Stellar Properties star GD 362 PG 1225-079 M∗ (M⊙) 0.72 0.58 T (K) log g (cm2 s−1) 10,540 10,800 8.24 8.00 D (pc) 51 26 log Mcvz/M∗ a Ref -6.71 -5.02 (1) (2) (3) (4) a Newly-derived mass of the convective zone (see section 4). References.(1) Kilic et al. (2008); (2) Zuckerman et al. (2007); (3) Klein et al. (2011); (4) Farihi et al. (2005). is dominated by C I 1930.9 A. For PG 1225-079, our derived carbon abundance4 [C]/[He] = -7.80 ± 0.10 agrees with the IUE upper limit of -7.5 (Wolff et al. 2002). For GD 362, the largest uncertainty is from the low SNR of the data; the measured continuum flux is (3.1 ± 1.0) × 10−15 erg s−1 cm−2 A−1. It is hard to measure the EW of C I 1930.9 A directly from the noisy data. Instead, we computed model spectra with different carbon abundance to match the observed spectrum. In Figure 1, we present three best-fit models with [C]/[He] = - 6.4, [C]/[He] = -6.7, [C]/[He] = -7.0 and a continuum flux at 4.1 × 10−15 erg s−1 cm−2 A−1, 3.1 × 10−15 erg s−1 cm−2 A−1, 2.1 × 10−15 erg s−1 cm−2 A−1, respectively. The final abundance is [C]/[He] = -6.7 ± 0.3 and the EW reported in Table 1 is measured from the model spectra. 3.2. Sulfur There are two useful sulfur lines, S I 1807.3 A and S I 1820.3 A. However, at best, we have only a tentative detection of sulfur in each star. S I 1807.3 A, the stronger line, is adjacent to Si II 1808.0 A. Fortunately, for GD 362, the silicon abundance is determined from previous optical data (Zuckerman et al. 2007); for PG 1225-079, other ultraviolet lines can be used to derive the silicon abundance (see section 3.4). The data and model atmosphere spectra for GD 362 and PG 1225-079 are presented in Figures 3 and 4, respectively. Considering the apparent match between the model and data for both S I lines, tentative sulfur abundances of -6.7 for GD 362 and -9.5 for PG 1225-079 can be assigned. Conservatively, these results 4Here, log n(X)/n(Y) is abbreviated as [X]/[Y]. -- 8 -- Table 4: Atmospheric Abundances for GD 362 Z log n(Z)/n(He)a -1.14 ± 0.10 -6.70 ± 0.30 < -4.14 < -5.14 -7.79 ± 0.20 -5.98 ± 0.25 -6.40 ± 0.20 -5.84 ± 0.30 . -6.70d -6.24 ± 0.10 -10.19 ± 0.30 -7.95 ± 0.10 -8.74 ± 0.30 -7.41 ± 0.10 -7.47 ± 0.10 -5.65 ± 0.10 -8.50 ± 0.40 -7.07 ± 0.15 -9.20 ± 0.40 -10.42 ± 0.30 H C∗ N O Na Mg Al Si S∗ Ca Sc Ti V Cr Mn Fe Co Ni Cu Sr Total b tset (105 yr) ... 2.1 2.2 2.2 2.2 2.2 1.6 1.2 0.79 0.99 0.93 0.94 0.95 1.0 1.0 1.1 0.99 1.0 0.83 0.56 M(Zi)c (g s−1) ... 2.5 × 107 < 9.0 × 109 < 1.1 × 109 3.7 × 106 2.5 × 108 1.5 × 108 7.2 × 108 . 1.7 × 108 5.1 × 108 6.8 × 104 1.2 × 107 2.1 × 106 4.3 × 107 4.0 × 107 2.5 × 109 4.1 × 106 1.1 × 108 1.1 × 106 1.3 × 105 4.4 × 109 ∗ New measurements from this paper. The rest are from Zuckerman et al. (2007) but we reference abundances relative to He, the dominant element in GD 362's atmosphere, rather than H, as presented in Zuckerman et al. (2007). Consequently, there is a possible systematic offset up to 0.1 dex in all entries derived from that paper. a The final abundance of an element combining optical and ultraviolet data. b Newly-derived settling times in the convective zone (see section 4); they are typically a factor of 2-3 longer than previously-derived values in Koester (2009). c Accretion rates calculated from Equation (1). d The equality sign corresponds to the red model fit shown in figures. -- 9 -- Table 5: Atmospheric Abundances for PG 1225-079 Z log n(Z)/n(He) tset (106 yr) -4.05 ± 0.10 -7.80 ± 0.10 < -5.54 < -8.26 -7.50 ± 0.20 < -7.84 -7.45 ± 0.10 . -9.50 -8.06 ± 0.03 -11.29 ± 0.07 -9.45 ± 0.02 -10.41 ± 0.10 -9.27 ± 0.06 -9.79 ± 0.14 -7.42 ± 0.07 -8.76 ± 0.14 . -11.30 < -11.65 H C∗ O Na Mg∗ Al Si∗ S∗ Ca Sc Ti V Cr Mn Fe Ni Zn∗ Sr Total ... 5.5 4.5 4.4 4.8 3.6 3.0 1.7 1.9 1.8 1.8 1.8 1.9 2.0 2.1 2.3 2.2 1.2 M (Zi) (g s−1) ... 3.1 × 106 < 9.1× 108 < 2.6 × 106 1.4 × 107 < 9.5 × 106 3.0 × 107 . 5.2 × 105 1.6 × 107 1.1 × 104 8.3 × 105 9.6 × 104 1.3 × 106 4.0 × 105 9.0 × 107 4.0 × 106 . 1.3 × 104 < 1.4 × 104 1.6 × 108 ∗ New results from this paper. The rest are from Klein et al. (2011). Notes. The columns are defined the same as Table 4. -- 10 -- C I Fe II 6 5 4 3 2 1 ) 1 − A 2 − m c 1 − s g r e 5 1 − 0 1 ( λ F 0 1928 1929 1930 1931 Mn II 1932 λ(A) 1933 1934 1935 Fig. 1. -- HST/COS spectrum of GD 362. The black line is the data smoothed with a 3 pixel boxcar. The green, red, blue line represents the computed model spectrum with [C]/[He] = -6.4, -6.7, -7.0, respectively, placed at a different continuum level; the abundances of other elements are from Table 4. The adopted carbon abundance is -6.7 ± 0.3. The red labels represent lines that are used for abundance determinations. Wavelength is presented in the star's reference frame in vacuum. ) 1 − A 2 − m c 1 − s g r e 4 1 − 0 1 ( λ F 3 2.5 2 1.5 1 Fe II C I Fe II Al I Mn II 0.5 Mn II Fe II Fe II 0 1920 1922 1924 1926 1928 1932 1934 1936 1938 1940 1930 λ(A) Fig. 2. -- Similar to Figure 1 except for PG 1225-079 with abundances from Table 5. The data were smoothed with a 5 pixel boxcar. -- 11 -- are upper limits. 3.3. Magnesium and Zinc In PG 1225-079, Mg I 2026.4 A and Zn II 2026.1 A are heavily blended. As shown in Figure 5, our best fit model which matches the measured EW of the absorption feature requires [Mg]/[He] = -7.6 and [Zn]/[He] = -11.3. These values are individually taken as upper limits due to the blending. However, the reported magnesium abundance is -7.27 ± 0.06 from the optical data (Klein et al. 2011), which is largely based on three Mg lines but the detections for two lines are only 2σ. Wolff et al. (2002) reported [Mg]/[He] to be -7.6 ± 0.6 from the IUE data. Averaging these measurements, our final magnesium abundance is -7.50 ± 0.20. Because of the blending, the zinc abundance is only an upper limit. This provides the first stringent constraint on zinc in an extrasolar planetesimal. 3.4. Silicon In PG 1225-079, we measured two silicon lines, Si II 1808.0 A and Si II 1816.9 A, as shown in Figure 4. Si II 1808.8 A arises from the ground state and the photospheric line can be distorted by interstellar absorption. However, its measured EW is only 87 ± 11 mA in ζ Oph, a star at a distance of 112 pc with a large amount of foreground interstellar gas (Morton 1975). Considering PG 1225-079 is only 26 pc away, it has much less interstellar absorption. The measured EW is 936 ± 109 mA and we conclude that Si II 1808.0 A is largely photospheric and essentially free from interstellar absorption. The shape of Si II 1808.0 A in the model does not quite fit the data; but the measured EW of the data, which is key in the abundance determination, has a good agreement with that in the model. Using these two Si II lines, we derive a final silicon abundance of -7.45 ± 0.10, in agreement with, but much better than the reported IUE abundance of -7.5 ± 0.5 (Wolff et al. 2002) and the previous optical upper limit of -7.27 (Klein et al. 2011). 3.5. Iron In the COS data for PG 1225-079, there are six Fe II lines with EWs larger than 100 mA. Four of them are shown in Figures 2 and 5. We derived an iron abundance of -7.45 ± 0.23, in good agreement of the optical value of -7.42 ± 0.07, which is based on 28 high- SNR iron lines (Klein et al. 2011). Because the ultraviolet data are noisier, we adopt the -- 12 -- S I Fe II Si I Si II Fe II S I Fe II Si II Fe II 4 3.5 3 2.5 2 1.5 1 0.5 ) 1 − A 2 − m c 1 − s g r e 5 1 − 0 1 ( λ F 0 1805 1810 1815 λ(A) 1820 1825 Fig. 3. -- HST/COS spectrum of GD 362. All notations are the same as Figure 1 and the data are smoothed by a 5 pixel boxcar. S I 1807.3 A and 1820.3 A are used for constraining the sulfur abundance. S I Si II Si II Fe II S I Fe II 2.5 2 1.5 1 0.5 ) 1 − A 2 − m c 1 − s g r e 4 1 − 0 1 ( λ F 0 1805 1810 1815 λ(A) 1820 1825 Fig. 4. -- Similar to Figure 3 except for PG 1225-079. Si II 1808.0 A and 1816.9 A lines are used for determining the silicon abundance. -- 13 -- 2.5 2 1.5 1 0.5 ) 1 − A 2 − m c 1 − s g r e 4 1 − 0 1 ( λ F Fe II Fe II Fe II 0 2015 2020 Mg I Zn II 2025 λ(A) Fe II 2030 2035 Fig. 5. -- HST/COS spectrum of PG 1225-079 and the data are smoothed by a 5 pixel boxcar. All notations are the same as Figure 1. Zn II 2026.1 A and Mg I 2026.5 A are used for determining the Zn and Mg abundances, respectively. The green line presents the best fit model that matches the overall EW of the absorption feature at 2026 A ([Mg]/[He] = -7.6, [Zn]/[He] = -11.3); the red line is the adopted model combining with the optical results ([Mg]/[He] = -7.5, [Zn]/[He] = -11.3). Fe II 2019.4 A, 2021.4 A, 2033.0 A and 1926.0 A in Figure 2 are used for determining the iron abundance. optically-derived iron abundance. 4. COMPARISON WITH SOLAR SYSTEM OBJECTS Combined with previous data, we now have determined the abundances of 16 elements heavier than helium in the atmosphere of GD 362 and 11 heavy elements in PG 1225-079. However, the measured composition need not be identical to the composition of the accreted planetesimal because different elements gravitationally settle at different rates in a white dwarf atmosphere. Three major phases are proposed for a single accretion event: build-up, steady-state and decay (Dupuis et al. 1993; Koester 2009). Because an infrared excess is found for GD 362 and PG 1225-079 (Becklin et al. 2005; Kilic et al. 2005; Farihi et al. 2010), the accretion should be either in the build-up or steady- state phase. The timescale for build-up stage is comparable to the settling times (Koester -- 14 -- 2009); it is ∼ 105 yr, for GD 362 and PG 1225-079 (see Tables 4 and 5). The rest of the disk-host stage should all be under the steady-state approximation. The dust disk lifetime has been under intensive studies for a few years but the values are still very uncertain, including 105 yr (Farihi et al. 2009; Rafikov 2011b), 106 yr (Rafikov 2011a; Girven et al. 2012; Farihi et al. 2012) and up to 107 yr (Barber et al. 2012). The true disk lifetime might have a range but it is likely to be longer than the settling times. Furthermore, Zuckerman et al. (2010) suggested that steady-state approximation is the dominant situation for white dwarf accretion event based on a study of helium dominated stars; the settling times are only 0.1% of their cooling times but 30% of them show atmospheric pollution. GD 362 and PG 1225-079 are more likely to be under the steady-state approximation and that is the main focus of this paper. In the steady-state model, the observed concentration of an element is dependent on the time it takes to sink out of the convective envelope. To derive the theoretical settling times and obtain an improved understanding of the uncertainties, we formulated several numerical experiments with the code for the envelope structure and corrected two errors found in our previous calculations of diffusion timescales. In the course of changing the equations describing element diffusion from the version in Paquette et al. (1986) (Equation 4) to the one in Pelletier et al. (1986) (Equation 5), which is more accurate in the case of electron degeneracy, one of us (D.K.) discovered an error in the former paper. A factor of ρ1/3 is missing in the second alternative of Equation 21, which we had not noticed before. A rederivation of all our equations uncovered another error in our implementation of the contribution of thermal diffusion. These errors have only a very small effect in stars with relatively shallow convection zones, like the hydrogen-dominated white dwarfs. However, for helium-dominated white dwarfs with T < 15,000 K and a deep convection zone, the diffusion timescales can be slower by factors 2-3 relative to our earlier calculations5. The accretion rate M(Zi) of an element Z is calculated as (Koester 2009) M (Zi) = McvzX(Zi) tset(Zi) (1) where Mcvz is the mass of the convective envelope. X(Zi) is the mass fraction of the element Zi relative to the dominant element in the atmosphere, either hydrogen or helium; tset(Zi) is the settling time. A longer settling time corresponds to a lower diffusion flux. Fortunately, the relative timescales for different elements, which are important for the determination of the abundances in the accreted matter, change much less. 5Updated diffusion timescales can be obtained at http://www.astrophysik.uni-kiel.de/∼koester/astrophysics/ -- 15 -- For GD 362 and PG 1225-079, compared to previously published values, the settling times listed in Tables 4 and 5 typically increase by factors of 2-3 while the mass of the convective zone is 0.13 dex smaller for GD 362 and 0.05 dex larger for PG 1225-079 (Table 3). These corrections lead to smaller total accretion rates by a factor of 3 for both stars. The next step is to compare the composition of the accreted parent body with those of solar system objects. We choose the summed mass of all the major elements as the normal- ization factor so that the analysis is independent of the chemical property and abundance uncertainty of each individual element. However, one complication is that no oxygen lines are detected in either GD 362 or PG 1225-079 due to their low photospheric temperatures relative to other helium-dominated white dwarfs; only upper limits were obtained for this major element. Therefore, our approach is to compare the mass fraction of an element rela- tive to the summed mass of the common elements Mg, Si and Fe. For solar system objects, we include 80 representative and well-analyzed meteorite samples mostly from Nittler et al. (2004). We also include the bulk composition of Earth from All`egre et al. (2001) and an updated carbon abundance from Marty (2012). For our purpose, Earth appears to be chon- dritic and its bulk composition approaches CV chondrites even though Earth has experienced some post-nebular processing, such as differentiation and collisions. 4.1. GD 362: Accretion from a Mesosiderite Analog? In Figure 6, we compare the abundances of all 18 elements, including upper limits, of the accreted material in GD 362 with CI chondrites, which are the most primitive material in the solar system. The composition of CI chondrites is almost identical to the solar photosphere, with the exception of depletion of volatile elements C, N as well as H and noble gases. The parent body accreted onto GD 362 looks nothing like a CI chondrite, as first pointed out in Zuckerman et al. (2007). For the volatile elements, the mass fraction of C and S are depleted by at least a factor of 7 and 3, respectively, relative to CI chondrites; refractory elements, such as V, Ca, Ti and Al, are all enhanced. Though oxygen is not detected in GD 362, its stringent upper limit can still provide useful insights. Following Klein et al. (2010), we can calculate the required number of oxygen atoms to form oxides Zp(Z)Oq(Z) as n(O) = XZ q(Z) p(Z) n(Z) (2) Hydrogen is excluded here because GD 362 has an enormous amount and it might not be associated with the parent body or bodies currently in its atmosphere (see Appendix A). -- 16 -- GD362 Steady GD362 Build−up MES C O S Na Cu Mn Cr Si Fe Mg Co Ni V Sr Ca Ti Al Sc ) g M + e F + i S ( M / ) Z ( M g o l I C ) g M + e F + i S ( M / ) Z ( M g o l 1.5 1 0.5 0 −0.5 −1 −1.5 −2 −2.5 Fig. 6. -- Mass fractions of heavy elements in GD 362 with respect to the summed mass of silicon, iron and magnesium from Table 4. The abundances are normalized to those of CI chondrites. The elements are ordered by decreasing volatility. The filled stars represent the steady-state approximation while the open stars are in the build-up stage. 1σ error bars and arrows for upper limits are plotted only for the steady-state approximation for clarity. We also plot the average values for five mesosiderites (Emergy, Barea, Patwar, Dyarrl Island and ALH 77219) from Nittler et al. (2004); 1σ deviations are shown as blue error bars. The error bars are smaller than the symbol when they are not visible in the figure. There are no reported whole rock fractions of Cu, V, Sr and Sc in mesosiderites. Under the steady-state approximation, [O]/[He] = -5.07 is required to form MgO, Al2O3, SiO2 and CaO; this value is comparable to the observed oxygen upper limit of -5.14. However, Fe is the most abundant heavy element in the atmosphere of GD 362 and there is insufficient oxygen to tie it up in either FeO or Fe2O3. Thus, most, if not all the iron in the parent body is in metallic form, which is very different from CI chondrites where most iron is in oxides (Nittler et al. 2004). O'Neill & Palme (2008) suggested that [Mn]/[Na] can be used as an indicator of post- nebular processing. For example, [Mn]/[Na] is -0.79 for all chondrites as well as the solar photosphere while non-chondritic objects have a much higher value. Interestingly, [Mn]/[Na] is 0.65 ± 0.22 for GD 362, which is larger than -0.01 for Mars and 0.32 for the Moon (O'Neill & Palme 2008). This suggests that the planetesimal accreted onto GD 362 is likely to be non-chondritic and have experienced some post-nebular processing. Zuckerman et al. (2007) compared the [Na]/[Ca] ratio in GD 362 with solar system objects and reached a -- 17 -- similar conclusion; the accreted planetesimal was non-chonridtic. The only other polluted white dwarf with both Mn and Na detections is WD J0738+1835 wherein [Mn]/[Na]= -0.54 ± 0.19 (Dufour et al. 2012); this agrees with the chondritic value within the uncertainties. To find the best solar system analog to the parent body accreted onto GD 362, we calculated a reduced chi-squared value for each object in our sample (χ2 red), defined as: χ2 red = 1 N N Xi=1 (Mwd(Zi) − Mmtr(Zi))2 σ2 wd(Zi) (3) where N is the total number of elements considered in the analysis. Mwd(Zi) and Mmtr(Zi) represent the mass fraction of an element Zi relative to the summed mass of Mg, Si and Fe in the extrasolar planetesimal and solar system objects, respectively. σwd(Zi) is the propagated uncertainty in mass fraction. For GD 362, we calculated χ2 red for 11 heavy elements, C, Na, Mg, Al, Si, Ca, Ti, Cr, Mn, Fe and Ni, which have detections both in GD 362 and the meteorite sample6. The results are shown in Figure 7 for both steady-state and build-up approximations. There is no qualitative difference between these two models and mesosiderites provide the best fit considering all 11 elements. In particular, the mesosiderite ALH 77219 can match the overall abundance pattern to 95% confidence level. As shown in Figure 6, the abundance of individual elements agrees within 2σ between mesosiderites and the planetesimal accreted onto GD 362. Mesosiderites are a rare type of stoney-iron meteorite with equal amounts of silicates and metallic iron and nickel. One mystery about mesosiderties is that the Si-rich crust and Fe, Ni-rich core materials are abundant but the olivine Mg-rich mantle seems to be missing. One model for the formation of mesosiderites is that a 200-400 km diameter asteroid with a molten core was nearly catastrophically disrupted by a 50-150 km diameter projectile at 4.42-4.52 Gyr ago (Scott et al. 2001). The collision mixed the target's molten core with its crustal material but excluded the large and hot mantle fragments. The planetesimal accreted onto GD 362 may have been formed in a similar way. While mesosiderites may be a prototype for the accreted planetesimal onto GD 362, there are three major hurdles for this hypothesis to overcome. First, in the model of Scott et al. (2001), only half of the original mass of a 200-400 km diameter asteroid was maintained after the collision and the final product only contains about 10% mesosiderite-like material 6For a couple of meteorites with no reported carbon abundance, we compute the χ 2 red for the other 10 elements. -- 18 -- GD 362 18 16 14 12 10 Carbonaceous Chondrite Ordinary Chondrite Enstatie Chondrite R Chondrite Primitive Achondrite Asteroidal Meteorite Lunar Meteorite Martian Meteorite Pallasite Mesosiderite Bulk Earth 8 6 4 2 0 0 2 4 6 8 10 12 14 16 18 χ2 r ed (Steady) ) p u - d l i u B ( d e 2r χ Fig. 7. -- This figure shows χ2 red values defined in Equation (3), which compares the abun- dances of C, Na, Mg, Al, Si, Ca, Ti, Cr, Mn, Fe and Ni in the accreted material in GD 362 with 80 meteorites and bulk Earth. The x-axis denotes χ2 red calculated for the compo- sition under the steady-state approximation while the y-axis is for the build-up phase. The dashed lines represent 95% confidence level. There are a couple of meteorites with a large χ2 red and they are not shown in the current scale. There is no qualitative difference between the steady-state and build-up models. Mesosiderites, particuarly ALH 77219, provide the best fit to the abundance pattern observed in GD 362. Notes: "Chondritic" materials in- clude: (i) carbonaceous chondrites: CI, CK, CM, CO, CR and CV; (ii) ordinary chondrites: H, L and LL; (iii) enstatite chondrites: EH and EL; (iv) R chondrite. "Non-chondritic" materials consist of achondrites and stoney-iron meteorites. No iron meteorite is included in our analysis because their bulk composition is dominated by Fe and Ni, with very few trace elements reported. Achondrites include: (i) primitive achondrites: acapulcoites, lo- dranites, winonaites and ureilites; (ii) asteroidal meteorites: angrites, aubrites, brachinites and howardite-eucrite-diogenite; (iii) Martian meteorites: shergottie, Nakhlites and Chas- signites; (iv) Lunar meteorites. For stoney-iron meteorites, we include (i) mesosiderites; (ii) pallasites. Most of the meteorite data are from Nittler et al. (2004) and the compositions for some Martian meteorites are from McSween (1985). The bulk composition of Earth is from All`egre et al. (2001) and the carbon abundance is from a more recent study of Marty (2012), which is a factor of 3 lower than the lower limit reported in All`egre et al. (2001). -- 19 -- by mass. This is equivalent to a 75-150 km diameter object. Mesosiderites that fall on Earth are only small fragments and the 180 kg NWA 2924 is among the largest (Meteorite Bulletin Database7). However, the parent body accreted onto GD 362 has a minimum mass of 2.7 × 1022 g, ∼260 km in diameter for an assumed density of 3 g cm−3. It is unclear whether the same kind of collision can produce a mesosiderite parent body this big. Second, the mass fraction of hydrogen in mesosiderites is less than 0.2%; it cannot explain how there is 5 × 1024 g hydrogen in the atmosphere of GD 362. Possibly, hydrogen was accreted during earlier events and it has been atop the atmosphere ever since (see Appendix A for more discussion). Third, GD 362 is currently accreting from its circumstellar disk and the disk material should also resemble the composition of mesosiderites. However, the shape of the mid-infrared spectrum for mesosiderite, which is dominated by a sharp peak at 9.13 µm and several other bands at 10.6 µm and 11.3 µm (Morlok et al. 2012), cannot fully account for the broad 10 µm silicate emission feature observed for GD 362 (Jura et al. 2007). This does not completely exclude the mesosiderite hypothesis but emission from some additional material is required to fully reproduce the observed infrared spectrum for GD 362. Mesosiderites are a good candidate for the parent body accreted onto GD 362 but there are remaining unresolved issues. 4.2. PG 1225-079: Accretion from a Planetesimal with No Single Solar System Analog In Figure 8, we show a comparison of the mass fractions of 16 elements, including upper limits between PG 1225-079 and CI chondrites. Though the carbon abundance is approaching the chondritic value, the accreted planetesimal differs a lot from CI chondrites; the mass fraction of S is depleted by at least a factor of 40 while Zn is depleted by at least a factor of 8. In contrast, refractories, such as V, Ca, Ti and Sc are all enhanced. The overall pattern of relatively high carbon abundance and enhanced mass fractions of refractory elements does not follow a single condensation sequence and post-nebular processing is required. As shown in Figure 9, PG 1225-079 has a [C]/[S] value that is no smaller than the solar ratio, which is very different from other polluted white dwarfs and meteorites. Carbon and sulfur are among the most volatile elements that we can measure and their 50% condensation temperatures are 40 K and 655 K, respectively (Lodders 2003). Most of the meteorites as well as polluted white dwarfs have a [C]/[S] ratio lower than the solar value, which can be explained by condensation at a temperature between 40 and 665 K though this is not 7http://www.lpi.usra.edu/meteor/ -- 20 -- ) g M + e F + i S ( M / ) Z ( M g o l I C ) g M + e F + i S ( M / ) Z ( M g o l ) g M + e F + i S ( M / ) Z ( M g o l I C ) g M + e F + i S ( M / ) Z ( M g o l 1.5 1 0.5 0 −0.5 −1 −1.5 −2 −2.5 1.5 1 0.5 0 −0.5 −1 −1.5 −2 −2.5 PG1225 Steady PG1225 Build−up URE C O S Zn Na Mn Cr Si Fe Mg Ni V Ca Ti Al Sc (a) PG1225 Steady PG1225 Build−up URE+MES C O S Zn Na Mn Cr Si Fe Mg Ni V Ca Ti Al Sc (b) Fig. 8. -- (a) Similar to Figure 6 except for PG 1225-079 with the abundances in Table 5. Ureilites can match the high carbon low sulfur pattern in PG 1225-079 but fail with the other elements. The abundances for ureilites are from Warren et al. (2006). (b) The best-fit model for the composition in PG 1225-069 under the steady-state approximation is 30% by mass ureilite North Haig and 70% mesosiderite Dyarrl Island. There are no reported bulk compositions of Zn, V or Sc for the North Haig and Dyarrl Island meteorites. -- 21 -- necessarily true for all of them. The only solar system analog to PG 1225-079 with similar high carbon, low sulfur pattern is ureilites, a type of primitive achondrites. Ureilites are the second largest achondrite group and it is suggested that its high carbon abundance is derived from a carbon-rich parent body, but the exact formation mechanism is not well understood (Goodrich 1992). However, as can be seen in Figure 8(a), ureilites fail to match the overall composition of the parent body accreted onto PG 1225-079. We performed a χ2 red analysis between solar system objects and the accreted planetesimal in PG 1225-079, comparing 9 elements, C, Mg, Si, Ca, Ti, Cr, Mn, Fe and Ni8. The result is shown in Figure 10. There is no single solar system object that can match all nine elements; the closest is carbonaceous chondrite. Regardless, as shown in Figure 8, the accreted abundance in PG 1225-079 is not at all identical to CI chondrites. The infrared excess around PG 1225-079 corresponds to ∼500 K dust (Farihi et al. 2010); so far, only two white dwarfs are known to have such cool dust. The other 28 known disk-host stars all have ∼1000 K dust (Xu & Jura 2012). One hypothesis is that the inner disk region was recently impacted by another asteroid and all the material was dissipated (Farihi et al. 2010; Jura 2008). If that is the case, PG 1225-079 can be accreting from a blend of two planetesimals, rather than one single parent body. After testing different combinations of the 80 meteorites in our database, the best fit model to the steady-state approximation consists of 30% ureilite North Haig and 70% mesosiderite Dyarrl Island by mass. This blend is also marked in Figure 10. Detailed abundance comparison is shown in Figure 8(b); the abundances of S, Mn and Ca do not agree as well as the other elements but are all within 2σ. A possible scenario is that one extrasolar ureilite (mesosiderite) analog first got tidally disrupted and more recently, another mesosiderite (ureilite) analog impacted the disk and was blended with the previous material. 5. ASSESSING THE FORMATION MECHANISMS OF EXTRASOLAR PLANETESIMALS Having established that the parent bodies accreted onto GD 362 and PG 1225-079 are beyond primitive, we now extend our analysis to other extrasolar planetesimals. We are most interested in understanding the formation mechanisms of extrasolar planetesimals and whether these are dominated by nebular or post-nebular processing. Jura & Xu (2013) suggested collisional rearrangement is important in determining the final composition of 8Similar to the case of GD 362, for the meteorites with no reported carbon abundance, we only calculated 2 red for the other 8 elements. χ -- 22 -- Sun −0.5 −1 −1.5 −2 −2.5 ) e F + i S + g M ( m / ) S ( m g o l PG0843 G241−6 GD362 WD1226 GD40 Earth WD1929 PG1225 −3 −4 Chondritic Non−chondritic 0 log m(C)/m(Mg+Si+Fe) −3 −2 −1 Fig. 9. -- Mass fraction of S and C over the sum of Fe, Mg and Si for solar system objects as well as polluted white dwarfs with positive detection of carbon or sulfur or both. Due to the presence of an infrared excess except for G241-6 (see discussion in Appendix B), all white dwarfs are plotted under the steady-state approximation . 1σ uncertainties are plotted. The red dashed line denotes constant solar [C]/[S]. Most meteorites have a lower [C]/[S] than the solar value. However, [C]/[S] in PG 1225-079 is no smaller than solar and the closest solar system analog is ureilites. References: WD 1929+012, PG 0843+517, WD 1226+110: Gansicke et al. (2012); GD 40, G241-6: Jura et al. (2012); GD 362 and PG 1225-079: this paper; Solar abundance: Lodders (2003); for solar system objects, the references are listed in the caption of Figure 7. -- 23 -- PG1225−079 18 16 14 12 10 8 6 4 2 0 0 2 4 6 ) p u - d l i u B ( d e 2r χ Carbonaceous Chondrite Ordinary Chondrite Enstatie Chondrite R Chondrite Primitive Achondrite Asteroidal Meteorite Lunar Meteorite Martian Meteorite Pallasite Mesosiderite Bulk Earth URE+MES 8 10 12 14 16 18 χ2 r ed (Steady) Fig. 10. -- Similar to Figure 7 except for PG 1225-079 comparing 9 elements, C, Mg, Si, Ca, Ti, Cr, Mn, Fe and Ni. We see that no solar system object comes close to the overall abundance pattern in PG 1225-079. The black triangle represents 30% by mass ureilite and 70% mesosiderite as shown in Figure 8(b), which is the best match for the composition of PG 1225-079 in the steady-state approximation. extrasolar planetesimals based on the scatter in [Mg]/[Ca] ratios in 60 externally-polluted white dwarfs. Here, we compile a sample of well-studied externally-polluted white dwarfs with abundance determinations of at least 9 elements. There are 9 stars in total, as listed in Table 6 and now we assess the formation mechanism for individual objects. GD 40: As discussed in Jura et al. (2012) and Appendix B, the overall abundance pat- tern in GD 40 matches with carbonaceous chondrites and bulk Earth. Nebular condensation is sufficient to explain its observed composition. WD J0738+1835: Dufour et al. (2012) found that there is a correlation between the abundance of an element and its condensation temperature: refractory elements are de- pleted while volatile elements are enhanced compared to bulk Earth. This indicates that the accreted planetesimal might be formed in a low temperature environment under nebular condensation. Table 6: Summary of Planetesimal Formation Mechanisms in 9 Well-Studied White Dwarfs Dom. Dust Volatile Intermediate star GD 40 WD J0738+1835 PG 0843+517 PG 1225-079 NLTT 43806 GD 362 WD 1929+012 G241-6 HS 2253+8023 He He H He H He H He He Y Y Y Y N Y Y N N Cr,Si,Fe,Mg,Co,Ni V,Ca,Ti,Al,Sc Cr,Si,Fe,Mg,Ni Cr,Si,Fe,Mg,Ni Cr,Si,Fe,Mg,Ni Process primitive C,S:,O,Mn, P Cr,Si,Fe,Mg,Ni primitive O,Na,Mn beyond-primitive(?) C,S,O,P beyond-primitive C,Mn Na beyond-primitive C,Na,Cu,Mn Cr,Si,Fe,Mg,Co,Ni V,Sr,Ca,Ti,Al,Sc beyond-primitive C,S,O,Mn,P S,O,Mn,P O,Mn Cr,Si,Fe,Mg,Ni Cr,Si,Fe,Mg,Ni Cr,Si,Fe,Mg,Ni: ??? primitive primitive Ref 1,2 3,4 5 6,7 8 7,9 5,10,11,12 2,6,13 6 -- Refractory Ca,Ti,Al Al V,Ca,Ti,Sc Ca,Ti,Al Ca,Al Ca,Ti Ca,Ti Note. This is a compiled sample of externally polluted white dwarfs with detections of at least 9 elements heavier than helium. Columns are defined as follows. "Dom" lists the dominant element in the atmosphere. "Dust" indicates whether a star has an infrared excess ("Y") or not ("N"). Following the classification scheme in Lodders (2003), "Volatile" lists the detected volatile elements, defined as having a 50% condensation temperature lower than 1290 K in a solar-system composition gas (Lodders 2003); "Intermediate" lists the elements with a condensation temperature between 1290-1360 K -- the same range as that of the common elements, Si, Fe and Mg; "Refractory" elements have a 50% condensation temperature higher than 1360 K. The elements are ordered with increasing condensation temperature. "Process" shows our proposed dominant mechanism that determines the final composition of the accreted extrasolar planetesimal (see section 5). References. (1) Klein et al. (2010); (2) Jura et al. (2012); (3) Dufour et al. (2010); (4) Dufour et al. (2012); (5) Gansicke et al. (2012); (6) Klein et al. (2011); (7) this paper; (8) Zuckerman et al. (2011); (9) Zuckerman et al. (2007); (10) Vennes et al. (2010); (11) Vennes et al. (2011); (12) Melis et al. (2011); (13) Zuckerman et al. (2010). 2 4 -- -- 25 -- PG 0843+517: This star has the highest mass fraction of iron among all polluted white dwarfs. Gansicke et al. (2012) found that all core elements, including Fe, Ni, S and Cr are enhanced relative to the values for bulk Earth while lithophile refractory Al is depleted. This star might be accreting from the core of a differentiated object. Nevertheless, considering the uncertainty for each element is at least 0.2 dex, the conclusion is still preliminary. PG 1225-079: As discussed in section 4.2, this star has a near chondritic carbon abun- dance but also enhanced mass fractions of refractory elements relative to CI chondrite; it cannot be formed solely under nebular processing. NLTT 43806: Compared to chondritic values, the accreted planetesimal is depleted in Fe and enhanced in Al. Zuckerman et al. (2011) found that the best fit model corresponds to "30% crust 70% upper mantle". With detections of 9 elements, evidence is strong that NLTT 43806 has accreted the outer layer of a differentiated parent body. GD 362: As discussed in section 4.1, mesosiderite is the best solar system analog to the accreted parent body and post-nebular processing is required. WD 1929+012: Gansicke et al. (2012) showed that this star has a high iron content. However, the situation is perplexing in that different analyses yield different stellar parame- ters and atmospheric abundances. For example, both Melis et al. (2011) and Gansicke et al. (2012) derived that [Si]/[Fe] is -0.25 but Vennes et al. (2010) found that [Si]/[Fe] is 0.19. No final conclusion can be drawn before resolving such discrepancies. G241-6: This star is a near twin of GD 40 with a similar abundance pattern but without an infrared excess. One possible scenario is that G241-6 has accreted a planetesimal with a similar composition to GD 40 and now it is at the beginning of a decaying phase; all heavier elements appear to be depleted relative to GD 40 due to their short settling times (Klein et al. 2011; Jura et al. 2012). As discussed in Jura et al. (2012) and Appendix B, the overall abundances resemble those of chondrites and no post-nebular processing is required. HS 2253+8023: Klein et al. (2011) showed that the composition of its parent body agrees with bulk Earth, except for the enhanced calcium abundance. Nebular processing can produce the observed abundance pattern. As summarized in Table 6, at least 4 out of the 9 white dwarfs have accreted planetesi- mals that can be formed under nebular processing while post-nebular processing is required for another 3 of them. It should be noted that some objects that we identify as primitive might still have undergone some post-nebular processing. For example, GD 40 has accreted from a planetesimal that has a similar composition as bulk Earth, whose overall abundance pattern is chondritic. However, it is still possible that the parent body was differentiated; -- 26 -- when the entire object is accreted, the composition appears to be "chondritic". We can only put an upper limit on the number of objects formed under nebular condensation. From this sample of 9 stars, we see that post-nebular processing appears to play an important role in determining the final abundance of extrasolar planetesimals; beyond- primitive planetesimals might be as common as primitive planetesimals. In contrast, chon- drites comprise more than 90% of all meteorites found on Earth by number (Meteorite Bulletin Database9). Possibly, extrasolar planetesimals around white dwarfs have violent evolutionary histories with more collisions. This difference is not surprising since dynamical rearrangement of planetary systems at white dwarfs is expected to increase the frequency of collisions and produce more beyond-primitive extrasolar planetesimals. So far, 19 elements heavier than helium, including C, S, O, Na, Cu, Mn, P, Cr, Si, Mg, Fe, Co, Ni, V, Sr, Ca, Ti, Al and Sc, have been detected in the atmospheres of polluted white dwarfs, as shown in Table 6. In terms of mass fraction in the accreted planetesimal, the lowest limit is ∼5 ppm, for Sc in WD J0738+1835 (Dufour et al. 2012). Studying externally-polluted white dwarfs proves to be a very sensitive probe of the bulk compositions of extrasolar planetesimals. 6. CONCLUSIONS We present HST/COS ultraviolet observations for GD 362 and PG 1225-079, two heavily polluted helium white dwarfs. In GD 362, the mass fractions of carbon and sulfur are depleted by at least a factor of 7 and 3 respectively, compared to CI chondrites. In PG 1225-079, a similar volatile depletion pattern is found: C by a factor of 2, S by at least a factor of 40 and Zn by at least a factor of 8. We provide good evidence for the presence of beyond-primitive extrasolar planetesimals: 1. Mesosiderites provide a good match to the composition of the parent body accreted onto GD 362. However, there are several unresolved issues for this hypothesis, espe- cially the apparent difference between the mid-infrared spectrum of mesosiderites and the dust disk around GD 362. Additional material is required. 2. No single meteorite can reproduce the abundance pattern in PG 1225-079. A blend of 30% North Haig ureilite and 70% Dyarrl Island mesosiderite can provide a good fit to the overall composition. 9http://www.lpi.usra.edu/meteor/ -- 27 -- 3. Spectroscopic observations of externally-polluted white dwarfs enable sensitive mea- surement of the bulk compositions of extrasolar planetesimals, including 19 heavy elements down to a mass fraction of 5 ppm. Based on a sample of 9 well-studied white dwarfs, we find that post-nebular processing is as important as nebular condensation in determining the compositions of extrasolar planetesimals. Support for program # 12290 was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. This work also has been partly supported by NSF grants to UCLA to study polluted white dwarfs. APPENDIX A. The Herschel/PACS Observation of GD 362 While hydrogen is detected in some helium-dominated white dwarfs (Voss et al. 2007), GD 362 has an anomalously large amount. The helium-to-hydrogen number ratio is 14 in its convective zone, corresponding to 5 × 1024 g of hydrogen; this is lower than 7 × 1024 g reported in Jura et al. (2009) because the mass of the convective zone for GD 362 is 0.13 dex lower in the updated calculation (Table 3). The origin of the hydrogen is a mystery. Unlike heavy elements which have short settling times compared to the white dwarf cooling age, hydrogen never sinks and can be accumulated over the entire cooling history of the star (Bergeron et al. 2011; Jura & Xu 2012). If GD 362 has always been a helium-dominated white dwarf and all this hydrogen is from accretion of tidally disrupted objects, it can either be one Callisto-size object or ∼100 Ceres-like asteroids (Jura et al. 2009). In the latter case, likely there would be many more asteroids orbiting the star and mutual collisions among them would generate a cloud of cold dust. We were awarded 1.1 hours of Herschel/PACS (Poglitsch et al. 2010) observation time to look for cold dust around GD 362. The "mini-scan map" mode was used to observe in "blue" (85-125 µm) and "red" (125-210 µm) bands simultaneously with a medium scan speed of 20′′ s−1 and a scan leg length of 4′. The scan map size is 345′′ × 374′′ and the repetition number is 25. Two different scan angles, 45 degrees and 135 degrees were used and the total integration time was 1200 sec. Data reduction was performed using HIPE (Herschel Interactive Processing Environ- ment) on a combined mosaic of level 2 products from pipeline SPG 7.1.0. The pixel scale is 1′′ pixel−1 and 2′′ pixel−1 for the blue and red band, respectively. Correcting for its proper -- 28 -- motion, we expect GD 362 at α = 17:31:34.355, δ = +37:05:18.331 on the date of the obser- vation. Because there is no detection, aperture photometry was performed at 25 locations within 5 pixels of the nominal position of GD 362. The aperture radius was 20′′ with a sky annulus between 61′′ and 70′′. The background intensity was estimated using the median sky estimation algorithm (Herschel Data Analysis Guide10). Aperture correction factors are 0.949 for blue and 0.897 for red (PACS Observer's Manual11). Based on the dispersion of the 25 measurements, 3σ upper limits are 5.1 mJy for blue and 5.6 mJy for red. What does this imply about dust mass? GD 362 has shrunk in mass from 3 M⊙ on the main-sequence to its current mass of 0.72 M⊙ (Kilic et al. 2008). Consequently, asteroids initially at 3-5 AU are now orbiting at 13-21 AU. Currently, GD 362 has a stellar temperature of 10,540 K and cooling age ∼ 0.9 Gyr (Farihi et al. 2009). Extrapolating from white dwarf cooling models12 (Bergeron et al. 2011), for GD 362, its stellar temperature is lower than 20,000 K for 90% of its cooling time. We approximate the stellar luminosity as a time- averaged luminosity of 0.01 L⊙. Poynting-Robertson drag was able to remove particles smaller than 20 µm at a distance of 15 AU for a grain density of 3 g cm−3. We therefore assume a dust particle radius of 20 µm in the putative asteroid belt orbiting GD 362. If the grains function as blackbodies with negligible albedo, then their temperature can be calculated as Td = T∗r R∗ 2Dorb (A1) T∗, R∗ are the stellar temperature and radius; Dorb is the orbital distance. The dust temperature is 14-11 K between 13-21 AU. The mass of the dust disk is Md = Fν D2 χBν(T ) ∗ (A2) where D∗ is the distance to GD 362, 51 pc (Kilic et al. 2008) and χ is the dust opacity. For a particle radius of 20 µm, χ = 100 cm2 g−1 in the geometric optics limit. As shown in Figure 11, the upper limit of dust mass is between 1025 g and 1026 g at 13-21 AU; this mass is at 10http://herschel.esac.esa.int/hcss-doc-8.0/print/howtos/howtos.pdf 11http://herschel.esac.esa.int/Docs/PACS/pdf/pacs om.pdf 12http://www.astro.umontreal.ca/ bergeron/CoolingModels/ -- 29 -- least twice the hydrogen mass in the atmosphere of GD 362 and one order of magnitude larger than the mass of solar system's asteroid belt (Krasinsky et al. 2002). The upper limit is not stringent enough to rule out the hypothesis that hydrogen in GD 362 is from accretion of multiple asteroids. So, the large hydrogen abundance in GD 362 remains an unsolved puzzle. 100 µm 160 µm 30 29 28 27 26 25 ) g ( d M g o l 24 5 10 20 15 Dorb (AU) 25 30 Fig. 11. -- Upper limit for the mass of cold dust around GD 362 derived from PACS blue and red data, as a function of orbital radius. Given the hydrogen mass of 5 × 1024 g, the upper limit of dust mass 1025-1026 g at 13-21 AU is not stringent enough to rule out the accretion of multiple asteroids. B. Looking for Solar System Analogs to Extrasolar Planetesimals The χ2 red analysis has proven to be an effective way to look for solar system analogs to the compositions of extrasolar planetesimals. Two other helium-dominated white dwarfs have reported volatile and refractory abundances from high-resolution optical and ultraviolet -- 30 -- observations that are suitable for this kind of analysis13 -- GD 40 and G241-6. Updated settling times and accretion rates are listed in Table 7 while the mass of the convective zone stays the same. Since all the major elements are determined, we compare the mass fraction of an element relative to the sum of O, Mg, Si and Fe. Table 7: Updated Settling Times and Accretion Rates for GD 40 and G241-6 a tset (106 yr) 1.1 1.1 1.1 1.2 1.2 1.0 0.79 0.64 0.51 0.51 0.49 0.53 0.53 0.56 0.61 0.58 0.50 0.43 Z C N O Mg Al Si P S Cl Ca Ti Cr Mn Fe Ni Cu Ga Ge Total (g s−1) M (Z)G241−6 M (Z)GD40 (g s−1) 2.2 × 106 < 4.4 × 105 < 2.6 × 105 < 2.1 × 105 4.3 × 108 4.5 × 108 1.7 × 108 1.5 × 108 < 6.1 × 106 1.4 × 107 8.7 × 107 1.3 × 108 4.7 × 105 1.1 × 106 1.0 × 107: 5.6 × 107 < 8.0 × 105 < 5.8 × 105 5.1 × 107 1.3 × 108 3.2 × 106 1.4 × 106 4.5 × 106 6.4 × 106 2.4 × 106 3.1 × 106 4.4 × 108 2.0 × 108 1.8 × 107 8.9 × 106 < 1.8 × 105 < 1.8 × 105 < 2.9 × 104 < 2.9 × 104 < 1.4 × 105 < 1.4 × 105 9.9 × 108 1.4 × 109 a This column is for both GD 40 and G241-6 because their atmospheric conditions are similar. The total accretion rate for GD 40 is a factor of 2 lower than the value derived in Klein et al. (2010), but the relative abundances change much less. The result of a χ2 red analysis is presented in Figure 12. When including all 13 detected elements, both carbona- ceous chondrites and bulk Earth can match the composition to 95% confidence level for 13GD 61 also has high-resolution optical and ultraviolet observations (Desharnais et al. 2008; Farihi et al. red analysis. 2011). However, with a total of 5 detected elements, it is hard to make a comparison using the χ 2 -- 31 -- both steady-state and build-up approximations. The accreted planetesimal appears to be primitive and can be formed under nebular condensation, similar to what was concluded by Jura et al. (2012). 10 ) p u - d l i u B ( d e 2r χ 9 8 7 6 5 4 3 2 1 0 0 GD 40 Carbonaceous Chondrite Ordinary Chondrite Enstatie Chondrite R Chondrite Primitive Achondrite Asteroidal Meteorite Lunar Meteorite Martian Meteorite Pallasite Mesosiderite Bulk Earth 2 4 χ2 r ed (Steady) 6 8 10 Fig. 12. -- Similar to Figure 7 except for GD 40 comparing the mass fraction of 13 elements, including C, O, Mg, Al, Si, P, S, Ca, Ti, Cr, Mn, Fe and Ni, relative to the sum of Mg, Si, Fe and O. Both carbonaceous chondrites and bulk Earth are a good match to the parent body accreted onto GD 40. The newly-derived total accretion rate for G241-6 is about a factor of 2 lower than previously reported (Zuckerman et al. 2010). The non-detection of an infrared excess and the slight depletion of heavier elements suggest that it may be at the beginning of a decay phase (Xu & Jura 2012; Klein et al. 2011). We assess both steady-state and decay phase for the χ2 red analysis; in the latter case, we assume that accretion stopped 0.6 × 106 yr ago, approximately one settling time for Fe because its mass fraction is depleted by a factor of 2 relative to CI chondrites. The composition of the parent body is calculated following Zuckerman et al. (2011) and Equation (5) in Koester (2009). A fuller exploration of different time-varying models will be presented in the future in the spirit of Jura & Xu (2012). As shown in Figure 13, both carbonaceous chondrites and ordinary chondrites provide good matches to all 11 elements, including O, Mg, Si, P, S, Ca, Ti, Cr, Mn, Fe and Ni. However, the -- 32 -- carbon upper limit in G241-6, which is not included in the χ2 red analysis, is at least one order of magnitude lower than most carbonaceous chondrites (Jura et al. 2012). Thus, ordinary chondrites are a more promising solar system analog to the parent body accreted onto G241-6 and nebular condensation is sufficient to produce the observed abundance pattern. The χ2 red analysis for GD 40 and G241-6 confirms the previous results (Jura et al. 2012); the accreted extrasolar planetesimals can be formed under nebular condensation and their compositions resemble primitive chondrites in the solar system. 4 3.5 3 2.5 2 ) y a c e D ( d e 2r χ 1.5 1 0.5 0 0 G241−6 0.5 1 2 1.5 χ2 2.5 r ed (Steady) Carbonaceous Chondrite Ordinary Chondrite Enstatie Chondrite R Chondrite Primitive Achondrite Asteroidal Meteorite Lunar Meteorite Martian Meteorite Pallasite Mesosiderite Bulk Earth 3 3.5 4 Fig. 13. -- Similar to Figure 12 except for G241-6 comparing 11 elements -- O, Mg, Si, P, S, Ca, Ti, Cr, Mn, Fe and Ni for steady-state versus decay phase when the accretion stopped 0.6 × 106 yr ago. Both carbonaceous chondrites and ordinary chondrites produce a good fit to the parent body accreted onto G241-6. Ordinary chondrites are a relatively better match because of the low carbon abundance, which is not considered in this χ2 red plot because only an upper limit was reported for G241-6 (Jura et al. 2012). All`egre, C., Manh`es, G., & Lewin, E. 2001, Earth and Planetary Science Letters, 185, 49 REFERENCES -- 33 -- Barber, S. D., Patterson, A. J., Kilic, M., Leggett, S. K., Dufour, P., Bloom, J. S., & Starr, D. L. 2012, ApJ, 760, 26 Becklin, E. E., Farihi, J., Jura, M., Song, I., Weinberger, A. J., & Zuckerman, B. 2005, ApJ, 632, L119 Benz, W., Slattery, W. L., & Cameron, A. G. W. 1988, Icarus, 74, 516 Bergeron, P., Wesemael, F., Dufour, P., Beauchamp, A., Hunter, C., Saffer, R. A., Gianninas, A., Ruiz, M. T., Limoges, M.-M., Dufour, P., Fontaine, G., & Liebert, J. 2011, ApJ, 737, 28 Bonsor, A., Mustill, A. J., & Wyatt, M. C. 2011, MNRAS, 594 Chyba, C. F. 1990, Nature, 343, 129 Debes, J. H. & Sigurdsson, S. 2002, ApJ, 572, 556 Debes, J. H., Walsh, K. J., & Stark, C. 2012, ApJ, 747, 148 Desharnais, S., Wesemael, F., Chayer, P., Kruk, J. W., & Saffer, R. A. 2008, ApJ, 672, 540 Dufour, P., Kilic, M., Fontaine, G., Bergeron, P., Lachapelle, F., Kleinman, S. J., & Leggett, S. K. 2010, ApJ, 719, 803 Dufour, P., Kilic, M., Fontaine, G., Bergeron, P., Melis, C., & Bochanski, J. 2012, ApJ, 749, 6 Dupuis, J., Fontaine, G., Pelletier, C., & Wesemael, F. 1993, ApJS, 84, 73 Farihi, J., Becklin, E. E., & Zuckerman, B. 2005, ApJS, 161, 394 Farihi, J., Brinkworth, C. S., Gansicke, B. T., Marsh, T. R., Girven, J., Hoard, D. W., Klein, B., & Koester, D. 2011, ApJ, 728, L8 Farihi, J., Gansicke, B. T., Wyatt, M. C., Girven, J., Pringle, J. E., & King, A. R. 2012, MNRAS, 424, 464 Farihi, J., Jura, M., Lee, J., & Zuckerman, B. 2010, ApJ, 714, 1386 Farihi, J., Jura, M., & Zuckerman, B. 2009, ApJ, 694, 805 Gansicke, B. T., Koester, D., Farihi, J., Girven, J., Parsons, S. G., & Breedt, E. 2012, MNRAS, 424, 333 -- 34 -- Girven, J., Brinkworth, C. S., Farihi, J., Gansicke, B. T., Hoard, D. W., Marsh, T. R., & Koester, D. 2012, ApJ, 749, 154 Goodrich, C. A. 1992, Meteoritics, 27, 327 Jura, M. 2003, ApJ, 584, L91 -- . 2008, AJ, 135, 1785 -- . 2013, ArXiv e-prints Jura, M., Farihi, J., Zuckerman, B., & Becklin, E. E. 2007, AJ, 133, 1927 Jura, M., Muno, M. P., Farihi, J., & Zuckerman, B. 2009, ApJ, 699, 1473 Jura, M. & Xu, S. 2010, AJ, 140, 1129 -- . 2012, AJ, 143, 6 -- . 2013, AJ, in press Jura, M., Xu, S., Klein, B., Koester, D., & Zuckerman, B. 2012, ApJ, 750, 69 Kelleher, D. & Podobedova, L. 2008, JPCRD, 37, 267 Kilic, M., Thorstensen, J. R., & Koester, D. 2008, ApJ, 689, L45 Kilic, M., von Hippel, T., Leggett, S. K., & Winget, D. E. 2005, ApJ, 632, L115 -- . 2006, ApJ, 646, 474 Klein, B., Jura, M., Koester, D., & Zuckerman, B. 2011, ApJ, 741, 64 Klein, B., Jura, M., Koester, D., Zuckerman, B., & Melis, C. 2010, ApJ, 709, 950 Koester, D. 2009, A&A, 498, 517 -- . 2010, Mem. Soc. Astron. Italiana, 81, 921 Koester, D., Gansicke, B., Girven, J., & Farihi, J. 2012, ArXiv e-prints 1206:6036 Krasinsky, G. A., Pitjeva, E. V., Vasilyev, M. V., & Yagudina, E. I. 2002, Icarus, 158, 98 Kupka, F., Piskunov, N., Ryabchikova, T. A., Stempels, H. C., & Weiss, W. W. 1999, A&AS, 138, 119 -- 35 -- Larimer, J. W. 1988, Meteorites and the Early Solar System (University of Arizona Press), 19 -- 52 Lodders, K. 2003, ApJ, 591, 1220 Marty, B. 2012, Earth and Planetary Science Letters, 313, 56 McSween, H. Y. 1985, Reviews of Geophysics, 23, 391 Melis, C., Farihi, J., Dufour, P., Zuckerman, B., Burgasser, A. J., Bergeron, P., Bochanski, J., & Simcoe, R. 2011, ApJ, 732, 90 Morlok, A., Koike, C., Tomeoka, K., Mason, A., Lisse, C., Anand, M., & Grady, M. 2012, Icarus, 219, 48 Morton, D. C. 1975, ApJ, 197, 85 Nittler, L. R., McCoy, T. J., Clark, P. E., Murphy, M. E., Trombka, J. I., & Jarosewich, E. 2004, Antarctic Meteorite Research, 17, 231 O'Neill, H. S. C. & Palme, H. 2008, Royal Society of London Philosophical Transactions Series A, 366, 4205 Paquette, C., Pelletier, C., Fontaine, G., & Michaud, G. 1986, ApJS, 61, 197 Pelletier, C., Fontaine, G., Wesemael, F., Michaud, G., & Wegner, G. 1986, ApJ, 307, 242 Poglitsch, A., Waelkens, C., Geis, N., & et al. 2010, A&A, 518, L2 Rafikov, R. R. 2011a, ApJ, 732, L3+ -- . 2011b, MNRAS, 416, L55 Scott, E. R. D., Haack, H., & Love, S. G. 2001, Meteoritics and Planetary Science, 36, 869 Vennes, S., Kawka, A., & N´emeth, P. 2010, MNRAS, 404, L40 -- . 2011, MNRAS, 413, 2545 von Hippel, T., Kuchner, M. J., Kilic, M., Mullally, F., & Reach, W. T. 2007, ApJ, 662, 544 Voss, B., Koester, D., Napiwotzki, R., Christlieb, N., & Reimers, D. 2007, A&A, 470, 1079 Warren, P. H., Ulff-Møller, F., Huber, H., & Kallemeyn, G. W. 2006, Geochim. Cos- mochim. Acta, 70, 2104 -- 36 -- Wolff, B., Koester, D., & Liebert, J. 2002, A&A, 385, 995 Xu, S. & Jura, M. 2012, ApJ, 745, 88 Zuckerman, B., Koester, D., Dufour, P., Melis, C., Klein, B., & Jura, M. 2011, ApJ, 739, 101 Zuckerman, B., Koester, D., Melis, C., Hansen, B. M., & Jura, M. 2007, ApJ, 671, 872 Zuckerman, B., Melis, C., Klein, B., Koester, D., & Jura, M. 2010, ApJ, 722, 725 This preprint was prepared with the AAS LATEX macros v5.2.
1612.01579
1
1612
2016-12-05T22:31:31
Stability of sulphur dimers S2 in cometary ices
[ "astro-ph.EP" ]
S2 has been observed for decades in comets, including comet 67P/Churyumov-Gerasimenko. Despite the fact that this molecule appears ubiquitous in these bodies, the nature of its source remains unknown. In this study, we assume that S2 is formed by irradiation (photolysis and/or radiolysis) of S-bearing molecules embedded in the icy grain precursors of comets, and that the cosmic ray flux simultaneously creates voids in ices within which the produced molecules can accumulate. We investigate the stability of S2 molecules in such cavities, assuming that the surrounding ice is made of H2S or H2O. We show that the stabilization energy of S2 molecules in such voids is close to that of the H2O ice binding energy, implying that they can only leave the icy matrix when this latter sublimates. Because S2 has a short lifetime in the vapor phase, we derive that its formation in grains via irradiation must occur only in low density environments such as the ISM or the upper layers of the protosolar nebula, where the local temperature is extremely low. In the first case, comets would have agglomerated from icy grains that remained pristine when entering the nebula. In the second case, comets would have agglomerated from icy grains condensed in the protosolar nebula and that would have been efficiently irradiated during their turbulent transport towards the upper layers of the disk. Both scenarios are found consistent with the presence of molecular oxygen in comets.
astro-ph.EP
astro-ph
Stability of sulphur dimers (S2) in cometary ices O. Mousis1, O. Ozgurel2, J. I. Lunine3, A. Luspay-Kuti4, T. Ronnet1, F. Pauzat2, A. Markovits2, and Y. Ellinger2 Received ; accepted 1Aix Marseille Universit´e, CNRS, LAM (Laboratoire d'Astrophysique de Marseille) UMR 7326, 13388, Marseille, France [email protected] 2Laboratoire de Chimie Th´eorique, Sorbonne Universit´es, UPMC Univ. Paris 06, CNRS UMR 7616, F-75252 Paris CEDEX 05, France 3Department of Astronomy and Carl Sagan Institute, Space Sciences Building Cornell University, Ithaca, NY 14853, USA 4Department of Space Research, Southwest Research Institute, 6220 Culebra Rd., San Antonio, TX 78228, USA -- 2 -- ABSTRACT S2 has been observed for decades in comets, including comet 67P/Churyumov- Gerasimenko. Despite the fact that this molecule appears ubiquitous in these bodies, the nature of its source remains unknown. In this study, we assume that S2 is formed by irradiation (photolysis and/or radiolysis) of S-bearing molecules embedded in the icy grain precursors of comets, and that the cosmic ray flux simultaneously creates voids in ices within which the produced molecules can ac- cumulate. We investigate the stability of S2 molecules in such cavities, assuming that the surrounding ice is made of H2S or H2O. We show that the stabilization energy of S2 molecules in such voids is close to that of the H2O ice binding energy, implying that they can only leave the icy matrix when this latter sublimates. Be- cause S2 has a short lifetime in the vapor phase, we derive that its formation in grains via irradiation must occur only in low density environments such as the ISM or the upper layers of the protosolar nebula, where the local temperature is extremely low. In the first case, comets would have agglomerated from icy grains that remained pristine when entering the nebula. In the second case, comets would have agglomerated from icy grains condensed in the protosolar nebula and that would have been efficiently irradiated during their turbulent transport to- wards the upper layers of the disk. Both scenarios are found consistent with the presence of molecular oxygen in comets. Subject headings: comets: general -- comets: individual (67P/Churyumov- Gerasimenko) -- solid state: volatile -- methods: numerical -- astrobiology -- 3 -- 1. Introduction The nature of the source of sulphur dimers (S2) observed in comets is still unknown. The first detection of S2 in a celestial body was in the UV spectra of Comet IRAS-Araki-Alcock (C/1983 H1) acquired with the International Ultraviolet Explorer (IUE) space observatory (A'Hearn et al. 1983). Emission bands of S2 were subsequently identified in many comets observed with IUE in the eighties, including 1P/Halley (Krishna Swamy & Wallis 1987). S2 was also identified in comets Hyakutake (C/1996 B2), Lee (C/1999 H1), and Ikeya-Zhang (C/2002 C1) (Laffont et al. 1998; Kim et al. 2003; Boice & Reyl´e 2005). More recently, S2 has been detected in comet 67P/Churyumov-Gerasimenko (hereafter 67P/C-G) by the ROSINA mass spectrometer aboard the Rosetta spacecraft at a distance of ∼3 AU from the Sun in October 2014 (∼4 -- 13 ×10−6 with respect to water; Le Roy et al. 2015; Calmonte et al. 2016). All these observations suggest that S2 is ubiquitous in comets. Because the lifetime of S2 is very short in comae (∼a few hundreds seconds at most; Reyl´e & Boice 2003), two main scenarios have been invoked in the literature to account for its presence in comets. In the first scenario, S2 is the product of reactions occurring in the coma. Ethylene was thus proposed to act as a catalyst allowing the formation of S2 molecules in the inner coma (Saxena & Misra 1995; Saxena et al. 2003). Also, the presence of atomic S (as the photodissociation product of CS2) reacting with OCS was suggested to form S2 in comae (A'Hearn et al. 2000). However, models depicting the chemistry occurring in cometary comae show that these two mechanisms do not account for the observed levels of S2 (Rogers & Charnley 2006). In the second scenario, S2 molecules are believed to be of parent nature and reside in cometary ices (A'Hearn et al. 1983; A'Hearn & Feldman 1985; Grim & Greenberg 1987; Feldman 1987; A'Hearn 1992). A'Hearn & Feldman (1985) proposed that the UV photolysis of S-bearing species embedded in ISM ices could form sufficient amounts of S2 -- 4 -- that remains trapped in the icy matrix. Since then, a number of mechanisms based on UV or X-ray irradiation have been proposed, starting mainly from H2S (the most abundant S-bearing volatile observed in comets; Irvine et al. 2000; Bockel´ee-Morvan et al. 2004) and H2S2, and involving radicals like HS and HS2 (Grim & Greenberg 1987; Jim´enez-Escobar & Munoz-Caro 2011; Jim´enez-Escobar et al. 2012). It has also been proposed that S2 could be formed from the radiolysis of S -- bearing compounds in cometary ices (A'Hearn & Feldman 1985; Calmonte et al. 2016) despite the fact that so far, there is no experimental proof showing that this mechanism is effective. In the present study, we postulate that S2 is formed from H2S molecules embedded in icy grains by irradiation of UV, X-ray and cosmic ray fluxes (CRF), whether icy grain precursors of comets formed in the protosolar nebula or the ISM. Because radiolysis generated by the impact of cosmic rays simultaneously creates voids in ices within which the produced molecules can accumulate (Carlson et al. 2009; Mousis et al. 2016a), we investigate the stability of S2 molecules in such cavities, assuming that the surrounding ice is made of H2S or H2O. We show that the stabilization energy of S2 molecules in such voids is close to that of the H2O ice binding energy, implying that they can only leave the icy matrix when this latter sublimates. We finally discuss the implications of our results for the origin of cometary grains, with a particular emphasis on those agglomerated by comet 67P/C-G. 2. Irradiation of icy grains Three irradiation mechanisms leading to the formation of S2 are considered in this study. The first two mechanisms, namely UV and X-ray irradiation, have been proven to produce S2 from H2S and H2S2 (Grim & Greenberg 1987; Jim´enez-Escobar & Munoz-Caro 2011; Jim´enez-Escobar et al. 2012). Experiments have shown that S2 can be produced -- 5 -- and stabilized in icy grains over thicknesses of a few tenths of microns. Despite the lack of experimental data, radiolysis has also been considered as a potential candidate for S2 formation from S -- bearing compounds in cometary icy grains (A'Hearn & Feldman 1985). This mechanism has recently been proposed to explain the detection of S2 in 67P/C-G (Calmonte et al. 2016) and is often invoked to account for its presence in Europa's exosphere (Carlson et al. 1999; Cassidy et al. 2010). Cosmic rays reach deeper layers than photon irradiation and simultaneously creates voids in which some irradiation products such as O2 or here S2 can be sequestrated (Mousis et al. 2016a). Whatever the irradiation process considered, we assume that, once S2 has been created and trapped in the microscopic icy grains, the latter agglomerated and formed the building blocks of comets. 3. Stability of S2 molecules in an icy matrix The S2 stabilization energy arises from the electronic interaction between the host support (H2O ice or H2S ice) and the S2 foreign body. The stabilization energy is evaluated as Estab = (Eice + ES2) − E, (1) where ES2 is the energy of the isolated molecule, Eice the energy of the pristine solid host and E is the total energy of the [host + S2] complex, with all entities optimized in isolation. All simulations are carried out by means of the Vienna ab-initio simulation package (VASP) (Kresse & Hafner 1993; 1994; Kresse & Furthmuller 1996; Kresse & Joubert 1999). The long range interactions in the solid and the hydrogen bonding being the critical parameters in the ices, we use the PBE generalized gradient approximation (GGA) functional (Perdew et al 1996), in the (PBE+D2) version corrected by Grimme et al. (2010) -- 6 -- that has been specifically designed to deal with the present type of problem. This theoretical tool has proved to be well adapted to model bulk and surface ice structures interacting with volatile species (Lattelais et al. 2011, 2015; Ellinger et al. 2015; Mousis et al. 2016a). More details on the computational background can be found in the aforementioned publications. Since S2 is created well inside the icy grain mantles, the initial description of the irradiated ice is taken as that of the internal structures of ice clusters obtained from Monte-Carlo simulations of ice aggregates constituted of hundreds of water molecules. The important point in the simulations by Buch et al (2004) is that the core of the aggregates consists in crystalline domains of apolar hexagonal ice Ih. However, in the present context, the irradiation creates significant defects inside the ice, namely, voids and irradiation tracks that, at least locally, modify the crystalline arrangement. 3.1. S2 embedded in H2O ice Because H2O is the dominant volatile in comets (Bockel´ee-Morvan et al. 2004), most of the cavities created by CRF irradiation are expected to be surrounded by H2O molecules. Table 1 shows the stabilization energy of S2 as a function of the size of these cavities. How the S2 stabilization evolves as a function of their size is summarized below. 1. Starting with no H2O removed, we find no stabilization for the inclusion of S2 in the ice lattice. It is in fact an endothermic process, as it is for O2 inclusion (Mousis et al. 2016a). 2. With one H2O removed, we have an inclusion structure whose stabilization is negative, meaning that S2 cannot stay in such a small cavity. 3. With somewhat larger cavities obtained by removing 2 to 4 adjacent H2O molecules from the ice lattice, we obtain increasing stabilization energies from 0.3 to 0.5 eV. -- 7 -- 4. With larger cavities that form along the irradiation track, the stabilization energies are found to be at least of the order of 0.5 eV. In short, as soon as the space available is sufficient, the energy stabilizes around 0.5 eV. This stabilization energy is i) higher (more stabilizing) than what is found in the case of O2 (0.2 -- 0.4 eV; Mousis et al. 2016a) and ii) larger than that of a water dimer (∼0.25 eV). Hence, the presence of S2 should not perturb the ice structure until it is ejected into the coma via sublimation with the surrounding H2O molecules. The results of our computations are consistent with the laboratory experiments of Grim & Greenberg (1987) who showed that S2 remains trapped in icy grains until they are heated up to ∼160 K, a temperature at which water ice sublimates at PSN conditions. Table 1: Computed stabilization energies (eV) of S2 interacting with H2O ice or H2S ice Environment H2O ice H2S ice Adsorption Inclusion (n=1)∗ Inclusion (n=2) Inclusion (n=4) Inclusion (fine track) Inclusion (large track) 0.28 -0.12 0.28 0.50 0.51 0.53 0.30 0.45 0.40 0.41 0.50 ∗n = number of H2O or H2S molecules destroyed to create the void in which S2 is trapped. 3.2. S2 embedded in H2S ice H2S behaves similarly to H2O because of its ability to establish hydrogen bonds. This implies that small domains of H2S could have formed in the bulk of the ice and served as local sources for the formation of S2. The stabilization of these aggregates is addressed by -- 8 -- numerical simulations in which H2S entities are progressively introduced by replacing an equal number of H2O molecules in the water-ice lattice. Table 2 shows the stabilization energies with values around 0.5 and 0.75 eV for neighboring and far away H2S, respectively. Consequently, substituting several neighboring H2O by H2S is a possibility to be considered if the H2S is abundant enough, thus creating small islands of H2S within the water ice. Table 2: Computed stabilization energies (eV) of H2S interacting with H2O ice Environment Adsorption Substitution (n=1)∗ Substitution (n=2 far away) Substitution (n=2 close) Substitution (n=3 close) SH2 Estab 0.61 0.77 0.73 0.56 0.50 Substitution (irradiation track) 0.51 ∗n = number of H2O molecules replaced by H2S. If small clumps of H2S ices in the bulk of water ice are a plausible hypothesis, as suggested by the aforementioned numbers, then the proper conditions are realized for the in situ formation of S2 by deep irradiation. The case in which H2S molecules replace H2O along the irradiation track is a less favorable situation but it could be at the origin of the Sn oligomers observed in some laboratory experiments (Meyer et al. 1972; Jim´enez-Escobar et al. 2012). We evaluate the stabilization of S2 in H2S clumps, assuming that they behave as pure condensates. The results, presented in Table 1 and summarized below, are quite close to those derived for water ice. 1. With one H2S removed, we have a substitution structure whose stabilization is on the order of 0.30 eV. -- 9 -- 2. With larger cavities obtained by removing 2 to 4 adjacent H2S molecules, we obtain increasing stabilization energies between 0.40 and 0.45 eV. 3. With even larger cavities, extended in the direction of the irradiation, the stabilization energies are found to be similar to the preceding ones, between 0.40 and 0.50 eV. Again we find that the presence of S2 should not perturb the ice structure, even when trapped in H2S clumps, until the latter sublimate, due to increasing local temperature. 4. Implications for cometary ices It has been recently shown that the radiolysis of icy grains in low-density environments such as the presolar cloud may induce the production of amounts of molecular oxygen high enough to be consistent with the quantities observed in 67P/C-G (Mousis et al. 2016a). Higher density environments such as the PSN midplane were excluded because the timescales needed to produce enough O2 in cometary grains exceeded by far their lifetimes in the disk. Also, the efficiency of ionization by cosmic rays in the PSN midplane is now questioned because of the deflection of galactic CRF by the stellar winds produced by young stars (Cleeves et al. 2013, 2014). On the other hand, because the lifetime of S2 is very short in the gas phase (∼a few hundred seconds at most; Reyl´e & Boice 2003), its formation conditions are even more restrictive than those required for O2. Assuming that S2 indeed formed from H2S or any other S-bearing molecule via UV, X-ray or CRF irradiation, this implies that this molecule never left the icy matrix in the time interval between its formation and trapping. In other words, S2 never condensed from the PSN before being trapped in cometary grains. This stringent constrain requires S2 to form within icy grains irradiated by CRF in low density environments such as ISM, where the local temperature is extremely cold. In this -- 10 -- picture, comets, including 67P/C-G, would have agglomerated in the PSN from icy grains originating from ISM, whose compositions and structures remained pristine when entering the nebula. Alternatively, because the CRF irradiation should be poorly attenuated in the upper layers of the PSN, these regions also constitute an adequate low-density environment allowing the formation of S2 in cometary grains. Turbulence plays an important role in the motion of small dust grains that are well coupled to the gas. Micron-sized grains initially settled in the midplane are entrained by turbulent eddies and diffuse radially and vertically with an effective viscosity roughly equal to that of the gas for such small particles (see Ciesla 2010, 2011 for details). Consequently, solid particles follow a gaussian distribution in the vertical direction. The scale height of dust (corresponding to the standard deviation of the distribution) is a fraction of the gas scale height, this fraction being larger and possibly equal to the gas scale height in the cases of small grains and higher degrees of turbulence (Dubrulle et al. 1995; Youdin & Lithwick 2007). The vertical transport of solids exposes them to very different disk environments. Dust grains are stochastically transported to high altitude and low-density regions above the disk midplane. Ciesla (2010) developed a numerical simulation to integrate the motion of individual particles and showed that micron-sized grains spent ∼32% of their lifetime at altitudes above the scale height of the disk, including ∼5% at heights above four times its scale height, regardless the distance from the Sun. In such low-density environments, photochemistry plays a primordial role, as demonstrated by Ciesla & Sandford (2012), because UV photons are weakly attenuated at those heights. This also holds for the CRF irradiation of grains that should be substantially enhanced compared to the dose received by particles residing in the midplane. Under those circumstances, the production of S2 should be favored in icy grains over several cycles of vertical transport towards the surface -- 11 -- of the disk. This scenario should also favor the formation of O2 from irradiation of H2O ice (see Mousis et al. 2016a for details). 5. Discussion and conclusions It is reasonable to assume that the multiple forms of irradiation hitting the microscopic icy grains in low density environments such as ISM or the upper layers of protoplanetary disks can lead both to the formation of S2 molecules and the development of cavities in these grains, in which the molecule remains sequestered. The same scenario has been proposed for O2 formation and stabilization in cometary icy grains (Mousis et al. 2016a). In the case of S2 formation, the possibility of forming the dimer via the radiolysis of S-bearing ices remains an open question. Future experimental work is needed to check the viability of this mechanism. The possible formation of S2 in icy grains via their irradiation in ISM, together with the short lifetime of this molecule in the gas phase, leads to the plausible possibility that comets agglomerated from pristine amorphous grains that never vaporized when entering the PSN, as already envisaged for the origin of 67P/C-G's material (Rubin et al. 2015a; Mousis et al. 2016a). On the other hand, the formation of S2 in icy grains that migrated towards the upper layers of the disk is compatible with their condensation in the PSN midplane. This mechanism leaves open the possibility that these grains are made of crystalline ices and clathrates, as proposed by Mousis et al. (2016b) and Luspay-Kuti et al. (2016) to account for several pre-perihelion compositional measurements made by the Rosetta spacecraft in 67P/C-G. The same process could explain the presence of O2 measured in situ in comets 67P/C-G and 1P/Halley (Bieler et al. 2015; Rubin et al. 2015b). Interestingly, whatever the ice structure considered for the icy grains, the voids allowing the stabilization of S2 can be considered as analogs of clathrates in terms of cage sizes and intermolecular interactions. -- 12 -- The fact that one H2S replacing one H2O has little influence on the stability of the solid lattice is a favorable situation for the formation of a mixed ice. It is plausible that some segregation occurs with the formation of H2S islands in the bulk of crystalline or amorphous water ice. Then, the proper conditions would be realized for the in situ formation of S2, especially if we remember that the formation of one S2 requires at least the destruction of two imprisoned sulphur species. The plausible formation of H2S clumps is a strong argument in favor of a non uniform distribution of S2 within cometary ices. Note that in the case of irradiation of crystalline grains condensed in the PSN and transported towards the upper layers of the disk, the formed S2 may be entrapped in clathrates (Grim & Greenberg 1987), also forming a solid phase distinct from water ice in cometary grains. The immediate consequence of the presence of distinct S2-bearing solid phases is the difficulty to predict the S2 correlation with H2O or H2S in 67P/C-G from Rosetta measurements. The S2/H2O abundance ratio is directly linked to the region of the comet whose desorption is observed. Contrary to O2 whose apparent good correlation with H2O is explained by its trapping in water ice (Bieler et al. 2015; Mousis et al. 2016a), no global trend should be drawn between the variation of S2 and H2O abundances if S2 is distributed within both the S-bearing and H2O ices. Indeed, S2 may be released simultaneously from the H2O layer present close to the surface and from H2S clusters localized deeper in the subsurface. Our results are supported by the ROSINA data collected between May 2015 (equinox) and August 2015 (perihelion), showing that there is no clear correlation of S2 with H2O or H2S in 67P/C-G (Calmonte et al. 2016). These observations allow us to exclude the trapping of S2 in a dominant ice reservoir. If S2 was mainly trapped in H2S -- bearing ice, then the outgassing rates of S2 and H2S should have been well correlated during the period sampled by the ROSINA instrument. The same statement applies if S2 had been essentially trapped in water ice. -- 13 -- O.M. acknowledges support from CNES. This work has been partly carried out thanks to the support of the A*MIDEX project (no ANR-11-IDEX-0001-02) funded by the "Investissements d'Avenir" French Government program, managed by the French National Research Agency (ANR). This work also benefited from the support of CNRS-INSU national program for planetology (PNP). J.I.L appreciates support from NASA through the JWST project. A.L.-K acknowledges support from NASA JPL (subcontract no. 1496541). -- 14 -- Fig. 1. -- Illustration of the vertical transport of small icy grains towards disk regions where they are efficiently irradiated. Dust is concentrated in the midplane of the disk due to gravitational settling and gas drag. However, turbulent eddies lift the icy grains toward the upper regions and also drag them down as the direction of the velocity is random and coherent during a timecale comparable to the local keplerian period. Small dust grains finally spend a non negligible fraction of their lifetime in the disk's upper regions, where the irradiation attenuation is low. Higher irradiation regionsUV, X-ray and CRF IRRADIATIONVertical transportDisk midplane -- 15 -- REFERENCES A'Hearn, M. F., Arpigny, C., Feldman, P. D., et al. 2000, Bulletin of the American Astronomical Society, 32, 44.01 A'Hearn M. F., 1992, IAUS, 150, 415 Ahearn M. F., Feldman P. D., 1985, ASIC, 156, 463 Ahearn, M. F., Schleicher, D. G., & Feldman, P. D. 1983, ApJ, 274, L99 Bieler, A., Altwegg, K., Balsiger, H., et al. 2015, Nature, 526, 678 Bockel´ee-Morvan, D., Crovisier, J., Mumma, M. J., & Weaver, H. A. 2004, Comets II, 391 Boice D. C., Reyl´e C., 2005, HiA, 13, 501 Buch, V., Sigurd, B., Devlin, J. P., Buck, U., & Kazimirski, J. K. 2004, International Reviews in Physical Chemistry, 23, 375 Calmonte, U., Altwegg, K., Balsiger, H., et al. 2016, MNRAS, in press Carlson, R. W., Calvin, W. M., Dalton, J. B., et al. 2009, Europa, Edited by Robert T. Pappalardo, William B. McKinnon, Krishan K. Khurana ; with the assistance of Ren´e Dotson with 85 collaborating authors. University of Arizona Press, Tucson, 2009. The University of Arizona space science series ISBN: 9780816528448, p.283, 283 Carlson, R. W., Johnson, R. E., & Anderson, M. S. 1999, Science, 286, 97 Cassidy, T., Coll, P., Raulin, F., et al. 2010, Space Sci. Rev., 153, 299 Ciesla, F. J., & Sandford, S. A. 2012, Science, 336, 452 Ciesla, F. J. 2011, ApJ, 740, 9 -- 16 -- Ciesla, F. J. 2010, ApJ, 723, 514 Cleeves, L. I., Bergin, E. A., Alexander, C. M. O., et al. 2014, Science, 345, 1590 Cleeves, L. I., Adams, F. C., & Bergin, E. A. 2013, ApJ, 772, 5 Dubrulle, B., Morfill, G., & Sterzik, M. 1995, Icarus, 114, 237 Ellinger, Y., Pauzat, F., Mousis, O., et al. 2015, ApJ, 801, L30 Feldman P. D., 1987, IAUS, 120, 417 Grim R. J. A., & Greenberg J. M., 1987, ApJ, 321, L91 Grimme, S., Antony, J., Ehrlich, S., & Krieg, H. 2010, J. Chem. Phys., 132, 154104 Irvine, W. M., Schloerb, F. P., Crovisier, J., Fegley, B., Jr., & Mumma, M. J. 2000, Protostars and Planets IV, 1159 Jim´enez-Escobar, A., Munoz Caro, G. M., & Ciaravella, A., et al. 2012, ApJ, 751, L40 Jim´enez-Escobar, A., & Munoz Caro, G. M. 2011, A&A, 536, A91 Kim S. J., A'Hearn M. F., Wellnitz D. D., Meier R., & Lee Y. S., 2003, Icarus, 166, 157 Kresse, G., & Joubert, D. 1999, Phys. Rev. B, 59, 1758 Kresse, G., & Furthmuller, J. 1996, Phys. Rev. B, 54, 11169 Kresse, G., & Hafner, J. 1994, Phys. Rev. B, 49, 14251 Kresse, G., & Hafner, J. 1993, Phys. Rev. B, 48, 13115 Krishna Swamy K. S., & Wallis M. K., 1987, MNRAS, 228, 305 Laffont, C., Boice, D. C., Moreels, G., et al. 1998, Geophys. Res. Lett., 25, 2749 -- 17 -- Lattelais, M., Pauzat, F., Ellinger, Y., & Ceccarelli, C. 2015, A&A, 578, A62 Lattelais, M., Bertin, M., Mokrane, H., et al. 2011, A&A, 532, A12 Le Roy L., et al., 2015, A&A, 583, A1 Luspay-Kuti, A., Mousis, O., Hassig, M., et al. 2016, Science Advances, 2, e1501781 (2016). Luspay-Kuti, A., Hassig, M., Fuselier, S. A., et al. 2015, A&A, 583, A4 Meyer, B., Stroyer-Hansen, T., & Oommen, T. V. 1972, Journal of Molecular Spectroscopy, 42, 335 Mousis, O., Ronnet, T., Brugger, B., et al. 2016a, ApJ, 823, L41 Mousis, O., Lunine, J. I., Luspay-Kuti, A., et al. 2016b, ApJ, 819, L33 Perdew, J. P., Burke, K., & Ernzerhof, M. 1996, Physical Review Letters, 77, 3865 Reyl´e, C., & Boice, D. C. 2003, ApJ, 587, 464 Rodgers S. D., & Charnley S. B., 2006, AdSpR, 38, 1928 Rubin, M., Altwegg, K., Balsiger, H., et al. 2015a, Science, 348, 232 Rubin, M., Altwegg, K., van Dishoeck, E. F., & Schwehm, G. 2015b, ApJ, 815, L11 Saxena P. P., Singh M., & Bhatnagar S., 2003, BASI, 31, 75 Saxena P. P., & Misra A., 1995, MNRAS, 272, 89 Youdin, A. N., & Lithwick, Y. 2007, Icarus, 192, 588 This manuscript was prepared with the AAS LATEX macros v5.2.
1805.01915
2
1805
2018-07-03T16:11:14
A gap in the planetesimal disc around HD 107146 and asymmetric warm dust emission revealed by ALMA
[ "astro-ph.EP" ]
While detecting low mass exoplanets at tens of au is beyond current instrumentation, debris discs provide a unique opportunity to study the outer regions of planetary systems. Here we report new ALMA observations of the 80-200 Myr old Solar analogue HD 107146 that reveal the radial structure of its exo-Kuiper belt at wavelengths of 1.1 and 0.86 mm. We find that the planetesimal disc is broad, extending from 40 to 140 au, and it is characterised by a circular gap extending from 60 to 100 au in which the continuum emission drops by about 50%. We also report the non-detection of the CO J=3-2 emission line, confirming that there is not enough gas to affect the dust distribution. To date, HD 107146 is the only gas-poor system showing multiple rings in the distribution of millimeter sized particles. These rings suggest a similar distribution of the planetesimals producing small dust grains that could be explained invoking the presence of one or more perturbing planets. Because the disk appears axisymmetric, such planets should be on circular orbits. By comparing N-body simulations with the observed visibilities we find that to explain the radial extent and depth of the gap, it would be required the presence of multiple low mass planets or a single planet that migrated through the disc. Interior to HD 107146's exo-Kuiper belt we find extended emission with a peak at ~20 au and consistent with the inner warm belt that was previously predicted based on 22$\mu$m excess as in many other systems. This warm belt is the first to be imaged, although unexpectedly suggesting that it is asymmetric. This could be due to a large belt eccentricity or due to clumpy structure produced by resonant trapping with an additional inner planet.
astro-ph.EP
astro-ph
MNRAS 000, 1–19 (2015) Preprint 4 July 2018 Compiled using MNRAS LATEX style file v3.0 A gap in the planetesimal disc around HD 107146 and asymmetric warm dust emission revealed by ALMA S. Marino,1(cid:63), J. Carpenter,2 M. C. Wyatt1, M. Booth3, S. Casassus4,5, V. Faramaz6, V. Guzman2, A. M. Hughes7, A. Isella8, G. M. Kennedy9, L. Matr`a10, L. Ricci11 and S. Corder2,12. 1Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 2Joint ALMA Observatory (JAO), Alonso de Cordova 3107 Vitacura - Santiago de Chile, Chile 3Astrophysikalisches Institut und Universitatssternwarte, Friedrich-Schiller-Universitat Jena, Schillergasschen 2-3, 07745 Jena, Germany 4Departamento de Astronomi`I (cid:44)Aa, Universidad de Chile, Casilla 36-D, Santiago, Chile 5Millennium Nucleus a AIJProtoplanetary Disksa A I, Chile 6Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove drive, Pasadena CA 91109, USA. 7Department of Astronomy, Van Vleck Observatory, Wesleyan University, 96 Foss Hill Drive, Middletown, CT 06459, USA 8Department of Physics and Astronomy, Rice University, 6100 Main Street, MS-108, Houston, Texas 77005, USA 9Department of Physics, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK 10Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA 11Department of Physics and Astronomy, California State University Northridge, 18111 Nordhoff Street, Northridge, CA 91130, USA 12National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA 22903-2475, USA Accepted XXX. Received YYY; in original form ZZZ ABSTRACT While detecting low mass exoplanets at tens of au is beyond current instrumenta- tion, debris discs provide a unique opportunity to study the outer regions of planetary systems. Here we report new ALMA observations of the 80-200 Myr old Solar ana- logue HD 107146 that reveal the radial structure of its exo-Kuiper belt at wavelengths of 1.1 and 0.86 mm. We find that the planetesimal disc is broad, extending from 40 to 140 au, and it is characterised by a circular gap extending from 60 to 100 au in which the continuum emission drops by about 50%. We also report the non-detection of the CO J=3-2 emission line, confirming that there is not enough gas to affect the dust distribution. To date, HD 107146 is the only gas-poor system showing multiple rings in the distribution of millimeter sized particles. These rings suggest a similar distribution of the planetesimals producing small dust grains that could be explained invoking the presence of one or more perturbing planets. Because the disk appears axisymmetric, such planets should be on circular orbits. By comparing N-body sim- ulations with the observed visibilities we find that to explain the radial extent and depth of the gap, it would require the presence of multiple low mass planets or a single planet that migrated through the disc. Interior to HD 107146's exo-Kuiper belt we find extended emission with a peak at ∼ 20 au and consistent with the inner warm belt that was previously predicted based on 22µm excess as in many other systems. This warm belt is the first to be imaged, although unexpectedly suggesting that it is asymmetric. This could be due to a large belt eccentricity or due to clumpy structure produced by resonant trapping with an additional inner planet. Key words: circumstellar matter - planetary systems - planets and satellites: dy- namical evolution and stability - techniques: interferometric - methods: numerical - stars: individual: HD 107146. 8 1 0 2 l u J 3 . ] P E h p - o r t s a [ 2 v 5 1 9 1 0 . 5 0 8 1 : v i X r a (cid:63) E-mail: [email protected] © 2015 The Authors 1 INTRODUCTION While exoplanet campaigns have discovered thousands of close in planets in the last decade, at separations greater 2 S. Marino et al. than 10 au it has only been possible to detect a few gas gi- ants, mainly through direct imaging 1 (Marois et al. 2008; Lagrange et al. 2009; Rameau et al. 2013). Protoplane- tary disc observations, on the other hand, have shown that enough mass in both dust and gas to form massive planets resides at large stellocentric distances (see review by An- drews 2015). In addition, the detection of cold dusty debris discs at tens of au shows that planetesimals can and do form at tens and hundreds of au in extrasolar systems (e.g. Su et al. 2006; Hillenbrand et al. 2008; Wyatt 2008; Carpenter et al. 2009; Eiroa et al. 2013; Absil et al. 2013; Matthews et al. 2014a; Thureau et al. 2014; Montesinos et al. 2016; Hughes et al. 2018), although the exact planetesimal belt formation mechanism is a matter of debate (e.g. Matr`a et al. 2018a). It is natural then to wonder how far out can planets form? In situ formation of the imaged distant gas giants is challenging as the growth timescale of their cores can easily take longer than the protoplanetary gas-rich phase (Pollack et al. 1996; Rafikov 2004; Levison et al. 2010). Gravitational instability was thought to be the only potential pathway towards in-situ formation at tens of au (Boss 1997; Boley 2009), but the revisited growth timescale of embryos through pebble accretion could be fast enough to form ice giants or the core of gas giants during the disc lifetime (Johansen & Lacerda 2010; Ormel & Klahr 2010; Lambrechts & Jo- hansen 2012; Morbidelli & Nesvorny 2012; Bitsch et al. 2015; Johansen & Lambrechts 2017). Alternatively, the observed giant planets at tens of au may have formed closer in and evolved to their current orbits by migrating outward (Crida et al. 2009), as could be the case for HR 8799 with four gas giants in mean motion resonances (Marois et al. 2008, 2010) surrounded by an outer debris disc (Su et al. 2009; Matthews et al. 2014b; Booth et al. 2016), or may have been scattered from closer in onto a highly eccentric orbit (Ford & Rasio 2008; Chatterjee et al. 2008; Juri´c & Tremaine 2008), as has been suggested for Fomalhaut b (Kalas et al. 2008, 2013; Faramaz et al. 2015). On the other hand, after the disper- sal of gas and dust, planetesimals could continue growing to form icy planets at tens of au over 100 Myr timescales; how- ever, numerical studies show that once a Pluto size object is formed at 30-150 au within a disc of planetesimals, these are inevitably stirred, stopping growth and the formation of higher mass planets through oligarchic growth (Kenyon & Bromley 2002, 2008, 2010). Thus, it is not yet clear how far from their stars planets can form. Moreover, the discovery of vast amounts of gas (possibly primordial) in systems with low, debris-like levels of dust (so-called "hybrid discs", e.g. Mo´or et al. 2017) has opened the possibility for long lived gaseous discs that could facilitate the formation of both ice and gas giant planets at tens of au. Broad debris discs provide a unique tool to investigate planet formation at tens of au. Planets formed at large radii or evolved onto a wide orbit should leave an imprint in the parent planetesimal belt, and thus in the dust distribution around the system. Gaps have been tentatively identified in a few young debris discs using scattered light observa- tions suggesting the presence of planets at large orbital radii clearing their orbits from debris, e.g. HD 92945 (Golimowski 1 http://exoplanet.eu/ et al. 2011) and HD 131835 (Feldt et al. 2017). However, alternative scenarios without planets that could also repro- duce the observed structure have not been ruled out yet in these systems. For example, multiple ring structures can arise from gas-dust interactions if gas and dust densities are similar (Lyra & Kuchner 2013; Richert et al. 2017), which might explain HD 131835's rings since large amounts of CO gas (likely primordial origin) have been found in this sys- tem (Mo´or et al. 2017). Moreover, the double ring structure around HD 92945 and HD 131835 has only been identified in scattered light images, tracing small dust grains whose dis- tribution can be highly affected by radiation forces (Burns et al. 1979), therefore not necessarily tracing the distribution of planetesimals (e.g. Wyatt 2006). Only HD 107146, an ∼ 80 − 200 Myr old G2V star (Williams et al. 2004, and references therein) at a dis- tance of 27.5 ± 0.3 pc (Gaia Collaboration et al. 2016a,b), has a double debris ring structure tentatively identified at longer wavelengths thanks to the Atacama Large Mil- limeter/submillimeter Array (ALMA, Ricci et al. 2015a). At these wavelengths observations trace mm-sized dust for which radiation forces are negligible, therefore indicating that the double ring structure is imprinted in the planetesi- mal distribution as well. Moreover, these observations ruled out the presence of gas at densities high enough to be re- sponsible for such structure. The debris disc surrounding HD 107146 was first discovered by its infrared (IR) excess using IRAS data (Silverstone 2000), but it was not until re- cently that the disc was resolved by the Hubble Space Tele- scope (HST) in scattered light, revealing a nearly face on disc with a surface brightness peak at 120 au and extending out to ∼ 160 au (Ardila et al. 2004; Ertel et al. 2011; Schneider et al. 2014). Despite HST's high resolution, limitations in subtracting the stellar emission or a smooth distribution of small dust likely kept the double ring structure hidden. Us- ing ALMA's unprecedented sensitivity and resolution, Ricci et al. (2015a) showed that this broad disc extended from about 30 to 150 au, but that it had a decrease in the dust density at intermediate radii, which could correspond to a gap produced by a planet of a few Earth masses clearing its orbit at 80 au through scattering. Finally, analysis of Spitzer spectroscopic and photometric data revealed the presence of an extra unresolved warm dust component in the system, at a temperature of ∼120 K and thus inferred to be located between 5-15 au from the star (Morales et al. 2011; Kennedy & Wyatt 2014). Despite the tentative evidence of planets producing these gaps around their orbits, neither the HD 107146 ALMA observations, nor the scattered light observations of HD 92945 and HD 131835, ruled out alternative scenarios in which planets are not orbiting within these gaps, but similar structure is created in the dust distribution through differ- ent mechanisms. These different scenarios have important implications for the inferred dynamical history of the system and planet formation. While a planet formed in situ could explain the data reasonably well, questions arise regarding how a planet of a few Earth masses could have formed at such large separations, where coagulation and planetesimal growth timescales are significantly longer. Alternative sce- narios such as the one suggested by Pearce & Wyatt (2015) to explain HD 107146's gap, could avoid these issues. In that scenario a broad gap is produced by secular interactions be- MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 3 Table 1. Summary of band 6 and band 7 (12m and ACA) observations. The image rms and beam size reported corresponds to briggs weighting using robust=0.5. Observation Dates Band 6 - 12m Band 7 - 12m Band 7 - ACA Band 7 - 12m+ACA 24, 27-30 Apr 2017 11 Dec 2016 20 Oct and 2 Nov 2016 22 Mar and 13 Apr 2017 - tsci [min] 236.4 48.8 135.1 - Image rms beam size (PA) [µJy] 6.8 30.0 245 31.1 0.(cid:48)(cid:48)67 × 0.(cid:48)(cid:48)66 (−2.9◦) 0.(cid:48)(cid:48)46 × 0.(cid:48)(cid:48)37 (21.2◦) 4.(cid:48)(cid:48)3 × 3.(cid:48)(cid:48)4 (−76.4◦) 0.(cid:48)(cid:48)47 × 0.(cid:48)(cid:48)38 (20.8◦) Min and max baselines [m] (5th and 95th percentiles) Flux calibrator Bandpass calibrator Phase calibrator 41 and 312 48 and 410 9 and 44 J1229+0203 J1229+0203 Titan J1229+0203 J1229+0203 J1256+0547 J1215+1654 J1215+1654 J1224+2122 - - - - tween a planetesimal disc and a similar mass planet on an eccentric orbit, which formed closer in and was scattered out by an additional massive planet. That scenario also predicts that the planet's orbit should become nearly circular and the planet would be located at the inner edge of the disc at the current epoch. The model also predicts the presence of asymmetries in the disc such as spiral features that would be detectable in deeper ALMA observations. A second alternative scenario was proposed by Golimowski et al. (2011) to explain the double-ring struc- ture around HD 92945 seen in scattered light. As shown by Wyatt (2003), planet migration can trap planetesimals in mean motion resonances; resulting in overdensities that are stationary in the reference frame co-rotating with the planet. Small dust released from these trapped planetesimals can exit the resonances due to radiation pressure, forming a double-ring structure that could be observable in scattered light (Wyatt 2006). On the other hand, in this planet migra- tion scenario the distribution of mm-sized dust should match the planetesimal distribution, with prominent clumps that could be seen in millimetre observations. Finally, secular resonances produced by a single eccen- tric planet in a massive gaseous disc (Zheng et al. 2017) or by two planets formed interior to the disc could also explain some of the wide gaps (Yelverton et al. in prep), but possibly also leaving asymmetric features. Hence debris disc observa- tions at multiple wavelengths can disentangle these different scenarios, and provide insights into the dynamical history of the outer regions of planetary systems, testing the exis- tence and origin of planets at tens of au, which otherwise would remain invisible. Since radiation forces acting on the smallest dust grains have a significant effect on their dis- tribution, the planetary perturbations discussed above can be best studied using ALMA observations which trace the distribution of large (∼0.1-10mm) grains for which radiation forces are negligible, and thus follow the distribution of their parent planetesimals. In this paper we present new ALMA observations of HD 107146 in both band 6 and 7 (1.1 and 0.86 mm). These observations resolve the broad debris disc around this system at higher sensitivity and resolution than the data presented by Ricci et al. (2015a), which showed tentative evidence of a gap as commented above. This paper is outlined as fol- lows. In §2 we present the new ALMA observations of the dust continuum and line emission of HD 107146. Then in §3 we model the data using both parametric models and the output of N-body simulations to quantify the disc struc- ture and assess whether a single planet could explain the observations. In §4 we discuss our results, the origin of the gap and implications for planet formation; the total mass of HD 107146's outer disc; the detection of an inner component MNRAS 000, 1–19 (2015) that could be warm dust; and our gas non-detection. Finally the main conclusions of this paper are summarised in §5. 2 OBSERVATIONS HD 107146 was observed both in band 7 (0.86mm, project 2016.1.00104.S, PI: S. Marino) and band 6 (1.1mm, project 2016.1.00195.S, PI: J. Carpenter). Band 7 observations were carried out between October and December 2016 (see Ta- ble 1) both using the 12m array and the Atacama Compact Array (ACA) to recover small and large scale structures. The total number of antennas for the 12m array was 42, with baselines ranging from 48 to 410 m (5th and 95th per- centiles), and between 9 to 11 ACA 7m antennas with base- lines ranging from 9 to 44 m. The correlator was set up with two spectral windows centered at 343.13 and 357.04 GHz with 2 GHz bandwidths and 15.625 MHz spectral resolu- tion, and the other two centered at 345.03 and 355.14 GHz with 1.875 GHz bandwidths and 0.976 MHz spectral res- olution. The four windows are used together to study the dust continuum emission, while the latter two are also used specifically to search for line emission from CO and HCN molecules in the disc (see §2.2). The weather varied between the multiple ACA band 7 observations, with average PWV values of 0.38, 0.72, 0.89 and 1.2 mm. The average PWV during the single 12m observation was 0.56 mm. Band 6 observations were carried out in April 2017 (see Table 1) using the 12m array only. We requested observa- tions using two different 12m antenna configurations to re- cover well the large scale structure and, at the same time, achieve high spatial resolution, but due to time constraints the compact configuration observations were never carried out. We find however that the observations with the ex- tended configuration have baselines short enough to recover the large scale structure (see §2.1 below). The total number of 12m antennas varied between 41 and 42, with baselines ranging from 41 to 312 m (5th and 95th percentiles). The correlator was set up with four spectral windows centered at 253.60, 255.60, 269.61 and 271.61 GHz with 2 GHz band- widths and 15.625 MHz spectral resolution. The four are used together to study the dust continuum emission only. The weather also varied between the multiple band 6 ob- servations, with PWV values of 0.89, 0.33, 0.30, 0.82 and 1.56 mm. Despite these variations we decided to use all the data sets to obtain the highest possible S/N. Calibra- tions were applied using the pipeline provided by ALMA and CASA 4.7. The total time on source for band 6 was 236 min, and 184 min for band 7 (49 and 135 min for the 12m array and ACA, respectively). Below we present the image analysis of continuum and line observations. 4 S. Marino et al. 2.1 Continuum Continuum maps at band 6 and 7 are created using the clean task in CASA 4.7 (McMullin et al. 2007) and are presented in Figure 1. We adopt natural weighting for band 6 and 7 (12m+ACA combined) for a higher signal to noise. These maps have a rms of 6.3 and 27 µJy beam−1 at the center, which increases towards the edges as the maps are corrected by the primary beam, reaching values of 7.0 and 34 µJy beam−1 at 5(cid:48)(cid:48) (140 au) for band 6 and 7, respectively. The band 6 and band 7 synthesised beams have dimensions of 0.(cid:48)(cid:48)80 × 0.(cid:48)(cid:48)79 and 0.(cid:48)(cid:48)58 × 0.(cid:48)(cid:48)472, respectively, reaching an approximate resolution of 22 au for band 6 and 15 au for band 7. The higher resolution and sensitivity of these new data reveal a nearly axisymmetric broad disc with a large decrease or gap in the surface brightness centered at a radius of ∼ 80 au as suggested by Ricci et al. (2015a). The disc ex- tends from nearly 40 to 180 au, being one of the widest discs resolved at millimetre wavelengths (Matr`a et al. 2018a). We compute the total flux by integrating the disc emission in- side an elliptical mask with a 180 au semi-major axis and the same aspect ratio as the disc (see §3.1). We find a total flux of 16.1 ± 1.6 and 34.4 ± 3.5 mJy at 1.1 and 0.86 mm (includ- ing 10% absolute flux uncertainties), leading to a spatially unresolved millimetre spectral index of 2.64 ± 0.48, consis- tent with results from Ricci et al. (2015b) which combined ALMA and ATCA observations. The total flux in band 6 is consistent with that measured by Ricci et al. (2015a) at similar wavelengths and using a more compact ALMA con- figuration, thus proving that our band 6 observations do not suffer from flux loss or miss large scale structure. In order to study in more detail the disc radial struc- ture, the top panel of Figure 2 shows the radial intensity profile, computed by azimuthally averaging the disc emis- sion over ellipses as in Marino et al. (2016, 2017a,b). Both in band 6 and 7, we find that the disc surface brightness peaks near the disc inner edge at ∼ 45 au, from which it decreases reaching a minimum at 80 au that is deeper in the band 7 profile likely due to the higher resolution. Be- yond this minimum, the surface brightness increases until 120 au where it peaks and then decreases steeply with ra- dius. No significant positive emission is recovered beyond 180 au. Within the gap, we find that the radial profile is not symmetric with respect to the minimum, with the outer sec- tion (80-100 au) having a steeper slope than the inner part (60-80 au), a feature that is present both in band 6 and 7 data. This is also visible in Figure 1 and is an important fea- ture that could shed light on the origin of this gap. We also compute a spectral index map using multi-frequency clean (nterms=2) and natural weights. The bottom panel of Fig- ure 2 shows the azimuthally averaged radial profile of the spectral index (αmm). From 40 to 150 au the disc has spec- tral index of roughly 2.7±0.1 (assuming αmm is constant over radii), consistent with the overall spectral index estimated above. Note that uncertainties on αmm do not include the 10% absolute flux uncertainties as we are only interested in relative differences as a function of radius. In addition to the outer disc, we detect emission from 2 Note that the rms and beam sizes are different to the one re- ported in Table 1 as the imaging is done with natural weights to increase the S/N within 30 au that peaks near the stellar position in the az- imuthally averaged profile. However, it has a much higher level than the photospheric emission if we extrapolate this from available photometry at wavelengths shorter than 10µm using a Rayleigh-Jeans spectral index (Fν(cid:63) = 32 ± 1 and 18± 1 µJy at 0.86 and 1.1 mm, respectively). Moreover, between 20 and 30 au we also find that αmm has a peak of ∼ 3.7, although still within 2σ from the average spectral in- dex. In §3.1 we recover this inner component in more detail after subtracting the disc emission using a parametric model and we find that it is inconsistent with point source emis- sion, it is significantly offset from the stellar position, and unlikely to be a background sub-millimetre galaxy. The fact that the disc emission is consistent with being axisymmetric disfavours the scenario in which a planet is scattered out from the inner regions opening a gap through secular interactions with the disc (Pearce & Wyatt 2015, see their figure 6). In their model, spiral density features are present in the planetesimal disc for hundreds of Myr, which should be imprinted in the mm-sized dust distribution as well, thus being detectable by our observations. In §3.1 we fit parametric models to the data to study with more detail the level of axisymmetry of HD 107146's disc. Moreover, in §3.2 we compare our observations with N-body simulations of a planet on a circular orbit clearing a gap in a planetesimal disc. 2.2 CO J=3-2 and HCN J=4-3 Despite the increasing number of CO gas detections in nearby debris discs (e.g. Dent et al. 2014; Mo´or et al. 2015; Marino et al. 2016, 2017a; Lieman-Sifry et al. 2016; Matr`a et al. 2017b; Mo´or et al. 2017), no CO v=0 J=3-2 emis- sion is detected around HD 107146 in the dirty continuum- subtracted data cube. This non detection is, however, not surprising since HD 107146 is significantly older and fainter than the other systems in which primordial gas has been found (i.e. the hybrid discs). Furthermore, the sensitivity of these observations is expected to be insufficient to detect CO gas if being released through collisions of volatile-rich plan- etesimals (Kral et al. 2017). We search more carefully for CO emission by applying a matched filter technique (Matr`a et al. 2015; Marino et al. 2016, 2017a; Matr`a et al. 2017b) in which we integrate the emission over an elliptic mask with the same orientation as the disc on the sky (see §3), but in each spatial pixel we integrate only over those frequencies (i.e. radial velocities) where line emission is expected taking into account the Doppler shift due to Keplerian rotation. For this, we assume a stellar mass of 1 M(cid:12), inclination of 19◦, PA of 153◦, and deprojected minimum and maximum radii of 40 and 150 au, respectively. Using this method we obtain an integrated line flux 3σ upper limit of 74 mJy km s−1. Note that we consider the two possible directions of the rotation, obtaining similar limits. We also search for emission that could be present at a specific radius by integrating the emis- sion azimuthally, however no significant emission is found. In §4.5, we use this total flux upper limit to estimate an up- per limit on the CO gas mass that could be present in the disc. Similarly, we search for HCN emission, finding no sig- nificant emission. Based on this non detection we place a 3σ upper limit of 91 mJy km s−1, which we also use in §4.5 MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 5 Figure 1. Clean images of HD 107146 at 1.1 (left) and 0.86 mm (right) using natural weights and primary beam corrected. The band 7 image is obtained after combining the 12m array and ACA data. The stellar position is marked with a white cross at the center of the image, while the beams of band 6 (0.(cid:48)(cid:48)80× 0.(cid:48)(cid:48)79) and band 7 (0.(cid:48)(cid:48)58× 0.(cid:48)(cid:48)47) are represented by black ellipses in the bottom left corners. The image rms at the center is 6.3 and 27 µJy beam−1 increasing with distance from the center and reaching values of 7.0 and 34 µJy beam−1 at 5(cid:48)(cid:48) from the center. to constrain its abundance in planetesimals in this system. HCN is of particular interest for exocometary studies as, besides being abundant in Solar System comets (Mumma & Charnley 2011), it has recently been suggested that it could play a key role for prebiotic chemistry in habitable planets (Patel et al. 2015; Sutherland 2017). 3 MODELLING In this section we model the data using parametric models to constrain the density distribution of solids in the system (§3.1), and N-body simulations of a planet embedded in a planetesimal disc tailored to HD 107146 to constrain the mass and orbit of a putative planet carving the observed gap (§3.2). In both approaches we model the central star as a G2V type star with a mass of 1 M(cid:12), an effective temperature of 5750 K and a radius of 1 R(cid:12). 3.1 Parametric model We first use a set of parametric models to study the un- derlying density distribution of mm-sized dust in the sys- tem, which we fit directly to the observed visibilities as in Marino et al. (2016, 2017a,b). Inspired by the radial profile of the dust emission (Figure 2), we first choose as a disc model an axisymmetric disc with a surface density that is parametrized as a triple power law that divides the disc into Figure 2. Average intensity (top) and spectral index (bottom) profile computed azimuthally averaging the Clean and spectral in- dex maps over ellipses oriented as the disc in the sky. The blue and orange lines are obtained from the band 6 and band 7 Clean im- ages, using Briggs (robust=0.5) weights, while the spectral index is computed using natural weights. The shaded areas represent the 68% confidence region over a resolution element (represented by circles spaced by 18 au for band 6, 13 au for band 7 and 22 au for the spectral index). MNRAS 000, 1–19 (2015) 050100Intensity[µJybeam−1]B6(1.1mm)B7(0.86mm)050100150200Radius[au]234αmm 6 S. Marino et al. an inner edge, an intermediate section (where the bulk of the dust mass is) and an outer edge. On top of this, the triple power law surface density distribution has a gap, which we parametrise with a Gaussian profile, to reproduce the de- pression seen in the ALMA data. This parametrization in- troduces a total of 9 parameters that define the surface den- sity as follows, Σ(r) = Σ0 fgap(r) fgap(r) = 1 − δg exp (cid:17)γ2 r < rmin, rmin < r < rmax, (1) r > rmax, (2) ,  rmin rmin (cid:16) r (cid:17)γ1 (cid:16) r (cid:17)γ2 (cid:17)γ3(cid:16) rmax (cid:16) r (cid:34) (cid:35) −(r − rg)2 2σ2 g rmax rmin 2(cid:112)2 ln(2)σg. We leave as a free parameter the total dust mass where rmin and rmax are the inner and outer radii of the disc, γ1,2,3 determine how the surface density varies interior to the disc inner radius, within the disc and beyond the disc outer radius. The gap is parametrized with a fractional depth δg, a center rg and a full width half maximum (FWHM) wg = Md, which is the surface integral of Σ(r). Although the disc is close to face on, we still model the dust distribution in three dimensions adding the scale height h as an extra parameter (i.e. vertical standard deviation of hr), and imposing a prior of h > 0.03. We solve for the dust equilibrium temperature and com- pute images at 0.86 and 1.13 mm using RADMC-3D3. We assume a weighted mean dust opacity corresponding to dust grains made of a mix of astrosilicates (Draine 2003), amor- phous carbon and water ice (Li & Greenberg 1998), with mass fractions of 70%, 15% and 15%, respectively, and as- suming a size distribution with an exponent of -3.5 and min- imum and maximum sizes of 1µm and 1cm. This translates to a dust opacity of 1.5 cm2 g−1 at 1.1 mm. We note that these choices in dust composition and size distribution have no significant effect on our modeling apart from the derived total dust mass. We then use these images to compute model visibilities at the same uv points as the 12m band 6, 12m band 7 and ACA band 7 observations by taking the Fast Fourier Transform after multiplying the images by the cor- responding primary beam. Additionally, we leave as free pa- rameters the disc inclination (i), position angle (PA), RA and Dec offsets for the three observation sets, and a disc spectral index (αmm) that sets the flux at 0.86 mm given the dust mass and opacity at 1.1 mm, i.e. the size distribution is assumed to be the same throughout the disc. In total, our model has 19 parameters, 10 for the density distribution and 9 for the disc centre, orientation, and spectral index. To find the best fit parameters we sample the param- eter space using the PYTHON module emcee, which im- plements Goodman & Weare's Affine Invariant MCMC En- semble sampler (Goodman & Weare 2010; Foreman-Mackey et al. 2013). The posterior probability distribution is de- fined as the product of the likelihood function (proportional to exp[− χ2/2]) and prior distributions which we assume uni- form, although we impose a lower limit for h of 0.03 due to model resolution constraints. In computing the χ2 over the three visibility sets we applied three constant re-weighting 3 http://www.ita.uni-heidelberg.de/ dullemond/software/radmc- 3d/ Table 2. Best fit parameters of the ALMA data for the differ- ent parametric models. The quoted values correspond to the me- dian, with uncertainties based on the 16th and 84th percentiles of the marginalised distributions or upper limits based on 95th percentile. Parameter best fit value description 3-power law + Gaussian gap Md [M⊕] rmin [au] rmax [au] γ1 γ2 γ3 rg [au] wg [au] δg h PA [◦] i [◦] αmm 0.250 ± 0.004 46.6+1.4−1.5 135.6+1.1−1.2 2.6+0.3−0.2 0.26+0.08 −0.10 −10.5+0.9−1.0 75.5+1.1−1.2 38.6+4.5−3.6 0.52+0.03 −0.02 0.12+0.04 −0.05 153 ± 3 19.3 ± 1.0 2.57 ± 0.11 millimetre spectral index total dust mass disc inner radius disc outer radius inner edge's slope disc slope outer edge's slope radius of the gap FWHM of the gap fractional depth of the gap scale height disc PA disc inclination from face-on Step gap rg [au] wg [au] δg Eccentric disc 75.4+0.8−0.7 42.2+1.7−2.2 0.43 ± 0.02 radius of the gap width of the gap depth of the gap ed <0.03 disc global eccentricity Inner component rmin [au] γ1 γ2 rg [au] wg [au] δg Mc [M⊕] rc [au] ∆rc [au] ωc [◦] σφ [◦] 41.9+1.2−1.4 11.6+3.0−2.7 0.03+0.19 −0.26 72.1+2.2−2.9 51+12−8 0.58+0.08 −0.06 3.0+0.9−0.6 × 10−3 19.3+2.8−2.8 35.8+9.1−6.7 85+9−9 94+15−12 disc inner radius inner edge's slope disc slope radius of the gap FWHM of the gap fractional depth of the gap dust mass inner component radius of inner component radial width of inner component PA of inner component south of disc PA, and in the disc plane azimuthal width of inner component factors for band 6, 12m band 7 and for ACA band 7 that en- sures that the final reduced χ2 of each of the three sets is ap- proximately 1 without affecting the relative weights within each of these data sets that are provided by ALMA. The re-scaling is necessary as the absolute uncertainty of ALMA visibilities can be offset by a factor of a few, even after re- weighting the visibilities with the task statwt in CASA 4.7. These factors could be alternatively left as free param- eters by adding an extra term to the likelihood function, however we find no differences in our results compared to leaving them fixed during multiple tries. Therefore we opt for leaving them fixed. In Table 2 we present the best fit parameters of our 3- power law model with a Gaussian gap. We find a disc inner radius of 47 au that is significantly larger than the inner edge of ∼ 25 − 30 au derived by Ricci et al. (2015a). This is because in their model they considered a sharp inner edge, MNRAS 000, 1–19 (2015) while in ours we allow for the presence of dust within this inner radius, but decreasing towards smaller radii. We find that rmin matches well the radius at which the surface den- sity peaks as seen in Figure 2. Similarly, our estimate of the outer radius (136 au) is significantly smaller than their outer edge (150 au). We find that the disc inner edge has a power law index (slope hereafter) of ∼ 2.6, while the outer edge is much steeper with a slope of −11. The intermediate component has a very flat slope of ∼ 0.3, 2.7σ flatter than the previous estimate (0.59), although the difference could be simply due to different parametrizations, i.e. considering a three vs a single power law parametrization or leaving the gap's depth as a free parameter vs fixed to 1. Note that the slope derived from the intermediate component does not im- ply that the mass surface density of planetesimals is flat. In fact, from collisional evolution models we expect the surface density of mm-sized dust and optical depth to have a lower slope than the total mass surface density in regions of the disc where the largest planetesimals are not yet in collisional equilibrium (i.e. have a lifetime longer than the age of the system, Schuppler et al. 2016; Marino et al. 2017b; Geiler & Krivov 2017). Moreover, we also expect that in the in- ner regions where the largest planetesimals are in collisional equilibrium, the surface density of material should have a slope close to 7/3 (Wyatt et al. 2007; Kennedy & Wyatt 2010), as we find interior to rmin. This suggests that the re- gions interior to 47 au might be relatively depleted of solids simply due to collisional evolution rather than clearing by planets or inefficient planetesimal formation. We compare the derived inner radius and surface density of millimetre grains (∼ 3 × 10−6 M⊕ au−2 at 50 au) with collisional evo- lution models by Marino et al. (2017b), which include how the size distribution evolves at different radii, to estimate the maximum planetesimal size and initial total surface density of solids. We find that the best match has a maximum plan- etesimal size of ∼ 10 km and an initial disc surface density of 0.015 M⊕ au−2 at 50 au, i.e. 5 times the surface density of the Minimum Mass Solar Nebula (Weidenschilling 1977; Hayashi 1981) extrapolated to large radii, or a total solid disc mass of 300 M⊕. This implies a very massive initial disc and efficient planetesimal formation at large radii. These conclusions as- sume, however, that the observed structure in the surface brightness profile arise from collisional evolution neglecting alternative origins. Finally, regarding the disc orientation, we find values that are consistent with previous estimates, but with tighter constraints (see Table 2). Our results show that the gap is centered at 75.5+1.1−1.2 au, consistent with the previous estimate. However, we find a FWHM of ∼40 au that is much larger than the previous estimate of 9 au, likely due to Ricci et al. (2015a) assuming a gap depth of 100%. Instead, we fit the depth of the gap finding a best fit value of 0.5. We also fit an alternative model in which the gap is a step function with a constant depth, finding best fit values that are similar to the model with the Gaussian gap (see Table 2). The difference in widths between the previous study and this work is interesting as Ricci et al. (2015a) derived the mass of a putative planet clearing a gap in the disc based on the gap's width and assuming that it should be roughly equal to the planet's chaotic zone. A 3-4 times wider gap would imply a much higher planet mass as the width of the chaotic zone scales as (Wisdom 1980; Duncan et al. 1989). Such planet mass M2/7 p MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 7 Figure 3. Deprojected and binned visibilities assuming a disc po- sition angle of 153◦ and inclination of 19◦. The real components of the band 6 and band 7 data are presented in the top and mid- dle panels, respectively. The imaginary components of the 12m data are displayed in the lower panels. The errorbars represent the binned data with their uncertainty estimated as the standard deviation in each bin divided by the square root of the number of independent points. The continuous line shows the triple power law best fit model with a Gaussian gap. The dashed line repre- sents the best fit model with the same parametrization, but with an additional inner component. Note that the scale in the left and right panels is different. estimates assume that the system is in steady state, that the gap is devoid of material, and that the gap's width is simply equal to the chaotic zone's width. However, given the age of the system (∼80-200 Myr), and that any planet may be younger, it is reasonable to consider that the distribution of particles in the planet's vicinity could still be evolving. Therefore, instead of using the gap's width to estimate a planet mass, in §3.2 we estimate this by comparing with N-body simulations tailored to HD 107146. Using the best fit PA and i we deproject the observed visibilities and bin them to compare them with our axisym- metric model in Figure 3 (continuous line). The 12m and ACA band 7 observations are consistent with each other within errors, with some systematic differences due to the different primary beams. We find that our best fit model fits well both the real and imaginary components, with the −20246810Real,1.1mm[mJy]12mband6−0.50.00.5−50510152025Real,0.86mm[mJy]12mband7ACAband7−0.50.00.5050100150−0.50.00.5Imag[mJy]200250300350−0.50.00.5DeprojectedBaseline[kλ] 8 S. Marino et al. Figure 4. Best fit model images (1.1 mm) and dirty map of residuals. Left column: 3-power law model with a Gaussian gap. Right column: 3-power law model with a Gaussian gap and a warm inner component. The residuals of band 6 (middle row) and band 7 (12m+ACA, bottom row) are computed using natural weights and with an outer taper of 0.(cid:48)(cid:48)8, leading to a synthesised beam of 0.(cid:48)(cid:48)80 × 0.(cid:48)(cid:48)79 and 1.(cid:48)(cid:48)0 × 0.(cid:48)(cid:48)9 and an rms of 6.1 and 41 µJy beam−1, respectively. Contours represent 3, 5 and 8σ. imaginary part being consistent with zero as expected for an axisymmetric disc. However, we find a significant deviation between the real components of the 12m band 7 data and model at around 50 kλ (corresponding to an angular scale of 4(cid:48)(cid:48)), with the model visibility predicting a slightly lower value than the data. This could be due to large scale varia- tions in the spectral index as the same model fits well those baselines in the band 6 data. This deviation is the same when comparing the model with a step gap, which has an indistinguishable visibility profile for baselines shorter than 200 kλ. Beyond 150 kλ we also find deviations between the data and model, expected since as shown in Figure 2 the radial profile seems to have structure that is more complex than a simple Gaussian gap. A more intuitive way to study the goodness of fit of our model is to look at the dirty map of the residuals. The left column of Figure 4 shows the dirty maps of the residuals af- ter subtracting our best fit model from the data (in visibility space). These are computed using natural weights, plus an outer taper of 0.(cid:48)(cid:48)8 for band 7 to obtain a synthesised beam of similar size compared to the band 6 beam. Both band 6 and 7 residual images show significant residuals (with peaks of 9 and 5 times the rms) within the disc inner edge and with a peak that is significantly offset from the stellar po- sition by ∼ 0.(cid:48)(cid:48)5, corresponding to a projected distance of ∼ 15 au. Moreover, the emission is marginally resolved, ex- tending radially by more than a beam (i.e. larger than 10 au) and with an integrated flux significantly higher than the pre- dicted photospheric emission at these wavelengths assuming a Rayleigh-Jeans spectral index (18 and 32 µJy at 1.1 and 0.86 mm). We estimate the flux of these components by inte- grating the dirty maps over a circular region of 2(cid:48)(cid:48) diameter, finding an integrated flux of 85 ± 13 µJy and 290 ± 70 µJy at 1.1 and 0.86 mm, respectively (without including 10% absolute flux uncertainties). Because this inner emission is resolved, it cannot arise from a compact source such as a planet or circumplanetary material. We identify two plausible origins for this inner com- ponent. It could be the same warm dust that was inferred based on Spitzer data (Morales et al. 2011), as the clump is at a radial distance that is consistent with the inferred black body radius (5-15 au) from its spectral energy dis- tribution (SED), although our detection would imply that the warm component has an asymmetric distribution (see discussion in §4.4). On the other hand, the residual could be also due to a background sub-millimeter galaxy that are often detected incidentally as part of ALMA deep observa- tions (see Simpson et al. 2015; Carniani et al. 2015; Marino et al. 2017b; Su et al. 2017). Sub-millimeter galaxies have typical sizes of the order of 1(cid:48)(cid:48) and spectral indices ranging between ∼ 3−5 at mm-wavelengths, thus consistent with the observed clump. We estimate a probability of ∼ 50% of find- ing a sub-millimeter galaxy as bright as 0.3 mJy at 0.86 mm or 0.1 mJy at 1.1 mm, respectively, within the entire band 6 and 7 primary beams (Simpson et al. 2015; Carniani et al. 2015). Hence, the detection of such a bright background ob- ject is not a rare event. However, we find that the probability of finding such a bright sub-millimeter galaxy at 0.86 mm and 1.1 mm and co-located with warm dust (i.e within 1(cid:48)(cid:48) from the star) is only 0.6% and 1%, respectively; therefore favouring the warm dust scenario. In fact, we do not detect any other compact emission above 5σ within the band 7 and 6 primary beams, consistent with the number counts of sub-millimeter galaxies. We also discard that this emission could originate from the galaxy detected using HST in 2004, 2005 and 2011 (Ardila et al. 2004; Ertel et al. 2011; Schnei- der et al. 2014), as its position would only have changed by 1.4(cid:48)(cid:48), and therefore still lie at ∼ 5(cid:48)(cid:48) (∼ 140 au) from the star in 2017, and is not detected in our observations. 3.1.1 Disc global eccentricity As shown by Pearce & Wyatt (2014, 2015), a planet on an eccentric orbit can force an eccentricity in a disc of plan- etesimals through secular interactions. Here we aim to as- sess if HD 107146's debris disc could have a global eccen- tricity and pericentre, using the same parametrization as in Marino et al. (2017a), i.e. taking into account the expected apocentre glow (Pan et al. 2016). We find that the disc is consistent with being axisymmetric, with a 2σ upper limit of 0.03 for the forced eccentricity. We find though that the marginalised distribution of ed peaks at 0.02 with a pericen- tre that is opposite to the residual inner clump. This peak is likely produced as the residuals are lower when the disc is ec- MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 9 as bright as this inner emission at 0.86 and 1.1 mm, respec- tively, and within 1(cid:48)(cid:48) from the star (Simpson et al. 2015; Carniani et al. 2015). These results also confirm that the peak of this inner component is significantly offset from the star and it is incompatible with an axisymmetric inner com- ponent. If this emission is produced by warm dust, then it could bring valuable insights into the origin of warm dust emission in general, as it is inconsistent with an axisymmet- ric asteroid belt (see §4.4). Because the new estimate of the inner edge slope is too steep to be consistent with being set by collisional evolu- tion, the maximum planetesimal size is only constrained to be (cid:38) 10 km. Despite this, the relative brightness between the inner and outer components can still be explained sim- ply by collisional evolution. This has also been found for other systems with warm dust components, e.g. q1 Eridani (Schuppler et al. 2016), suggesting the presence of planets clearing the material in between (Shannon et al. 2016). centric and with an apocentre oriented towards the clump's position angle due to apocentre glow. We therefore conclude that the fit is biased by the inner clump and that the disc is probably not truly eccentric. 3.1.2 Inner component In order to constrain the geometry or distribution of the inner emission found in the residuals, we add an extra in- ner component by introducing five additional parameters to our reference parametric model (3-power law surface density with a Gaussian gap). We parametrise its surface density as a 2D Gaussian in polar coordinates, with a total dust mass Mc and centered at a radius rc and azimuthal angle ωc (mea- sured in the plane of the disc from the disc PA and increasing in an anti-clockwise direction). The width of this Gaussian is parametrized with a radial FWHM ∆rc and an azimuthal standard deviation σφ. Best fit values are presented in Ta- ble 2. We find that this inner component is extended both radially and azimuthally, but concentrated around 19± 3 au and orthogonal to the disc PA (ωc ∼ 90◦), as we found in the residuals of our axisymmetric model. In the right panel of Figure 4 we present the model image of the best fit model and its residuals, which are below 3σ in both band 6 and 7 within the disc inner edge. We also find that the total dust mass of this inner component is 3.0+0.9−0.6 × 10−3 M⊕ (∼1% of the outer disc mass). When adding this extra inner component we find that the slope of the inner edge is steeper and the inner radius smaller compared with our previous model (symmetric disc model hereafter), as it was probably compensating for the emission within 30 au with a less steep inner edge. We also find a slightly smaller gap radius of 72 au, a larger and deeper gap (51 au wide and 0.58 deep), and a flatter surface density slope of 0.0±0.2. These differences in the gap's structure and disc slope are overall consistent within 3σ with our previous estimates, but significantly improve the fit at large baselines as Figure 3 shows (dashed blue and orange lines). These im- provements in the fit at large baselines are not due to the addition of the clump, but due to the different best fit sur- face density profile of the outer disc. In fact, the visibilities of the inner component are negligible beyond 200 kλ. The rest of the parameters are consistent within 1σ with the values presented for the symmetric disc model. To have a better estimate of the flux of this inner component, we subtract the new best fit of the outer component (its inner edge is steeper), finding an integrated flux over a circular region of 3(cid:48)(cid:48) diameter of 0.82 ± 0.14 and 0.31 ± 0.04 mJy at 0.86 and 1.1 mm, respectively (including 10% absolute flux uncertain- ties)4. Note that this flux is a factor two higher than that estimated from the residuals of the symmetric disc model, as without the inner component the model tries to compensate for the emission interior to 40 au. From these fluxes we esti- mate a spectral index of 3.3 ± 0.6, thus still consistent with the typical observed spectral indices of debris discs. Based on this new flux estimate at 0.86 mm, we find an even lower probability of 0.1% and 0.3% of finding a sub-mm galaxy 4 We also measured the integrated flux of the inner component using the task uvmodelfit in CASA 4.7, obtaining values of 0.35± 0.03 and 0.81 ± 0.14 mJy at 1.1 and 0.86 mm, respectively. MNRAS 000, 1–19 (2015) 10 S. Marino et al. 3.2 N-body simulations In this section we compare the observations with dynamical N-body simulations of a planet embedded in a planetesimal disc. To simulate the gravitational interactions between the planet and particles we use the N-body software package REBOUND (Rein & Liu 2012), using the hybrid integrator Mercurius5 that switches from a fixed to a variable time- step when a particle is within a given distance from the planet (here chosen to be 8 Hill radii). The fixed time step is chosen to be 4% of the planet's period, which is lower than 17% of the orbital period of all particles in the simulation. Although the total mass of the disc could be tens of M⊕, and thus comparable with the mass of the simulated planets, we assume particles have zero or a negligible mass. In §4.3 we discuss the effect of considering a massive planetesimal disc on planet disc interactions. Particles are initially randomly distributed in the sys- tem with a uniform distribution in semi-major axis (a) be- tween 20 and 170 au (i.e. with a surface density proportional to r−1), with an eccentricity and inclination uniformly dis- tributed between 0-0.02 and 0-0.01 radians, respectively. We use a total number of 104 particles, sufficient to sample the 150 au span in semi-major axes and recover smooth images of the disc density distribution. The planet is placed on a circular orbit at 80 au and assumed to have a bulk density of 1.64 g cm−3 (Neptune's density). We integrate the evolution of the system up to 200 Myr, roughly the upper limit for the age of the system. We run a set of simulations with planet masses varying from 10 to 100 M⊕ with a 5 M⊕ spacing. In order to translate the outcome of these simulations to density distributions used to produce synthetic images, we populate the orbits of each particle with 200 points ran- domly distributed uniformly in mean anomaly as in Pearce & Wyatt (2014), but in the frame co-rotating with the planet in order to see if there are resonant structures (e.g. loops or Trojan regions). We find however no significant azimuthal structures due to first or second order resonances in the de- rived surface density of particles. We weight the mass of each particle based on its initial semi-major axis to impose an initial surface density proportional to rγ between a min- imum and maximum semi-major axis (amin, amax). Finally, because we only run simulations with a fixed planet semi- major axis of 80 au to save computational time, we scale all the distances in the output of the simulation and leave ap as a free parameter (only varying roughly between 75-80 au). This linear scaling is motivated by the fact that some of the features that we are interested in should scale with semi- major axis (e.g. Hill radius, mean-motion resonances, chaotic zone), although some other important quantities, such as the scattering diffusion timescale (Tremaine 1993), have a dependency on a that is different from linear. In Appendix 5 We also used Hermes in several trial runs; however, we decided to use Mercurius instead because with Hermes we obtained re- sults that differ significantly from other integrators such as ias15, whfast and the hybrid integrator within Mercury. We found that some particles outside the chaotic zone were driven to unstable orbits and did not conserve their Tisserand parameter. The pos- sibility of a potential bug in Hermes was confirmed by private communication with Hanno Rein. A we test the validity of the scaling approximation for the narrow range of planet semi-major axis that we explore. In Figure 5 we present, as an example, the evolution of a, e and the surface density of particles for planet masses of 10, 30 and 90 M⊕. The width of the chaotic zone is overlayed in grey (note that this is a region in semi-major axis rather than radius), to compare with the gap in the surface density cleared by the planet. Although by 100 Myr the chaotic zone is almost empty of material (except for particles in the co- rotation zone) the surface density is not zero inside the gap as some particles have apocentres or pericentres within this region while being scattered by the planet. Some particles also remain on stable tadpole or Trojan orbits until the end of the simulation, creating an overdensity within the gap at 80 au. Interior and exterior to the planet, a small fraction of particles are in mean motion resonance with the planet and have their eccentricities increased. In agreement with previous work, we find that the gap's width approximates to the chaotic zone, and its width does not vary after 10 Myr for the planet masses explored. To compare with observations, we use snapshots of the simulations at 50, 100 and 200 Myr since the age of the sys- tem is uncertain and could vary roughly within this range. These ages assume that the putative planet formed (or grew to its current mass) early during the evolution of this sys- tem, rather than recently, or that this is the time since the planet formed. For each assumed epoch, we explore the pa- rameter space using the same MCMC technique as in §3.1, varying amin and amax, the surface density exponent γ, the semi-major axis of the planet, and its mass by interpolat- ing the resulting surface density between two neighbouring simulations. We also leave as free parameters the disc ori- entation, pointing offsets and spectral index, and instead of varying the disc scale height we assume a flat disc. We find a best fit mass of 30±5 M⊕ and semi-major axis of 77±1 au, independently of the assumed age since after 10 Myr there is no significant evolution in the orbits of particles near the planet (see Figure 5). We find, however, that our best fit cannot reproduce well the width and depth of the gap. This is illustrated in the top and middle panels in Figure 6, where the model has a narrower gap compared with the band 6 ra- dial profile, and thus overall larger residuals. Although larger planet masses could produce wider gaps, these would also be significantly deeper and thus inconsistent with the observa- tions. Moreover, the planet gap could be even deeper if, for example, we assume a different starting condition with a de- pleted surface density within the planet's feeding zone as it accreted a large fraction of that material while growing. To test this, we repeat the fitting procedure, but removing those particles that start the simulation within two Hill radii from the planet's orbit (planet's feeding zone). We find a slightly lower best fit planet mass of 26 ± 3 M⊕, but overall the fit is worse with a difference of ∼ 300 in the total χ2, strongly preferring the model with particles starting near the planet. Therefore, we conclude that a single planet on a circular or- bit that was born within the outer disc is unable to explain the ALMA observations. MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 11 Figure 5. Evolution of massless particles in a system with a single planet on a circular orbit at 80 au. Left column: eccentricity and semi-major axis of particles after 0, 10 and 100 Myr of evolution. The dotted and dashed vertical lines represent first order mean motion resonances and the semi-major axis of the planet. Right column: Surface brightness of particles assuming an initial surface density proportional to r1/4 and a dust temperature profile decreasing with radius as r−1/2. The grey shaded region represents the chaotic zone approximated by ap ± 1.5ap(Mp/M(cid:63))2/7. The top, middle and bottom panels show systems with planet masses of 10, 30 and 90 M⊕, respectively. 4 DISCUSSION 4.1 The gap's origin As discussed in §1, we identify multiple scenarios where a single planet might open a gap in a planetesimal disc. Below we discuss these and how well they could fit the data. 4.1.1 Single planet on a circular orbit As we showed in §3.2, a single planet on a circular orbit and formed inside the gap could clear its orbit of debris through scattering, opening a gap with a width similar to the chaotic zone. Such a planet, however, needs to be very massive to produce a ∼ 40 au wide gap ((cid:38) 3 MJup), which, on the other hand, results in a gap that is significantly deeper than our observations suggest. Based on these constraints we find that MNRAS 000, 1–19 (2015) a 30 M⊕ planet is the best trade-of between the gap's width and depth. If, however, the planet could migrate (inwards or out- wards) then these two observables could become compati- ble. A low mass planet that migrated through the disc could carve a sufficiently wide gap to explain the ∼ 40 au gap's width. Migration could also explain the relative depth of ∼ 50% that is observed as after migrating, the planet would also leave behind debris that was scattered onto excited or- bits, but that do no longer cross the planet's orbit (e.g. Kirsh et al. 2009), therefore the gap would still contain a signifi- cant fraction of the original material. Planet migration could be induced by planetesimal scattering, which is discussed in §4.3. 0501001502000.00.10.2EccentricityMp=10M⊕Mp=10M⊕Mp=10M⊕T=0MyrT=10MyrT=100Myr0501001502000.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale]T=0MyrT=10MyrT=100Myr0501001502000.00.10.2EccentricityMp=30M⊕Mp=30M⊕Mp=30M⊕0501001502000.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale]050100150200Semi-majoraxis[au]0.00.10.2EccentricityMp=90M⊕Mp=90M⊕Mp=90M⊕050100150200Radius[au]0.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale] 12 S. Marino et al. Figure 6. Top: Average intensity profile computed azimuthally averaging the disc emission over ellipses oriented as the disc in the sky. The blue and orange lines are obtain from the band 6 and band 7 Clean images (continuous lines) and simulated Clean im- ages based on an N-body simulation of one 30 M⊕planet (dashed lines), using Briggs (robust=0.5) weights. The black continuous line represents the surface brightness profile with a 1 au resolution and displayed at an arbitrary scale. The vertical dotted line repre- sents the orbital radius of the planet. Middle: Azimuthally aver- aged residuals in band 6. Bottom: Azimuthally averaged residuals in band 7. The shaded areas represent the 68% confidence region in the top panel and 99.7% confidence region in the middle and bottom panel, over a resolution element (18 au for band 6 and 13 au for band 7). 4.1.2 Multiple planets on circular orbits If we assume planet migration is negligible (e.g. disc mass around the planet's orbit is much lower than its mass), then multiple planets would need to be present to carve such a wide and shallow gap. Let us assume that the gap was carved by multiple equal mass planets orbiting between 60 and 90 au. As shown by Shannon et al. (2016) and in §3.2, the width of the gap and age of the system place tight con- straints on the minimum mass of a planet to clear the re- gion surrounding its orbit, and on the maximum number of planets that could be orbiting within the gap based on a stability criterion. Given HD 107146's age limits of ∼ 50 and 200 Myr, only planets with masses greater than 10 M⊕ would have enough time to clear their orbits via scattering. On the other hand, only planet masses (cid:46) 30 M⊕ can create a gap that is not too deep compared with our observations. Given this range of planet masses, we estimate that a maximum of three 10 M⊕ planets could orbit between 60 and 90 au spaced by 8 mutual Hill radii at the limits of long term sta- bility (Chambers et al. 1996; Smith & Lissauer 2009). If we require planets to be spaced by more than 8 mutual Hill radii, this multiplicity reduces considerably to two or only one planet if planet masses are slightly higher. We tested this by running new simulations with the same parameters as in §3.2, but instead with a pair of 30 or 10 M⊕ planets and semi-major axes ranging between 60 to 90 au. After a few iterations we found a good fit using two 10 M⊕ planets with semi-major axes of 60 and 83 au (i.e. spaced by 12 Mutual Hill radii). Figure 7 shows the radial profile for this model reproducing the width and depth of the gap. The relative depth of the gap is similar to the one observed as there are still particles on stable orbits at around 70 au and trapped in the co-orbital regions of the two planets - note that there are residuals above 3σ at the inner and outer edge of the disc because this model has sharp boundaries in semi-major axis. If we assumed that no particles were present near the planet at the start of the simulation as commented in §3.2, then even lower mass planets creating narrower gaps would be needed in order to achieve an overall gap depth of 50%. 4.1.3 Planet(s) on eccentric orbits Although a single or multiple low mass planets might ex- plain the observed gap, the formation of an ice giant planet at ∼ 80 au encounters significant difficulties compared to within a few tens of au (see §4.2). The scenario proposed by Pearce & Wyatt (2015), where a planet opens a broad gap through secular interactions, circumvents some of these problems as the planet is formed closer in at 10-20 au, and scattered out by a more massive planet onto an eccentric or- bit with a larger semi-major axis. This scenario was able to fit the mean radial profile derived by Ricci et al. (2015a) and predicted the presence of asymmetries in the form of spiral features and a small offset between the inner and outer re- gions of the disc due to secular interactions with the planet. However, no significant offset between the inner and outer regions is present in our observations. By fitting an ellipse to the inner and outer bright arcs in the band 6 image (roughly at 50 and 110 au), we constrain the offset to be 0.(cid:48)(cid:48)05± 0.(cid:48)(cid:48)02, (i.e. consistent with zero and lower than 1.6 au). This trans- lates to a maximum eccentricity of 0.03 if we assume that the outer bright arc is circular while the inner arc is eccen- tric and thus offset from the star. Moreover, our observations suggest that the disc inner edge is much steeper than pre- dicted by Pearce & Wyatt (2015), thus inconsistent with their model. A last recently proposed scenario involves two interior planets with low but non-zero eccentricity, which open a gap in the outer regions due to secular resonances (see Yelverton et al. in prep). Although this scenario can produce an observable gap, in its simplest form the model produces a gap that is narrower than seen for HD 107146. 4.1.4 Dust-gas interactions There are other scenarios that might not require the pres- ence of planets to produce a gap or multiple ring structures in the dust distribution. Photoelectric instability (Klahr & Lin 2005; Besla & Wu 2007; Lyra & Kuchner 2013; Richert et al. 2017) is one of those scenarios and has received par- ticular attention lately as the presence of vast amounts of MNRAS 000, 1–19 (2015) 050100150Intensity[µJybeam−1]B6(1.1mm)30M⊕-100Myr−10010050100150200Radius[au]−25025B7(0.87mm)30M⊕-100Myr ALMA observations of HD 107146 13 4.2 Planet formation at tens of au Based on these new ALMA observations, if the gap was cleared by a single or multiple planets, these must have a low mass ((cid:46)30 M⊕) and formed between 50 and 100 au. Oth- erwise, if the planets were formed closer in and scattered out onto an eccentric orbit, the disc would appear asym- metric (Pearce & Wyatt 2015). Such low planet masses and orbits at tens of au resemble the ice giant planets in the Solar System, but with a semi-major axis 2-3 times larger than Neptune's. Could such planets have formed in situ as these observations suggest? HD 107146's broad and massive debris disc indicates that planetesimals efficiently formed at a large range of radii from 40 to 140 au. Moreover, the mass of these putative planets ((cid:46) 30 M⊕) is consistent with the solid mass available in their feeding zones (∼ 4 Hill radii wide), based on the dust surface density derived in §3.1 and extrapolating it to the total mass surface density (assum- ing a size distribution dN ∝ D−3.5dD and planetesimals up to sizes of 10 km). Nevertheless, in situ formation encoun- ters the two following problems. First, although a planet at ∼ 80 au might grow through pebble accretion fast enough to form an ice giant before gas dispersal (Johansen & Lac- erda 2010; Ormel & Klahr 2010; Lambrechts & Johansen 2012; Morbidelli & Nesvorny 2012; Bitsch et al. 2015), it re- quires the previous formation of a massive planetesimal (or protoplanet) of 10−2 − 10−1 M⊕ (so-called transition mass). However, newly born planetesimals through streaming insta- bility (Youdin & Goodman 2005) have characteristic masses of rather 10−6 − 10−4 M⊕ (Johansen et al. 2015; Simon et al. 2016). These planetesimals can grow through the accretion of pebbles and smaller planetesimals, but this growth (so- called Bondi accretion) is very slow for low mass bodies at large stellocentric distances (Johansen et al. 2015). A possi- ble solution to this problem is that the protoplanetary disc around HD 107146 was unusually massive and long-lived, which would increase the chances of forming such a proto- planet. Alternatively, the low mass protoplanet could have formed closer in and been scattered out and circularised dur- ing the protoplanetary disc phase. The second problem that in-situ formation faces is related to planet migration. A pro- toplanet growing to form an ice giant is expected to mi- grate inwards through type-I migration (Tanaka et al. 2002). While this might imply that the single or multiple putative planets might have formed at larger radii, our observations suggest that they would have attained most of their mass within the observed gap, i.e. within 100 au. If pebble accre- tion is fast enough, the planet could grow fast and migrate only slightly before disc dispersal. Why would planets form only between 60-90 au within this 100 au wide disc of planetesimals? While no planets might have formed beyond 90 au as planetesimal accretion was too slow for a planetesimal to reach the transition mass and efficiently accrete pebbles, this is not the case for plan- etesimals formed at smaller radii between the inner edge of the disc and the orbit of the innermost putative planet. Why did only planetesimals form between 40-60 au? Planetesimal growth might have been hindered there if the orbits of large solids were stirred by inner planets, making their accretion rate slower. Alternatively, planetesimal growth between 60- 90 au could have been more efficient due to a local enhance- ment in the available solid mass. As stated before, it is also Figure 7. Top: Average intensity profile computed azimuthally averaging the disc emission over ellipses oriented as the disc in the sky. The blue and orange lines are obtain from the band 6 and band 7 Clean images (continuous lines) and simulated Clean im- ages based on an N-body simulation of two 10 M⊕planets (dashed lines), using Briggs (robust=0.5) weights. The black continuous line represents the surface brightness profile with a 1 au resolution and displayed at an arbitrary scale. The vertical dotted line repre- sents the orbital radii of the planets. Middle: Azimuthally aver- aged residuals in band 6. Bottom: Azimuthally averaged residuals in band 7. The shaded areas represent the 68% confidence region in the top panel and 99.7% confidence region in the middle and bottom panel, over a resolution element (18 au for band 6 and 13 au for band 7). gas around a few debris discs suggests that this mechanism might be common. However, this mechanism is only impor- tant when dust-to-gas ratios are comparable and it is not yet clear if relevant for the dust distribution of large millimetre- sized grains that are not well coupled to the gas. Assuming some residual primordial gas could still be present in the disc, we convert our CO gas mass upper limit of 5×10−6 M⊕ (see §4.5 below) to a total gas mass upper limit of ∼ 0.05 M⊕ (assuming a CO/H2 ratio of 10−4) or a surface density of ∼ 10−5 g cm−2. This upper limit is comparable to the to- tal dust mass in millimetre grains, thus photoelectric insta- bility could occur. However, using this gas surface density upper limit we estimate a Stokes number of 105 for millime- tre grains, thus the stopping time is much longer than the collisional lifetime of millimetre dust and of the order of the age of the system. Therefore, we conclude that photoelectric instability does not play an important role in the formation of the structure observed by ALMA around HD 107146. MNRAS 000, 1–19 (2015) 050100150Intensity[µJybeam−1]B6(1.1mm)10M⊕-100Myr−10010050100150200Radius[au]−25025B7(0.87mm)10M⊕-100Myr 14 S. Marino et al. possible that the low mass protoplanet did not form in situ, but it was scattered out from further in. 4.3 Massive planetesimal disc Here we discuss the effect that a massive planetesimal disc could have on the conclusions stated above where we as- sumed a disc of negligible mass ((cid:28) 10 M⊕). As shown in §3, this debris disc is likely very massive and thus affects the dynamics of this system, e.g. because of the gravitational force on the planet and disc self-gravity. A massive disc can induce planetesimal driven migration where the planet mi- grates through a planetesimal disc due to the angular mo- mentum exchange in close encounters (e.g. Fernandez & Ip 1984; Ida et al. 2000; Gomes et al. 2004). This type of mi- gration has been well studied in the context of the outer Solar System, as it could have driven an initially compact orbital configuration to a more extended and current con- figuration (Hahn & Malhotra 1999, 2005), or towards an or- bital instability (e.g. Tsiganis et al. 2005). In all these mod- els the outer planets scatter material in to Jupiter which ejects most of it, leading to an outward migration of the outer planets and a small inward migration of Jupiter. How- ever, Ida et al. (2000) showed that even in the absence of interior planets self-sustained migration could have led Nep- tune's orbit to expand due to the asymmetric planetesimal distribution around it. A more recent work by Kirsh et al. (2009), however, showed that a single planet embedded in a planetesimal disc migrates preferentially inwards due to the timescale difference between the inner and outer feeding zones (see also Ormel et al. 2012). Therefore, we expect that the putative planet at ∼80 au around HD 107146 is likely migrating or has migrated inwards since it formed. Simple scaling relations from Ida et al. (2000) predict that the migration rate should be of the order of (cid:12)(cid:12)(cid:12)(cid:12) da dt (cid:12)(cid:12)(cid:12)(cid:12) ≈ 4πΣa2 M(cid:63) a T (3) which agrees roughly with numerical simulations (e.g. Kirsh et al. 2009). Given the expected surface density of planetes- imals at 80 au around HD 107146 (Σ (cid:38) 10−2 M⊕ au−2 for a maximum planetesimal size of 10 km), Equation 3 pre- dicts that the migration rate is such that the planet would have crossed the whole disc reaching its inner edge in only a few Myr. Kirsh et al. (2009) found, however, that when the planet mass exceeds that of the planetesimals within a few Hill radii, the migration rate decreases strongly with planet mass. Assuming that the gap is indeed caused by a single or multiple planets between 60-90 au, given the 30 M⊕ upper limit for the planet mass we conclude that the surface den- sity of planetesimals must be much lower than 0.01 M⊕ au−2 to hinder planetesimal driven migration, i.e. a total disc mass (cid:46) 430 M⊕. If not the surface density profile would probably be significantly different without a well defined gap. This dynamical upper limit together with the lower limit derived from collisional models constrain the total disc mass to be between ∼ 300 − 400 M⊕, which is close to the maximum solid mass available in a protoplanetary disc under standard assumptions (e.g. disc-to-stellar mass ratio of 0.1 and gas-to- dust mass ratio of 100). This particularly high disc mass at the limits of feasibility is not unique, but a confirmation of the so-called "disc mass problem" as many other young and bright discs need similar or even higher masses according to collisional evolution models (see discussion in Krivov et al. 2018). Although we expect that the gap produced by a mi- grating planet might look significantly different compared with the no migration scenario (e.g. wider and radially and perhaps azimuthally asymmetric), none of the above stud- ies provided a prediction for the resulting surface density of particles during planetesimal driven migration that we could compare with our observations. In experimental runs with a 10 or 30 M⊕ planet and a similar mass planetesimal disc (using massive test particles) we find that the planet mi- grates inward ∼5-40 au in 100 Myr depending on the planet and disc mass, and that the gap has an asymmetric radial profile that is significantly distinct from the no-migration scenario, which could explain why we find a slight asymme- try within the gap (see §2.1). Future comparison with nu- merical simulations of planetesimal-driven migration could provide evidence for inward or outward migration, and thus set tighter constraints on the disc and planet mass. Dynam- ical estimates of the disc mass could also shed light on the disc mass problem, confirming the high mass derived from collisional models or rather indicating that these need to be revisited. 4.4 Warm inner dust component While IR excess detections can be generally explained by a single temperature black body, there is a large number of systems that show evidence for a broad range of tem- peratures that are hard to explain simply as due to a single dusty narrow belt, even with temperatures varying as a func- tion of grain size (e.g. Backman et al. 2009; Morales et al. 2009; Chen et al. 2009; Ballering et al. 2014; Kennedy & Wyatt 2014). Although this might be explained by material distributed over a broad range of radii (like in protoplane- tary discs), an alternative and very attractive explanation for these systems is the presence of a two-temperature disc, with an inner asteroid belt and an outer exo-Kuiper belt, analogous to the Solar System (e.g. Kennedy & Wyatt 2014; Schuppler et al. 2016; Geiler & Krivov 2017). HD 107146 is a good example of this type of system (Ertel et al. 2011; Morales et al. 2011), with a significant excess at 22 µm that cannot be reproduced by models of a single outer belt, but rather is indicative of dust located within ∼ 30 au produced in an asteroid belt. These type of belts are typically hard to resolve due to the small separation to the star (hinder- ing scattered light observations) where current instruments are not able to resolve and due to their low emissivity that peaks in mid-IR. Is the detected inner emission related to the ∼ 20 µm excess? In order to test if the inner emission seen by ALMA is compatible with being warm dust, we use the paramet- ric model developed in §3.1 to see if it can reproduce the available photometry of this system, including the mid-IR excess. We introduce, though, two changes to the model as- sumptions: first, we modify the size distribution index from -3.5 to -3.36 to be consistent with the derived spectral index in this work and previous studies (Ricci et al. 2015b); and second, we extend the size distribution from 1 to 5 cm. The latter is necessary to reproduce the photometry at 7 mm as the contribution from cm-sized grains is significant at these MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 15 Figure 9. Clean image of HD 107146 at 1.1 mm (band 6) us- ing natural weights. Overlayed in contours are the residuals after subtracting the best fit model presented in §3.1.2, but without the inner component. Contour levels are set to 4, 8 and 12 σ. The stellar position is marked with a white cross at the center of the image, while the beam (0.(cid:48)(cid:48)80× 0.(cid:48)(cid:48)79) is represented by a white ellipse in the bottom left corner. The image rms at the center is 6.3 µJy beam−1. For a better display we have adjusted the color scale with a minimum of 50% the peak flux. to be higher than 0.5. Assuming the disc eccentricity cor- responds to the forced eccentricity (i.e. free eccentricities are much smaller), then the perturbing planet should have a very high eccentricity (cid:38) 0.5. One potential problem with this scenario is that the outer disc does not show any hint of being influenced by an eccentric planet. This problem could be circumvented if the true eccentricity of the inner belt is lower than derived based on the image residuals (e.g. (cid:46) 0.2) and the inner planet has a semi-major axis of only a few au, as the forced eccentricity on the outer disc would be much smaller. Alternatively, if the disc or putative outer planet(s) are much more massive than the inner eccentric planet, the disc might remain circular as it is observed. A second potential scenario relates to a recent colli- sion between planetary embryos, which would release large amounts of dust at the collision point producing an asym- metric dust distribution that could last for ∼ 1000 orbits or ∼ 1 Myr at 20 au (Kral et al. 2013; Jackson et al. 2014). In such a scenario the pinch point would appear brightest since the orbits of the generated debris converge where the impact occurred and more debris is created from collisions. Thus, the pinch point would appear radially narrow. How- ever, both our band 6 and 7 datasets show that the emission is radially broad at its brightest point, spanning (cid:38) 20 au. Figure 8. Spectral Energy Distribution of HD 107146 (dark blue points, Kennedy & Wyatt 2014) and its inner component (light blue colours) obtained after subtracting the outer component of our model presented in §3.1.2. The grey, light blue and red lines represent the stellar, outer disc, inner component contribution to the total flux (black line), respectively. wavelengths. In Figure 8 we compare the model and ob- served SED, including the new ALMA photometric points of the inner component. Despite the simplicity of our model, it reproduces successfully both the photometry at 22 µm and at millimetre wavelengths, confirming that our detection is consistent with being warm dust emission. These new ALMA observations provide unique infor- mation on the nature of the warm dust emission as it is marginally resolved (see Figure 9). The surface brightness peaks at ∼ 20 au, thus roughly in agreement with the pre- dicted location based on SED modelling. The emission, how- ever, is far from originating in an axisymmetric asteroid belt, but is rather asymmetric with a maximum brightness towards the south west side of the disc. What could cause such an asymmetric dust distribution? We identify three sce- narios proposed in the literature that could cause long- or short-term brightness asymmetries in a disc or belt. First, the inner emission could correspond to an eccen- tric disc, which at millimetre wavelengths would be seen brightest at apocentre. This is known as apocentre glow, which is caused by the increase in dust densities at apoc- entre for a coherent disc (Wyatt 2005; Pan et al. 2016). Disc eccentricities can be caused by perturbing planets (e.g. Wyatt et al. 1999; Nesvold et al. 2013; Pearce & Wyatt 2014), as it has been suggested to explain Fomalhaut's ec- centric debris disc (Quillen 2006; Chiang et al. 2009; Kalas et al. 2013; Acke et al. 2012; MacGregor et al. 2017) and HD 202628 (Krist et al. 2012; Thilliez & Maddison 2016; Faramaz et al. in prep). At long wavelengths, the contrast between apocentre and pericentre brightness is expected to be approximately (Pan et al. 2016) (cid:18) 1 − e/2 (cid:19)(cid:18) 1 + e (cid:19) , 1 − e 1 + e/2 thus to reach a contrast higher than 2 (as observed for HD 107146's inner disc), the disc eccentricity would need MNRAS 000, 1–19 (2015) 100101102103Wavelength [µm]10-310-210-1100Flux density [Jy]Cold dustWarm dustStar 16 S. Marino et al. This could be circumvented if the collision occurred within a broad axisymmetric disc. Higher resolution observations are necessary to discard this scenario or confirm these two inner components. Finally, a third possible scenario is that the asymmetric structure is caused by planetesimals trapped in mean mo- tion resonances (typically 3:2 and 2:1) with an interior planet that migrated through the planetesimal disc (Wyatt & Dent 2002; Wyatt 2003, 2006; Reche et al. 2008). The exact dust spatial distribution depends on the planet mass, migration rate and eccentricity. The single clump inferred from our observations suggests that planetesimals would be trapped predominantly on the 2:1 resonance rather than 3:2 as the latter only creates a two clump symmetric structure. Simu- lations by Reche et al. (2008) showed that in order to trap planetesimals in the 2:1 resonance, a Saturn mass planet (or higher) with a very low eccentricity was needed, therefore, placing a lower limit on the planet mass if the asymmetry is due to resonant trapping. Future observations could readily distinguish between scenarios 1-2 and 3, as in the first two scenarios the orien- tation of the asymmetry should stay constant (precession timescales are orders of magnitude longer than the orbital period), while in the third scenario it should rotate at the same rate as the putative inner planet orbits the star. More- over, higher resolution and more sensitive observations could reveal if the disc is eccentric, smooth with a radially nar- row clump, or smooth with a radially broad clump (scenar- ios 1 to 3, respectively). Observations at shorter wavelength would also be useful. In the eccentric disc scenario we expect the disc to be eccentric and broader due to radiation pres- sure on small dust grains, while resonant structure would be completely absent and the disc should look axisymmetric as small grains are not trapped also due to radiation pressure. Finally, future ALMA observations should be able to defi- nitely rule out the possibility of the inner emission arising from a sub-mm galaxy. Given HD 107146's proper motion (-174 and -148 mas yr−1 in RA and Dec. direction, respec- tively), we would expect that in 2 years any background ob- ject should shift by 0.(cid:48)(cid:48)23 towards the north east with respect to HD 107146, thus enough to be measured with ALMA ob- servations of similar sensitivity and resolution to the ones presented here. 4.5 Gas non detections In §2.2 we search for CO and HCN secondary origin gas around HD 107146. Although we did not find any, here we use the flux upper limits, including also the 40 mJy km s−1 upper limit for CO J=2-1 from Ricci et al. (2015a), to de- rive an exocometary gas mass upper limit in this disc. It has been demonstrated that in the low density environments around debris discs, gas species are not necessarily in local thermal equilibrium (LTE), which typically leads to an un- derestimation of CO gas masses (Matr`a et al. 2015). Here we use the code developed by (Matr`a et al. 2018b) to esti- mate the population of the CO rotational level in non-LTE based on the radiation environment, densities of collisional partners, and also taking into account UV fluorescence. We consider as sources of radiation the star, the CMB and the mean intensity due to dust thermal emission within the disc (calculated using our model and RADMC-3D). We choose Figure 10. CO gas mass upper limit based on CO J=3-2 (con- tinuous line) and 2-1 (dashed line) flux upper limits, and as a function of electron densities and gas kinetic temperature. The blue, purple and red lines represent gas kinetic temperatures of 20, 50 and 80 K, respectively. 80 au as the representative radius, which is approximately the middle radius of the disc, and left as a free parameter the density of collisional partners (here assumed to be elec- trons). Based on our model, the dust temperature should roughly vary between 50 to 30 K between the disc inner and outer edge. Figure 10 shows the CO gas mass upper limit as a function of electron density for different kinetic temperatures. We find that for electron densities lower than ∼ 102 cm−3, the CO J=2-1 upper limit is more constraining than J=3-2 and vice versa. Overall, the CO gas mass must be lower than 5×10−6 M⊕. We then use equation 2 in Matr`a et al. (2017b) and the estimated photodissociation timescale of 120 yr to estimate an upper limit on the CO+CO2 mass fraction of planetesimals in the disc, but we find no mean- ingful constraint because this CO gas mass upper limit is much higher than the predicted 5 × 10−7 M⊕ if gas were re- leased in collisions of comet-like bodies (e.g. Marino et al. 2016; Kral et al. 2017; Matr`a et al. 2017a,b). In the absence of a tool to calculate the population of rotational levels for HCN, we estimate a mass upper limit assuming LTE. For temperatures ranging between 30-50 K, we find an upper limit of 3×10−9 M⊕, which translates to an upper limit on the mass fraction of HCN in planetesimals of 3%, an order of magnitude higher than the observed abun- dance in Solar System comets (Mumma & Charnley 2011). Note that this HCN upper limit could be much higher due to non-LTE effects, thus this 3% limit must be taken with caution. 5 CONCLUSIONS In this work, we have analysed new ALMA observations of HD 107146's debris disc at 1.1 and 0.86 mm, to study a pos- sible planet-induced gap suggested by Ricci et al. (2015a) with a higher resolution and sensitivity. These new obser- vations show that HD 107146, a 80-200 Myr old G2V star, is surrounded by a broad disc of planetesimals from 40 to MNRAS 000, 1–19 (2015) 10−1100101102103104105106Electrondensity[cm−3]10−610−5COgasmass[M⊕]CO(3-2)CO(2-1)T=20KT=50KT=80K 140 au, that is divided by a gap ∼40 au wide (FWHM), centered at 80 au and 50% deep, i.e. the gap is not devoid of material. We constrained the disc morphology, mass and spectral index by fitting parametric models to the observed visibilities using an MCMC procedure. We find that the disc is consistent with being axisymmetric, and we constrain the disc eccentricity to be lower than 0.03. The observed morphology of HD 107146's debris disc suggests the presence of a planet on a wide circular orbit opening a gap in a planetesimal disc through scattering. We run a set of N-body simulations of a planet embedded in a planetesimal disc that we compare with our observations. We find that the observed morphology is best fit with a planet mass of 30 M⊕, but significant residuals appear after subtracting the best fit model. We conclude that the ob- served gap cannot be reproduced by the dynamical clearing of such a planet as the gap it creates is significantly deeper and narrower than observed. We discuss that this could be circumvented by allowing the planet to migrate (e.g. due to planetesimal driven migration) or by allowing multiple plan- ets to be present. We discuss how a planet could have formed in situ if the primordial disc was massive and long-lived, and possibly grew to its final mass very quickly and by the end of the disc lifetime (e.g. through pebble accretion), avoid- ing significant inward migration and runaway gas accretion. Moreover, because the putative planet(s) could undergo very fast planetesimal driven migration, we set an upper limit on the surface density and total mass of the disc. These ALMA observations also revealed unexpected emission near the star that is best seen when subtracting our best fit parametric model of the outer disc. This inner component has a total flux of 0.8 and 0.3 mJy at 0.86 mm and 1.1 mm, respectively, a peak intensity that is signifi- cantly offset from the star by 0.(cid:48)(cid:48)5 (15 au), and we resolve the emission both radially and azimuthally. Its radial loca- tion indicates that it could be the same warm dust that had been inferred to be between 10-15 au to explain HD 107146's excess at 22µm. Indeed, we fit this emission with an extra in- ner asymmetric component finding a good match with these ALMA observations and also with HD 107146's excess at 22µm. We constrain its peak density at 19 au, and a radial width of at least 20 au. We hypothesise that this asymmetric emission could be due to a disc that is eccentric due to inter- actions with an eccentric inner planet, asymmetric due to a recent giant collision, or clumpy due to resonance trapping with a migrating inner planet. On the other hand, we find that this inner emission is unlikely to be a background sub- mm galaxy, as the probability of finding one as bright as 0.8 mJy at 0.86 mm within the disc inner edge (i.e. co-located with the warm dust) is 0.1%. Finally, although it had been demonstrated by Ricci et al. (2015a) that no primordial gas is present around HD 107146, we search for CO and HCN gas that could be released from volatile-rich solids throughout the collisional cascade in the outer disc. However, we find no gas, but we place upper limits on the total gas mass and HCN abun- dance inside planetesimals, being consistent with comet-like composition. MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 17 ACKNOWLEDGEMENTS We thank Bertram Bitsch for useful discussion on planet formation and pebble accretion. JC and VG acknowledge support from the National Aeronautics and Space Adminis- tration under grant No. 15XRP15 20140 issued through the Exoplanets Research Program. MB acknowledges support from the Deutsche Forschungsgemeinschaft (DFG) through project Kr 2164/15-1. VF's postdoctoral fellowship is sup- ported by the Exoplanet Science Initiative at the Jet Propul- sion Laboratory, California Inst. of Technology, under a con- tract with the National Aeronautics and Space Administra- tion. GMK is supported by the Royal Society as a Royal Society University Research Fellow. LM acknowledges sup- port from the Smithsonian Institution as a Submillimeter Array (SMA) Fellow. This paper makes use of the fol- lowing ALMA data: ADS/JAO.ALMA#2016.1.00104.S and 2016.1.00195.S. ALMA is a partnership of ESO (represent- ing its member states), NSF (USA) and NINS (Japan), to- gether with NRC (Canada) and NSC and ASIAA (Taiwan) and KASI (Republic of Korea), in cooperation with the Re- public of Chile. The Joint ALMA Observatory is operated by ESO, AUI/NRAO and NAOJ. Simulations in this paper made use of the REBOUND code which can be downloaded freely at http://github.com/hannorein/rebound. REFERENCES Absil O., et al., 2013, A&A, 555, A104 Acke B., et al., 2012, A&A, 540, A125 Andrews S. M., 2015, PASP, 127, 961 Ardila D. R., et al., 2004, ApJ, 617, L147 Backman D., et al., 2009, ApJ, 690, 1522 Ballering N. P., Rieke G. H., G´asp´ar A., 2014, ApJ, 793, 57 Besla G., Wu Y., 2007, ApJ, 655, 528 Bitsch B., Lambrechts M., Johansen A., 2015, A&A, 582, A112 Boley A. C., 2009, ApJ, 695, L53 Booth M., et al., 2016, MNRAS, 460, L10 Boss A. P., 1997, Science, 276, 1836 Burns J. A., Lamy P. L., Soter S., 1979, Icarus, 40, 1 Carniani S., et al., 2015, A&A, 584, A78 Carpenter J. M., et al., 2009, ApJS, 181, 197 Chambers J. E., Wetherill G. W., Boss A. P., 1996, Icarus, 119, 261 Chatterjee S., Ford E. B., Matsumura S., Rasio F. A., 2008, ApJ, 686, 580 Chen C. H., Sheehan P., Watson D. M., Manoj P., Najita J. R., 2009, ApJ, 701, 1367 Chiang E., Kite E., Kalas P., Graham J. R., Clampin M., 2009, ApJ, 693, 734 Crida A., Masset F., Morbidelli A., 2009, ApJ, 705, L148 Dent W. R. F., et al., 2014, Science, 343, 1490 Draine B. T., 2003, ApJ, 598, 1017 Duncan M., Quinn T., Tremaine S., 1989, Icarus, 82, 402 Eiroa C., et al., 2013, A&A, 555, A11 Ertel S., Wolf S., Metchev S., Schneider G., Carpenter J. M., Meyer M. R., Hillenbrand L. A., Silverstone M. D., 2011, A&A, 533, A132 Faramaz V., Beust H., Augereau J.-C., Kalas P., Graham J. R., 2015, A&A, 573, A87 Faramaz V., et al., in prep. Feldt M., et al., 2017, A&A, 601, A7 Fernandez J. A., Ip W.-H., 1984, Icarus, 58, 109 Ford E. B., Rasio F. A., 2008, ApJ, 686, 621 18 S. Marino et al. Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013, Matthews B. C., Krivov A. V., Wyatt M. C., Bryden G., Eiroa PASP, 125, 306 Gaia Collaboration et al., 2016a, A&A, 595, A1 Gaia Collaboration et al., 2016b, A&A, 595, A2 Geiler F., Krivov A. V., 2017, MNRAS, 468, 959 Golimowski D. A., et al., 2011, AJ, 142, 30 Gomes R. S., Morbidelli A., Levison H. F., 2004, Icarus, 170, 492 Goodman J., Weare J., 2010, Commun. Appl. Math. Comput. Sci., 5, 65 Hahn J. M., Malhotra R., 1999, AJ, 117, 3041 Hahn J. M., Malhotra R., 2005, AJ, 130, 2392 Hayashi C., 1981, Progress of Theoretical Physics Supplement, 70, 35 Hillenbrand L. A., et al., 2008, ApJ, 677, 630 Hughes A. M., Duchene G., Matthews B., 2018, preprint, (arXiv:1802.04313) Ida S., Bryden G., Lin D. N. C., Tanaka H., 2000, ApJ, 534, 428 Jackson A. P., Wyatt M. C., Bonsor A., Veras D., 2014, MNRAS, 440, 3757 Johansen A., Lacerda P., 2010, MNRAS, 404, 475 Johansen A., Lambrechts M., 2017, Annual Review of Earth and Planetary Sciences, 45, 359 Johansen A., Mac Low M.-M., Lacerda P., Bizzarro M., 2015, Science Advances, 1, 1500109 Juri´c M., Tremaine S., 2008, ApJ, 686, 603 Kalas P., et al., 2008, Science, 322, 1345 Kalas P., Graham J. R., Fitzgerald M. P., Clampin M., 2013, ApJ, 775, 56 Kennedy G. M., Wyatt M. C., 2010, MNRAS, 405, 1253 Kennedy G. M., Wyatt M. C., 2014, MNRAS, 444, 3164 Kenyon S. J., Bromley B. C., 2002, ApJ, 577, L35 Kenyon S. J., Bromley B. C., 2008, ApJS, 179, 451 Kenyon S. J., Bromley B. C., 2010, ApJS, 188, 242 Kirsh D. R., Duncan M., Brasser R., Levison H. F., 2009, Icarus, 199, 197 Klahr H., Lin D. N. C., 2005, ApJ, 632, 1113 Kral Q., Th´ebault P., Charnoz S., 2013, A&A, 558, A121 Kral Q., Matr`a L., Wyatt M. C., Kennedy G. M., 2017, MNRAS, 469, 521 Krist J. E., Stapelfeldt K. R., Bryden G., Plavchan P., 2012, AJ, 144, 45 Krivov A. V., Ide A., Lohne T., Johansen A., Blum J., 2018, MNRAS, 474, 2564 Lagrange A.-M., et al., 2009, A&A, 493, L21 Lambrechts M., Johansen A., 2012, A&A, 544, A32 Levison H. F., Thommes E., Duncan M. J., 2010, AJ, 139, 1297 Li A., Greenberg J. M., 1998, A&A, 331, 291 Lieman-Sifry J., Hughes A. M., Carpenter J. M., Gorti U., Hales A., Flaherty K. M., 2016, ApJ, 828, 25 Lyra W., Kuchner M., 2013, Nature, 499, 184 MacGregor M. A., et al., 2017, ApJ, 842, 8 Marino S., et al., 2016, MNRAS, 460, 2933 Marino S., et al., 2017a, MNRAS, 465, 2595 Marino S., Wyatt M. C., Kennedy G. M., Holland W., Matr`a L., Shannon A., Ivison R. J., 2017b, MNRAS, 469, 3518 Marois C., Macintosh B., Barman T., Zuckerman B., Song I., Patience J., Lafreni`ere D., Doyon R., 2008, Science, 322, 1348 Marois C., Zuckerman B., Konopacky Q. M., Macintosh B., Bar- man T., 2010, Nature, 468, 1080 Matr`a L., Pani´c O., Wyatt M. C., Dent W. R. F., 2015, MNRAS, 447, 3936 Matr`a L., et al., 2017a, MNRAS, 464, 1415 Matr`a L., et al., 2017b, ApJ, 842, 9 Matr`a L., Marino S., Kennedy G. M., Wyatt M. C., Oberg K. I., Wilner D. J., 2018a, preprint, (arXiv:1804.01094) Matr`a L., Wilner D. J., Oberg K. I., Andrews S. M., Loomis R. A., C., 2014a, Protostars and Planets VI, pp 521–544 Matthews B., Kennedy G., Sibthorpe B., Booth M., Wyatt M., Broekhoven-Fiene H., Macintosh B., Marois C., 2014b, ApJ, 780, 97 McMullin J. P., Waters B., Schiebel D., Young W., Golap K., 2007, in Shaw R. A., Hill F., Bell D. J., eds, Astronomical Society of the Pacific Conference Series Vol. 376, Astronomical Data Analysis Software and Systems XVI. p. 127 Montesinos B., et al., 2016, A&A, 593, A51 Mo´or A., et al., 2015, ApJ, 814, 42 Mo´or A., et al., 2017, ApJ, 849, 123 Morales F. Y., et al., 2009, ApJ, 699, 1067 Morales F. Y., Rieke G. H., Werner M. W., Bryden G., Stapelfeldt K. R., Su K. Y. L., 2011, ApJ, 730, L29 Morbidelli A., Nesvorny D., 2012, A&A, 546, A18 Mumma M. J., Charnley S. B., 2011, ARA&A, 49, 471 Nesvold E. R., Kuchner M. J., Rein H., Pan M., 2013, ApJ, 777, 144 Ormel C. W., Klahr H. H., 2010, A&A, 520, A43 Ormel C. W., Ida S., Tanaka H., 2012, ApJ, 758, 80 Pan M., Nesvold E. R., Kuchner M. J., 2016, ApJ, 832, 81 Patel B. H., Percivalle C., Ritson D. J., D. D., Sutherland J. D., 2015, Nat Chem, 7, 301 Pearce T. D., Wyatt M. C., 2014, MNRAS, 443, 2541 Pearce T. D., Wyatt M. C., 2015, MNRAS, 453, 3329 Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62 Quillen A. C., 2006, MNRAS, 372, L14 Rafikov R. R., 2004, AJ, 128, 1348 Rameau J., et al., 2013, ApJ, 772, L15 Reche R., Beust H., Augereau J.-C., Absil O., 2008, A&A, 480, 551 Rein H., Liu S.-F., 2012, A&A, 537, A128 Ricci L., Carpenter J. M., Fu B., Hughes A. M., Corder S., Isella A., 2015a, ApJ, 798, 124 Ricci L., Maddison S. T., Wilner D., MacGregor M. A., Ubach C., Carpenter J. M., Testi L., 2015b, ApJ, 813, 138 Richert A. J. W., Lyra W., Kuchner M., 2017, preprint, (arXiv:1709.07982) Schneider G., et al., 2014, AJ, 148, 59 Schuppler C., Krivov A. V., Lohne T., Booth M., Kirchschlager F., Wolf S., 2016, MNRAS, 461, 2146 Shannon A., Bonsor A., Kral Q., Matthews E., 2016, MNRAS, 462, L116 Silverstone M. D., 2000, PhD thesis, UNIVERSITY OF CALI- FORNIA, LOS ANGELES Simon J. B., Armitage P. J., Li R., Youdin A. N., 2016, ApJ, 822, 55 Simpson J. M., et al., 2015, ApJ, 799, 81 Smith A. W., Lissauer J. J., 2009, Icarus, 201, 381 Su K. Y. L., et al., 2006, ApJ, 653, 675 Su K. Y. L., et al., 2009, ApJ, 705, 314 Su K. Y. L., et al., 2017, AJ, 154, 225 Sutherland J. D., 2017, Nature Reviews Chemistry, 1, 12 Tanaka H., Takeuchi T., Ward W. R., 2002, ApJ, 565, 1257 Thilliez E., Maddison S. T., 2016, MNRAS, 457, 1690 Thureau N. D., et al., 2014, MNRAS, 445, 2558 Tremaine S., 1993, in Phillips J. A., Thorsett S. E., Kulkarni S. R., eds, Astronomical Society of the Pacific Conference Series Vol. 36, Planets Around Pulsars. pp 335–344 Tsiganis K., Gomes R., Morbidelli A., Levison H. F., 2005, Na- ture, 435, 459 Weidenschilling S. J., 1977, Ap&SS, 51, 153 Williams J. P., Najita J., Liu M. C., Bottinelli S., Carpenter J. M., Hillenbrand L. A., Meyer M. R., Soderblom D. R., 2004, ApJ, 604, 414 Wyatt M. C., Dent W. R. F., 2018b, ApJ, 853, 147 Wisdom J., 1980, AJ, 85, 1122 MNRAS 000, 1–19 (2015) ALMA observations of HD 107146 19 Wyatt M. C., 2003, ApJ, 598, 1321 Wyatt M. C., 2005, A&A, 440, 937 Wyatt M. C., 2006, ApJ, 639, 1153 Wyatt M. C., 2008, ARA&A, 46, 339 Wyatt M. C., Dent W. R. F., 2002, MNRAS, 334, 589 Wyatt M. C., Dermott S. F., Telesco C. M., Fisher R. S., Grogan K., Holmes E. K., Pina R. K., 1999, ApJ, 527, 918 Wyatt M. C., Smith R., Greaves J. S., Beichman C. A., Bryden G., Lisse C. M., 2007, ApJ, 658, 569 Yelverton B., et al., in prep. Youdin A. N., Goodman J., 2005, ApJ, 620, 459 Zheng X., Lin D. N. C., Kouwenhoven M. B. N., Mao S., Zhang X., 2017, ApJ, 849, 98 APPENDIX A: LINEAR SCALING OF SEMI-MAJOR AXIS In §3.2 we assumed that we can approximate the surface density of particles interacting with a planet at a given semi- major axis (between ∼ 75 − 80 au), by linearly scaling the distances in the output of a N-body simulation of a planet on a fixed semi-major axis of 80 au. In order to check the validity of this, we run three additional simulations with a planet semi-major axis of 75 au and masses of 10, 20 and 30 M⊕. Figure A1 compares the resulting surface density of these new simulations (dashed line) with simulations of a planet at 80 au, but scaled to 75 au. These three new simulations show that the linear scaling is a reasonable ap- proximation, matching very well the structure seen in the simulations with a planet at 75 au. Only minor differences are present, which are of the order of the expected noise due to the random initial semi-major axis of particles in our simulations. This paper has been typeset from a TEX/LATEX file prepared by the author. Figure A1. Surface brightness evolution of disc perturbed by a planet on a circular orbit at 75 au. The continuous and dashed lines represent simulations run with a planet with a semi-major axis of 80 au and then scaled to 75 au, and with a planet with a semi-major axis of 75 au without scaling the simulation, respec- tively. Surface brightness of particles assuming an initial surface density proportional to r0.25 and a dust temperature profile de- creasing with radius as r−1/2. The grey shaded region represents the chaotic zone approximated by ap ± 2ap(Mp/M(cid:63))2/7. The top, middle and bottom panels show systems with planet masses of 10, 20 and 30 M⊕, respectively. MNRAS 000, 1–19 (2015) 0501001500.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale]Mp=10M⊕Mp=10M⊕Mp=10M⊕T=0MyrT=10MyrT=100Myr0501001500.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale]Mp=20M⊕Mp=20M⊕Mp=20M⊕050100150Radius[au]0.00.20.40.60.81.01.2Σ(r)×r−1/2[arbitraryscale]Mp=30M⊕Mp=30M⊕Mp=30M⊕
1703.06386
2
1703
2018-11-22T21:50:56
Persistent or repeated surface habitability on Mars during the Late Hesperian - Amazonian
[ "astro-ph.EP" ]
Large alluvial fan deposits on Mars record relatively recent habitable surface conditions ($\lesssim$3.5 Ga, Late Hesperian - Amazonian). We find net sedimentation rate <(4-8) {\mu}m/yr in the alluvial-fan deposits, using the frequency of craters that are interbedded with alluvial-fan deposits as a fluvial-process chronometer. Considering only the observed interbedded craters sets a lower bound of >20 Myr on the total time interval spanned by alluvial-fan aggradation, >10^3-fold longer than previous lower limits. A more realistic approach that corrects for craters fully entombed in the fan deposits raises the lower bound to >(100-300) Myr. Several factors not included in our calculations would further increase the lower bound. The lower bound rules out fan-formation by a brief climate anomaly. Therefore, during the Late Hesperian - Amazonian on Mars, persistent or repeated processes permitted habitable surface conditions.
astro-ph.EP
astro-ph
Key Points: Abstract. 1. Introduction. Persistent or repeated surface habitability on Mars during the Late Hesperian - Amazonian Edwin S. Kite1,*, Jonathan Sneed1, David P. Mayer1, Sharon A. Wilson2 1. University of Chicago, Chicago, Illinois, USA (* [email protected]) 2. Smithsonian Institution, Washington, District of Columbia, USA Published in: Geophys. Res. Lett., 44, doi:10.1002/2017GL072660. • Alluvial fans on Mars did not form during a short-lived climate anomaly • We use the distribution of observed craters embedded in alluvial fans to place a lower limit on fan formation time of >20 Ma • The lower limit on fan formation time increases to >(100-300) Ma when model estimates of craters completely buried in the fans are included Large alluvial fan deposits on Mars record relatively recent habitable surface conditions (≲3.5 Ga, Late Hesperian -- Amazonian). We find net sedimentation rate <(4-8) μm/yr in the alluvial-fan deposits, using the frequency of craters that are interbedded with alluvial-fan deposits as a fluvial-process chronometer. Considering only the observed interbedded craters sets a lower bound of >20 Myr on the total time interval spanned by alluvial-fan aggradation, >103-fold limits. A more than previous realistic approach that corrects for craters fully entombed in the fan deposits raises the lower bound to >(100-300) Myr. Several factors not included in our calculations would further increase the lower bound. The lower bound rules out fan-formation by a brief climate anomaly. Therefore, during the Late Hesperian -- Amazonian on Mars, persistent or repeated processes permitted habitable surface conditions. Large alluvial fans on Mars record one or more river-supporting climates on ≲3.5 Ga Mars (e.g. Moore & Howard 2005, Grant & Wilson 2012, Kite et al. 2015, Williams et al. 2013) (Fig. 1). This climate permitted precipitation-sourced runoff production of >0.1 mm/hr that fed rivers with discharge up to 102 m3/s (Dietrich et al. 2017, Morgan et al. 2014). Such large water discharges probably require a water activity that would permit life. The large (>10 km2) alluvial fans with ≲2° slopes correspond to a relatively recent (Grant & Wilson 2011) epoch of Mars surface habitability (Williams et al. 2013). (Although small (<10 km2) alluvial fans with >10° slopes formed <5 Mya on Mars (e.g. Williams & Malin 2008), in this paper we focus on the large alluvial fans). Did these large fans result from a single anomalous burst of wet conditions, such as might result from an impact or volcanic eruption? Or, do the fans record persistent or repeated wet conditions, for example longer lower 1 as the result of a sustained warmer climate regime? Better constraints on the time span of alluvial fan formation would constrain models of Late Hesperian -- Amazonian climate. sediment Figure 1. Large alluvial fan deposit on Mars (28°S 333°E). The ≲3.5 Ga age of the fans is shown by their relatively low crater density. Aeolian erosion exposes layers and channels within the deposit. Previous work on the duration of the interval of fan build-up used sedimentology to estimate the time over which (e.g., Armitage et al. 2011, transport occurred Williams et al. 2011, Palucis et al. 2014). These sedimentologic methods require assumptions about flow intermittency or sediment:water ratio, which (for almost all Mars alluvial fans) are poorly constrained (Dietrich et al. 2017). Therefore, the lower limits from sedimentological methods are short -- obtained for example, ~3600 years (Morgan et al. 2014). Moreover, brief (1-100 yr) aggradation intervals have been proposed in age and volume fans for deltas on Mars that are similar to the alluvial (Kleinhans et al. 2010, Mangold et al. 2012a, Hauber et al. 2013). Another approach to estimating the interval of alluvial-fan build-up is to measure the density of craters 2 superimposed on different fans, and use the spread of crater-retention ages for the fan surfaces as a proxy for the range of fan-formation ages. This method is not reliable, because crater counts for areas <103 km2 cannot distinguish between ages of (for example) 2.5 Ga and 3.0 Ga (Warner et al. 2015). Therefore, the time span of habitable climates in the Late Hesperian-Amazonian remains an open question. Figure 2. Idealized cross-section through an alluvial fan deposit. Crater density within alluvial fans may be estimated using the frequency of visible interbedded craters at the exposed surface. Case A: If fluvial sediment accumulation was modest, some craters that formed late in the history of alluvial-fan aggradation are only partially buried and are visible today (though outnumbered by postfluvial craters; not shown). Case B: If fluvial accumulation was rapid, the temporal window for partial crater burial was correspondingly brief, and few interbedded craters are visible at the exposed surface. Case C: If postfluvial erosion was severe, the areal density of exposed embedded craters of a given diameter is still proportional to the volumetric density of those craters. In either erosional or well-preserved fan scenarios, assuming steady aggradation, a count of interbedded craters constrains the fan aggradation rate. 3 [1] τraw,D = ND/(fDa) To get a more accurate estimate of the interval over which alluvial fans formed, we used the embedded-crater method (Hartmann 1974, Kite et al. 2013a). For post-Noachian deposits on Mars, this method works as follows. Crater density on a stable planetary surface (a surface that is neither eroding nor aggrading) is proportional to exposure duration. This method may be extended to three dimensions: the total number of craters interbedded within a sedimentary deposit is proportional to the time spanned by active sedimentation (including any hiatuses) (Fig. 2). The greater the volumetric density of craters, the longer the time span of deposition. Of course, many or most of the interbedded craters may be completely buried (Fig. 2); therefore a complete count of interbedded craters is usually impossible. However, if an impact occurs near the end of active sedimentation, then the corresponding crater may be only partially buried. Smaller craters are more readily buried, and larger craters require more sediment to be completely obscured. The time needed to accumulate the population of visibly embedded (synfluvial) craters is where τraw,D is the minimum time required to build up the observed population of embedded craters (with minimum diameter D), ND is the number of observed embedded craters, fD is the past crater flux (#/km2/yr), and a is the count area (km2). Prefluvial craters (which are overlain by fan deposits, but that formed before the start of fluvial deposition; Irwin et al. 2015) are excluded. This procedure gives a strict lower limit on the interval of fan formation; it does not account for craters within the deposit that are fully entombed. To correct for fully entombed craters, we can assume steady aggradation and divide fan thickness Z by best-fit aggradation rate W to get duration of fan formation τsteady: The numerator in [2] corresponds to the required burial depth for obliteration. The amount of burial that is required is constrained by the geometry of small impact craters (Melosh 1989, Watters et al. 2015). φ is the obliteration depth fraction for a given crater, expressed relative to crater diameter. (In this paper, we define a crater as "obliterated" if it can no longer be identified in a high-resolution optical image; Kite & Mayer 2017). The factor of 1.33 corrects for the fact that, for any minimum-diameter D, the median diameter in a count will exceed D -- thus, Dφ is an underestimate of the required burial depth. This correction depends on the crater production function used. The correction is relatively small (1.3× -- 1.5×) in our size range of interest, because crater frequency falls off steeply with increasing diameter, and we represent it by a fixed factor. Post-depositional erosion of fan deposits does not affect the validity of [2-3]. A randomly oriented cut through the crater-containing volume intersects each crater with a probability proportional to that crater's size; the sample of craters partially exhumed at an erosional surface is biased towards larger impacts. Just as with partial burial, the number of craters that are exposed is proportional to the volumetric crater density (and inversely proportional to aggradation rate). Thus, the volumetric density of interbedded craters may be estimated from surface WD ≈ 1.33Dφ/τD τsteady,D = Z/WD [2] [3] 4 2. Methods and results. counts both on pristine fan surfaces and for fans that are deeply eroded (Kite et al. 2013a). Therefore, embedded-crater counts can be used as a Mars fluvial process speedometer. In order to set a lower bound on the time span of alluvial-fan formation, we searched 1.7×104 km2 of previously-catalogued fans (corresponding to most of the surface area of large alluvial fans on Mars; Wilson et al. 2012) using 6m-per-pixel CTX images to scout for candidate embedded craters. Candidate craters show possible evidence of interbedding with paleochannels or other fan deposits. Each candidate feature was reviewed by three of the authors (E.S.K, D.P.M., and J.S.) for final classification. Where available, 25cm-per-pixel HiRISE images and anaglyphs, plus CTX Digital Terrain Models (DTMs) were used to inspect candidates flagged in the initial CTX survey. Each candidate feature was categorized as quality level 1, 2, 3 (representing decreasing confidence that the crater was embedded), or it was discarded. A total of 25 embedded craters were found at <30° latitude (D = 0.08-5.0 km; Fig. 3, Figs. S1-S2). The latitude cut was selected to avoid glaciated craters. Those embedded craters were then classified (usually with the aid of CTX DTMs) as "synfluvial," "uncertain", or "prefluvial." These craters constitute a small fraction of the total number of craters on the surfaces of the fans, most of which appear to be postfluvial (Grant & Wilson 2012). found craters were Figure 3. Embedded synfluvial craters within alluvial fans on Mars from our catalog (Fig. S2). A supplementary HiRISE-only survey (570 km2) was carried out to check for resolution km; Figure effects. 13 S3). embedded (D=0.05 -- 0.22 For 0.08 km < D < 0.16 km, HiRISE embedded-crater densities ND/a are the same (within Poisson error) as CTX embedded-crater densities. Therefore, the HiRISE check provides no evidence that our conclusions would be changed by a HiRISE re-survey of the full area covered by our CTX survey. Diameter measurement error was estimated by blindly remeasuring craters and found to be negligible. Crater diameter change during degradation is ignored. The contribution of false positives to our catalog is likely small. Although polygonal faulting in Earth marine sediments can produce crater-like concentric layering (Tewksebury et al. 2014), this is unlikely for Mars alluvial-fan deposits. For example, embedded craters 5 Crater Name Comments Estimated maximum fan thickness are isolated (not space-filling), and frequently show preserved rims. However, there are certainly false negatives in our survey area: re-survey of a crater of interest found several additional candidates with scores ≤3 on panel inspection. It is possible that many embedded craters are easily identifiable as impact craters in CTX imagery, but have an expression that is indistinguishable from postfluvial craters at CTX scale. Therefore, our embedded crater densities are lower limits. We estimated fan thicknesses by differencing CTX DTM profiles across fans and analogous profiles across parts of the same fan-hosting craters (n=17) that lacked fans (Palucis et al. 2014) (Table 1). We found maximum fan thickness of 1.1km, with thicknesses ~1 km common. 900 m Flat floor approximation. 46W 0N Orson Welles 200 m 45W 24S Degana >200 m Conservative estimate Luba 37W 18S 1.1 km Flat floor approximation. 34W 26S Holden 1 km Flat floor approximation. 28W 26S Ostrov >300 m Conservative estimate "SE of Ostrov" 27W 28S 150 m 15W 25S - 300 m 67E, 22S Harris >1.1 km Flat floor approximation 73E 22S Saheki 500 m Flat floor approximation. "SE of Saheki" 74E 23S 250 m Alternative control gives max. thickness 450 m. 76E 26S Runanga 84E 29S - Majuro 84E 33S 300 m -- 400 m Taken from Mangold et al. 2012b >500 m 134E 1N - 250 m 142E 12N Eddie 300 m 170E 19N Kotka 200 m - 180E 5S Table 1. Alluvial fan thickness estimates (crater-by-crater maximum estimated thickness). Thicknesses ~1 km are shown in bold. The usual procedure for estimating crater-counting error is to use Poisson statistics (Michael et al. 2016). The results of this procedure are shown by the blue lines and blue error bars in Fig. 4. To generate these results, we assumed a fixed crater flux (Michael 2013), no change in atmospheric screening from today's Mars, a strong-rock target strength, and a fixed obliteration depth fraction φ = 0.1. The true error in:- true crater flux (Hartmann 2005, Johnson et al. 2016); target strength (Okubo 2007, Grindrod et al. 2010); filtering by a potentially thicker past atmosphere (Kite et al. 2014); the time of formation of the fans; and the amount of burial or erosion (expressed as a fraction of the crater's diameter) that is needed to prevent the crater from being detected at CTX resolution. larger than this because of uncertainty 3. Analysis. is 900 m 6 in Information). assumed in the lower Specifically, we the Supplementary Therefore, we adopted conservative prior probabilities on these parameters in a Monte Carlo simulation of our lower bound that also includes Poisson error (details are given (1) a factor-of-4 uncertainty in crater flux (log-uniform uncertainty between 0.5× and 2× the fluxes of Michael 2013); (2) log-uniform uncertainty in target strength between limits of 65 kPa and 10 MPa (Dundas et al. 2010); (3) log-uniform uncertainty in paleo- atmospheric pressure between limits of 6 mbar and 1000 mbar; (4) a uniform uncertainty between fan formation 2.0 Ga (low-end fan crater retention age) and 3.6 Ga (ages of large craters that host fans); and (5) a log-uniform prior for obliteration depth fraction (expressed as a fraction of D) from 0.05 (rim burial; Melosh 1989) and 0.2 (original crater depth; Watters et al. 2015). For each Monte Carlo trial, the effect of Poisson error is calculated analytically. Given the observations and the randomly-sampled parameters, each Monte Carlo trial yields an analytic probability for each candidate age (or each candidate aggradation rate) in each size bin. These probabilities are summed over 103 Monte Carlo trials and normalized. Results are shown by the gray bands and black stars in Fig. 4. The best-fit lower limit on the time-span of fan aggradation (Fig. 4a) increases with increasing diameter, as expected. Bins ≥1.4 km contain only 1 crater, and the limit. Poisson uncertainty constitutes most of the total uncertainty For the smaller diameters, the counting-statistics error is small compared to the total uncertainty in the lower limit, but systematic undercounting of embedded craters is most likely for craters that are smaller (and thus more easily buried and modified). For example, for 10 m diameter craters, the nominal time span to build up the observed craters is ~1 Myr, but the amount of burial needed to obliterate the crater is only ~1 m, and given fan thicknesses ~1 km (Table 1), it is very unlikely that there are zero craters entombed within the lower 99.9% of the fans. The largest >1 km-diameter bin contains 2 embedded craters and yields a 95% lower limit of >17 Myr, which we round to >20 Myr. Using the single ~5km crater found in our survey gives a >54 Myr lower bound. Turning to the rate plot (Fig. 4b), constant aggradation rates of <(4-8) μm/year match our data. Different crater diameters probe different depth ranges and thus aggradation rate over different timescales, but our data do not require a change in aggradation rate with timescale (Jerolmack & Sadler 2007). Dividing typical fan thicknesses by this rate gives 125-250 Myr (100-300 Myr to 1 significant figure). For both plots, the Monte Carlo procedure gives limits that are more uncertain, and slightly more permissive (Fig. 4). The time span of liquid water estimated from steady aggradation (>(100-300) Myr; Fig. 4b) is a more realistic estimate of the true fan-forming interval than the time span estimated from observed synfluvial craters (>20 Myr; Fig. 4a). For example, if the observed embedded craters represent the total embedded-crater population, then fan aggradation must have started very fast and decreased sharply near the end of fan build-up. Such an accumulation history would favor small-crater preservation relative to large-crater preservation -- opposite what is observed. Furthermore, after fluvial deposition stopped, many of the fan deposits underwent aeolian erosion (e.g. Fig. 1). Because postfluvial erosion would destroy some embedded craters, the observed embedded craters are very unlikely to represent the total population of craters that formed embedded within the fans. 7 (a) (b) 8 4. Discussion. 4.1. Factors not taken into account would increase our lower limit. Figure 4. (a). Minimum time-span of sedimentation based on observed synfluvial craters, binned by minimum diameter (cumulative plot). Blue error bars bracket the 90% confidence intervals on lower limit (by Poisson estimation). Full Monte Carlo fit corresponds to the gray band. Black zone is excluded with >95% confidence. White zone is excluded in <5% of trials. Black asterisks correspond to the median outcome of the Monte Carlo procedure. Red dashed lines assume a constant aggradation rate and that burial by 10% of a crater's diameter is sufficient to obscure the crater from orbiter-image surveys. (b) Corresponding aggradation rate estimates for alluvial fans, binned by minimum diameter (cumulative plot). Error bars are the same as in (a). Results are plotted excluding craters for which synfluvial versus prefluvial status could not be determined, but including craters of quality score 3. These choices have little effect on our conclusions. That is because the quality-3 embedded craters, and the craters whose synfluvial status is uncertain, correspond to small diameter bins for which our counts are probably incomplete. By fitting a single time-span (and separately fitting a single erosion rate) to the fan deposits, we implicitly assume that the probability of finding an embedded crater is spatially uniform. However, this assumption is inconsistent with the clumpiness of the observed spatial distribution of embedded craters. For example, Crater "W" (Kraal et al. 2008) has 20% of the embedded craters (5 out of 25) even though the fan area at that site is only 500 km2 (3% of total). This cluster is unlikely to be due to chance (assuming cratering is a Poisson process): if fan area is divided into (17000 km2)/(500 km2) = 34 equal-area sites, the expected number of craters per site is λ=25/34=0.73. The probability of finding ≥5 embedded craters in ≥1 sites is then only 1 -- (1 -- Σx=5 ∞f( x0.73))34 =3%, where f is the Poisson probability distribution function. To be consistent with data at the 50% level, we must increase λ by 200%. This might correspond to uniform fan age with spatially non- uniform detectability of embedded craters. Alternatively Crater "W" might record anomalously slow aggradation. In either case, our best estimate of the time span of fan- forming climates is 200% longer than in our lower limit. Spatial staggering of fan aggradation would increase our lower limit. Time-varying orbital forcing would favor snowmelt (e.g. Kite et al. 2013b) in different places at different times. Localized precipitation would not be globally correlated (Kite et al. 2011). If the fan deposits underwent erosion during the period of net aggradation, then this would destroy some embedded craters. Therefore, if erosion occurred during fan aggradation, our upper limit would increase further. Mars crater fluxes are extrapolated from Lunar radiogenic ages and corresponding crater densities (Neukum et al. 2001). Those crater densities have been argued to be incorrect 9 (Robbins 2014). Adopting Robbins' chronology would reduce flux uncertainty from (1-12)× modern (Neukum chronology function) to (1.5 -- 2.2)× modern. Because high aggradation rates in our Monte Carlo runs always correspond to high crater fluxes, including Robbins' chronology would raise our lower limit. D<100m embedded craters place an upper limit on paleoatmospheric pressure (Kite et al. 2014). Since D<50m impact craters are extremely rare on Earth (atmospheric column density 104 kg/m2), the existence of D<100m embedded craters in the Mars fans suggests atmospheric column density <2×104 kg/m2, i.e. P<1 bar around the time of fan aggradation (Fig. 5). Previous analyses of the interval over which alluvial fans formed have divided fan volume by the inferred fluvial sediment transport flux (e.g. Jerolmack et al. 2004). This duration is a lower bound on the interval over which alluvial fans formed, because not all years need produce runoff. Our lower bound exceeds sedimentological lower bounds by >1000-fold. Many alluvial fans are ~1km thick. Suppose a fan:alcove area ratio of 0.5. Typical water:sediment ratios on Earth are 103:1. 2000km of water at 0.5m snowmelt/year gives 4 Myr. These calculations are highly uncertain. For example, if the amount of snowmelt is limited by snow supply to 10 cm/yr, then the fan formation time is 20 Myr, equal to our strict lower limit on the total time span of alluvial fan formation. However, given quasi-periodic orbital variability, fine-tuning of Mars' hydrological cycle is required to produce small amounts of runoff every year, especially for our preferred lower limit of >(100-300) Myr. Slow net aggradation rates in areas of steep relief (Fig. 1) suggest intermittency. Intermittency in alluvial-fan-forming climate is further suggested by combining our data with other constraints. The paucity of mineralogic evidence for in-situ alteration of fan deposits (McKeown et al. 2013), the presence of hydrated silica (possibly opal; Carter et al. 2012), and the persistence of olivine (Stopar et al. 2006), when combined with the >20 Myr span of surface liquid water required by our data, suggest that climate conditions were cold and that intermittency further reduced liquid water interaction with soil. Cold conditions are also suggested by sedimentary-deposit mineralogy at Gale (McLennan et al. 2014, Siebach et al. 2017). Intermittency is also suggested by multiple pulses of fan formation at Holden (Irwin et al. 2008), Gale, and Melas Chasma (Williams & Weitz 2014). In summary, the data exclude any explanation that produces a single burst of habitability of <20 Myr duration. For example, the data exclude triggering by the thermal pulse caused by the impacts that formed the large craters which host the alluvial fans. The data disfavor fluvial sediment transport every year for >20 Myr. Among other possibilities, the data permit chaos rainfall); a a long-lived habitable environment trigger (Baker et al. 1991); or obliquity-paced fluvial intermittency (Kite et al. 2013b). (snowmelt or 10 4.2. Implications for paleohydrology and climate. 4.3. Implications for habitability. Figure 5. Geomorphic history of Mars (Howard 2007, Fassett & Head 2011). FSV = Fresh Shallow Valleys (Wilson et al. 2016). RSL = Recurring Slope Lineae. Pre-valley network fluvial sediment transport is from Irwin et al. (2013). The time span of alluvial-fan aggradation (§3) is a proxy for the time span of spatially associated paleolakes. Those paleolakes include candidate playa deposits at the toes of fans (e.g. Morgan et al. 2014), a >100m deep paleolake in Crater "P" (Kraal et al. 2008) suggested by a common fan-frontal-scarp elevation of -2700m (in our CTX DTMs), and the Eberswalde paleolake (which shares a drainage divide with fans at Holden; Irwin et al. 2015). In addition, rivers and lakes occurred during the Late Hesperian-Amazonian in Valles Marineris (e.g. Mangold et al. 2004) and Arabia Terra (Wilson et al. 2016). Lake deposits have good biosignature recovery potential (Summons et al. 2011), and biosignature recovery from Proterozoic lake deposits is routine (Peters et al. 2005). However, biosignatures would have been destroyed if lake waters were oxidizing. Explaining young alluvial fans on Mars is a challenge to climate models. To determine the time span of alluvial fan forming conditions, we counted embedded craters. We found a high density of embedded craters, which requires that the river-permitting climate(s) spanned >20 Myr. If aggradation was steady at <(4-8) μm/yr, which is consistent with our data, then fan build-up required >(100-300) Myr (Table 1). The data make the challenge of explaining the alluvial fans more severe, because they exclude a single short-lived anomaly as the cause of the alluvial fans. 5. Conclusions. 11 Acknowledgements. Supplementary Information. The Supplementary Information is too large to be included in the arXiv version of our article, but may be freely downloaded from the publisher's website via doi:10.1002/2017GL072660 , or from the website of the lead author. This supporting section provides an overview of the prior probabilities used in the Monte Carlo estimation as Text S1. Figure S1 provides a global map for each interbedded crater used in this study. Figure S2 provides images, location information, and ratings for each interbedded crater considered during the CTX-based search. Figure S3 provides comparable information regarding the supplementary HiRISE-based interbedded crater search. We thank Alan Howard, Marisa Palucis, Becky Williams, Ross Irwin, Jean-Pierre Williams, Bill Dietrich, and Brian Hynek. We thank two anonymous reviewers for their useful recommendations, and Andrew Dombard for editorial handling. All data may be obtained by contacting the lead author ([email protected]). We are grateful for financial support from the US taxpayer via NASA grants NNX16AG55G and NNX15AM49G. References. Armitage, John J., Warner, Nicholas H., Goddard, Kate, Gupta, Sanjeev, 2011. Timescales of alluvial fan development by precipitation on Mars. Geophys. Res. Lett. 38 (17). CiteID L17203. Baker, V.R., Strom, R.G., Gulick, V.C., Kargel, J.S., Komatsu, G. 1991. Ancient oceans, ice sheets and the hydrological cycle on Mars. Nature 352, 589-594. Carter, J., et al., 2012. Composition of Deltas and Alluvial Fans on Mars. In: 43rd Lunar and Planetary Science Conference. March 19 -- 23, 2012, Woodlands, TX. LPI Contribution No. 1659, id.1978. Dietrich, W.E., Palucis, M.C., Williams, R.M.E., Lewis, K.W., Rivera-Hernandez, F., and Sumner, D.Y. (2017), Fluvial gravels on Mars: Analysis and implications, p. 755 - 784 in D. Tsutsumi & J. Laronne, Gravel-bed rivers: Processes and disasters, John Wiley and Sons, Oxford. Dundas, C. M., Keszthelyi, L. P., Bray, V. J. & McEwen, A. S. Role of material properties in the cratering record of young platy-ridged lava on Mars. Geophys. Res. Lett. 37, L12203 (2010). Fassett, C. I., and J. W. Head, 2011, Sequence and timing of conditions on early Mars. Icarus , 211, 1204-1214. Grant, J.A., Wilson, S.A., 2011. Late alluvial fan formation in southern Margaritifer Terra, Mars. Geophys. Res. Lett. 38 (8). CiteID L08201. Grant, J.A., Wilson, S.A., 2012, A possible synoptic source of water for alluvial fan formation in southern Margaritifer Terra, Mars, Planetary & Space Science Grindrod, P.M.; Heap, M.J.; Fortes, A.D.; Meredith, P.G.; Wood, I.G.; Trippetta, F.; Sammonds, P.R. 2010, Experimental investigation of the mechanical properties of synthetic magnesium sulfate hydrates: Implications for the strength of hydrated deposits on Mars, J. Geophys. Res. 115, CiteID E06012. Hartmann, W. K., 1974, Geological observations of Martian arroyos., J. Geophys. Res. 79, 3951 -- 3957. Hartmann, W.K., 2005. Martian cratering 8: isochron refinement and the chronology of Mars. Icarus 174, 294 -- 320. 12 Hauber, E.; Platz, T.; Reiss, D.; Le Deit, L.; Kleinhans, M. G.; Marra, W. A.; Haas, T.; Carbonneau, P., 2013, Asynchronous formation of Hesperian and Amazonian-aged deltas on Mars and implications for climate, J. Geophys. Res.: Planets, 118, 1529-1544. Howard, A.D., 2007. Simulating the development of Martian highland landscapes through the interaction of impact cratering, fluvial erosion, and variable hydrologic forcing. Geomorphology 91, 332 -- 363. Irwin, R. P.; Grant, J. A.; Howard, A. D., 2008, The Alluvial Fan Complex in Holden Crater: Implications for the Environment of Early Mars, 39th Lunar and Planetary Science Conference, abstract id. 1869 Irwin, R.P., Lewis, K.W., Howard, A.D., Grant, J.A., 2015. Paleohydrology of Eberswalde crater, Mars. Geomorphology. http://dx.doi.org/10.1016/j.geomorph.2014.10.012. Irwin, Rossman P.; Tanaka, Kenneth L.; Robbins, Stuart J., 2013, Distribution of Early, Middle, and Late Noachian cratered surfaces in the Martian highlands: Implications for resurfacing events and processes, Journal of Geophysical Research: Planets, 118, 278-291. Jerolmack, Douglas J.; Mohrig, David; Zuber, Maria T.; Byrne, Shane, 2004, A minimum time for the formation of Holden Northeast fan, Mars, Geophysical Research Letters, 31, CiteID L21701. Jerolmack, D.J., Sadler, P., 2007. Transience and persistence in the depositional record of continental margins. J. Geophys. Res. (Earth Surface) 112, F03S13. Johnson, B.C.; Collins, G.S.; Minton, D.A.; Bowling, T.J.; Simonson, B.M.; et al. 2016, Spherule layers, crater scaling laws, and the population of ancient terrestrial impactors, Icarus, 271, 350-359. Kite, E.S., Michaels, T.I., Rafkin, S.C.R., Manga, M., & W.E. Dietrich, 2011. Localized precipitation and runoff on Mars, J. Geophys. Res. Planets, 116, E07002. Kite, E.S., Lucas, A., Fassett, C.I., 2013a, Pacing early Mars river activity: embedded craters in the Aeolis Dorsa region imply river activity spanned >∼(1-20) Myr. Icarus 225, 850 -- 855. Kite, E. S., I. Halevy, M. A. Kahre, M. Manga, and M. Wolff, 2013b, Seasonal melting and the formation of sedimentary rocks on Mars, Icarus, 223, 181 -- 210. Kite, E.S., Williams, J.-P., Lucas, A., Aharonson, O., 2014. Low palaeopressure of the martian atmosphere estimated from the size distribution of ancient craters. Nat. Geosci. 7, 335 -- 339. Kite, E.S., Howard, A., Lucas, A.S., Armstrong, J.C., Aharonson, O., Lamb, M.P., 2015. Stratigraphy of Aeolis Dorsa, Mars: stratigaphic context of the great river deposits. Icarus 253, 223-242. Kite, E.S.; Mayer, D.P., 2017, Mars sedimentary rock erosion rates constrained using crater counts, with applications to organic matter preservation and to the global dust cycle, Icarus, 286, 212-222. Kleinhans, M.G., van de Kasteele, H.E., Hauber, E., 2010. Palaeoflow reconstruction from fan delta morphology on Mars. Earth Planet. Sci. Lett. 294, 378 -- 392. Kraal, E.R., et al., 2008, Catalogue of large alluvial fans in martian impact craters, Icarus, 194, 101- 110. Mangold, N., et al., 2004. Evidence for precipitation on Mars from dendritic valleys in the Valles Marineris Area. Science 305, 78 -- 81. 13 Mangold, N.; Kite, E. S.; Kleinhans, M. G.; Newsom, H.; Ansan, V.; Hauber, E.; Kraal, E.; Quantin, C.; Tanaka, K., 2012a, The origin and timing of fluvial activity at Eberswalde crater, Mars, Icarus, 220, 530-551. Mangold, N.; Carter, J.; Poulet, F.; Dehouck, E.; Ansan, V.; Loizeau, D., 2012b, Late Hesperian aqueous alteration at Majuro crater, Mars, Planetary and Space Science, 72, 18-30. McKeown, N. K.; Rice, M. S.; Warner, N. H.; Gupta, S., 2013, A Detrital Source for the Phyllosilicates at Eberswalde Crater, 44th Lunar and Planetary Science Conference, held March 18-22, 2013 in The Woodlands, Texas. LPI Contribution No. 1719, p.2302. McLennan, S.M., et al., 2014. Elemental geochemistry of sedimentary rocks at Yellowknife Bay, Gale Crater, Mars. Science 343 (6169). http://dx.doi.org/10.1126/science.1244734 Melosh, H.J., 1989. Impact cratering: a geologic process. In: Research supported by NASA, 253. Oxford University Press, New York, p. 11. (Oxford Monographs on Geology and Geophysics, No.11), 1989. Michael, G. G.; Kneissl, T.; Neesemann, A., 2016, Planetary surface dating from crater size-frequency distribution measurements: Poisson timing analysis, Icarus, 277, 279-285. Moore, J.M., & Howard, A.D., 2005, Large alluvial fans on Mars, J. Geophys. Res., 110, E4, E04005. Michael, G.G., 2013. Planetary surface dating from crater size-frequency distribution measurements: multiple resurfacing episodes and differential isochron fitting. Icarus, 226, 885 -- 890. Morgan, A.M., et al., 2014. Sedimentology and climatic environment of alluvial fans in the martian Saheki crater and a comparison with terrestrial fans in the Atacama Desert. Icarus, 229, 131 -- 156. Neukum, G.; Ivanov, B. A.; Hartmann, W. K., 2001, Cratering Records in the Inner Solar System in Relation to the Lunar Reference System, Space Science Reviews, 96, 55-86. Okubo, C.H., 2007, Strength and deformability of light-toned layered deposits observed by MER Opportunity: Eagle to Erebus craters, Mars. Geophysical Research Letters 34, CiteID L20205. Palucis, M., et al. 2014, The origin and evolution of the Peace Vallis fan system that drains to the Curiosity landing area, Gale Crater, Mars, J. Geophys. Res. Planets., 119, 705-728 Peters, K.E., et al. (eds.), 2005, The biomarker guide: volume 2, Cambridge University Press, Cambridge, UK. Robbins, Stuart J., 2014, New crater calibrations for the lunar crater-age chronology, Earth and Planetary Science Letters, 403, 188-198. Siebach, K. L., M. B. Baker, J. P. Grotzinger, S. M. McLennan, R. Gellert, L. M. Thompson, J. A. Hurowitz, 2017, Sorting out Compositional Trends in Sedimentary Rocks of the Bradbury Group (Aeolus Palus), Gale Crater, Mars, J. Geophys. Res. Planets, DOI: 10.1002/2016JE005195 Stopar, J. D., G. J. Taylor, V. E. Hamilton, L. Browning (2006) Kinetic model of olivine dissolution and extent of aqueous alteration on Mars, GCA, 70, p. 6136-6152 Summons, R.E., et al., 2011. Preservation of Martian organic and environmental records: final report of the Mars Biosignature Working Group. Astrobiology 11, 157 -- 181 Tewksbury, B. J.; Hogan, J. P.; Kattenhorn, S. A.; Mehrtens, C. J.; Tarabees, E. A., 2014, Polygonal faults in chalk: Insights from extensive exposures of the Khoman Formation, Western Desert, Egypt Geology, 42(6), 479-482. 14 Warner, N.H.; Gupta, S.; Calef, F.; Grindrod, P.; Boll, N.; Goddard, K., 2015, Minimum effective area for high resolution crater counting of martian terrains, Icarus, 245, 198-240. Watters, W.A., Geiger, L.M., Fendrock, M., Gibson, R., 2015. Morphometry of small recent impact craters on Mars: size and terrain dependence, short-term modification. J. Geophys. Res. (Planets) 120, 226 -- 254. Williams, Rebecca M. E.; Malin, Michael C., 2008, Sub-kilometer fans in Mojave Crater, Mars, Icarus, 198, 365-383. Williams, R.M.E., Deanne Rogers, A., Chojnacki, M., Boyce, J., Seelos, K.D., Hardgrove, C., Chuang, F., 2011. Evidence for episodic alluvial fan formation in far western Terra Tyrrhena, Mars, Icarus 211 (1), 222 -- 237. Williams, R. M. E. et al., 2013, Martian Fluvial Conglomerates at Gale Crater, Science, 340, 1068-1072. Williams, R.M.E., Weitz, C.M., 2014. Reconstructing the aqueous history within the southwestern Melas Basin, Mars: clues from stratigraphic and morphometric analyses of fans. Icarus 242, 19 -- 37. Wilson, S. A.; Grant, J. A.; Howard, A. D., 2012, Distribution of Intracrater Alluvial Fans and Deltaic Deposits in the Southern Highlands of Mars, 43rd Lunar & Planet. Sci. Conf., abstract id.2462. Wilson, S.A.; Howard, A.D.; Moore, J.M.; Grant, J.A., 2016, A cold-wet middle-latitude environment on Mars during the Hesperian-Amazonian transition: Evidence from northern Arabia valleys and paleolakes, J. Geophys. Res.: Planets, 121, 1667-1694. 15
1603.04580
1
1603
2016-03-15T07:33:23
A sound nebula: the origin of the Solar System in the field of a standing sound wave
[ "astro-ph.EP" ]
According to the planetary origin conceptual model proposed in this paper, the protosun centre of the pre-solar nebula exploded, resulting in a shock wave that passed through it and then returned to the centre, generating a new explosion and shock wave. Recurrent explosions in the nebula resulted in a spherical standing sound wave, whose antinodes concentrated dust into rotating rings that transformed into planets. The extremely small angular momentum of the Sun and the tilt of its equatorial plane were caused by the asymmetry of the first, most powerful explosion. Differences between inner and outer planets are explained by the migration of solid matter, while the Oort cloud is explained by the division of the pre-solar nebula into a spherical internal nebula and an expanding spherical shell of gas. The proposed conceptual model can also explain the origin and evolution of exoplanetary systems and may be of use in searching for new planets.
astro-ph.EP
astro-ph
A sound nebula: the origin of the Solar System in the field of a standing sound wave S. Beck1, V. Beck1,* 1Alfa-Carbon R(cid:13) Beck, Berlin, Germany *Correspondence: [email protected] Abstract According to the planetary origin conceptual model proposed in this paper, the protosun centre of the pre-solar nebula exploded, resulting in a shock wave that passed through it and then returned to the centre, generating a new explosion and shock wave. Recurrent explosions in the nebula resulted in a spherical standing sound wave, whose antinodes concentrated dust into rotating rings that transformed into planets. The extremely small angular momentum of the Sun and the tilt of its equatorial plane were caused by the asymmetry of the first, most powerful explosion. Differences between inner and outer planets are explained by the migration of solid matter, while the Oort cloud is explained by the division of the pre-solar nebula into a spherical internal nebula and an expanding spherical shell of gas. The proposed conceptual model can also explain the origin and evolution of exoplanetary systems and may be of use in searching for new planets. 1 Introduction The classical theory of the origin of the Solar System is based on the Kant -- Laplace nebular hypothesis [1, 2] suggesting that the Sun and the planets condensed out of a spinning nebula of gas and dust. During the condensation process, the spin of the nebula accelerated (the 'pirouette' effect), and the resulting distribution of angular momentum caused it to form a disc. The center of the nebula evolved into a hot, highly compressed gas region -- the protosun. Concentration and coalescence of dust particles in the remainder of the spinning disc led to the formation of planets orbiting the Sun. This theory accounts for the general nature of the origin of the Solar System, but it cannot explain many of the observed facts. One problem is the angular momentum distribution: the Sun has more than 99.8% of the entire system mass, but only about 0.5% of the total angular momentum, with the remaining 99.5% residing in the orbiting planets. The hypothesis also fails to explain the 7◦ tilt of the Sun's equatorial plane relative to the average orbital plane of the planets. Another serious problem is the distinction between small solid-surface inner planets and outer gas giants. Moreover, the observed regularity in planetary distances from the Sun -- the so-called Titius-Bode law -- has no explanation. It is difficult to explain the existence of the Oort cloud beyond the Solar System's planets. 1 It consists of trillions of small objects composed of dust and water, ammonia and methane ice and it is believed that these objects were scattered outwards by the gas giants at the planetary formation stage and then acquired distant circular orbits (out to about one light year) as a result of gravitational forces due to nearby stars. Such an Oort cloud emergence scenario seems very unlikely for such a large number of bodies. Neptune is the most distant gas planet. Based on the decreasing series of giant planet masses Jupiter: 318 earth masses (M⊕); Saturn: 95.3 M⊕; Uranus: 14.5 M⊕ we could expect Neptune's mass to be several times smaller than that of Uranus. Such a mass distribution can be explained by decreasing density of the gas nebula from the centre to the periphery, so that each planet has less gaseous substance than the previous one. In reality, Neptune has a mass of 17.5 M⊕, which is greater than the mass of Uranus. Authors, such as for example, Chamberlin [3], Moulton [4] , Schmidt [5], von Weizscker [6], McCrea [7], Woolfson [8] and Safronov [9] have offered a variety of scenarios for the Solar System's formation. However, none of the existing theories is able to give a comprehensive picture of the emergence and development of the planetary system. A large number of recently discovered exoplanets (1955 confirmed planets by February 2016, NASA Exoplanet Archive1) allows us to estimate the number of planets in our galaxy as many billions, so that a compre- hensive understanding of the processes leading to the emergence of planetary systems becomes critically important. In this article, we present a new conceptual model of the Solar System's formation from a pre-solar gas and dust nebula in the field of a standing sound wave. 2 Theoretical construction 2.1 A spherical standing sound wave It is generally accepted that during a sufficiently strong process of heating, gravitational contraction in the centre of the pre-solar nebula started thermonuclear fusion of hydrogen into helium. A large amount of energy was emitted and radiation pressure prevented the further contraction of the gas nebula. However, thermonuclear fusion could not start quietly; the process resulted in an explosion that caused a spherical shock wave originating from the central region. The specific explosion mechanism accompanied by the thermonuclear fusion processes is beyond the scope of this article; we note here only that the power of the explosion would have been large enough for the shock wave to spread all over the pre-solar nebula2. A long time observed active processes, such as jets and outflows associated with star formation, e.g. [10]-[18], suggests that the explosions fundamentally inherent in the birth of stars and can start the emergence of planetary systems as described in our model. As the wave propagated, the gas particles in the pre-solar nebula oscillated radially, for two reasons: first, the difference in gas pressure; and secondly, the gravitational pull toward the centre of the nebula. The second factor began to play a crucial role at large distances from the central attracting mass concentrations: at some point, the accelerated gas particles at the 1http://exoplanetarchive.ipac.caltech.edu 2Appendix A outlines some arguments that allow us to make some assumptions about the evolution of protostars. 2 wave front would no longer return under the influence of gravity. Gas density at this distance from the centre would have been so low that the pressure difference could no longer cause the return movement of gas particles. The peripheral part separated, and the pre-solar nebula was thus divided into a central spherical region and an expanding spherical shell of gas (Figure 1). Figure 1: Separation of the pre-solar nebula into (1) a central spherical gas nebula and (2) an expanding spherical shell. Gas particles at the boundary of the central spherical nebula fell under gravity towards the centre then stopped when the pressure of the lower gas layers exceeded the gravitational pull, and began to move in the opposite direction. As a result of the interaction between gravity and contracting gas back-pressure, the boundary of the spherical nebula began to oscillate radially and a reverse spherical wave propagated from the periphery towards the centre. The sound wave would have appeared to reflect from the gravitational pause -- a spherical shell where the gas particles velocities were not high enough to overcome the gravitational pull from the centre of the nebula and the gas particles that had moved outwards returned due to oscillations; that is, their velocities were lower than the escape velocity: (cid:115) νe = 2GM r , (1) where G is the gravitational constant, M is the mass of the attraction centre and r is the distance of the gravitational pause from the centre. The centre of the pre-solar nebula became quiescent after the first massive explosion: the central part expanded, since the gravitational field was not yet intense enough to resist the much increased gas pressure, and the thermonuclear fusion of hydrogen combustion diminished. A large mass of gas was ejected into the surrounding space while the compact region of compressed and heated gas -- the protosun -- remained at the nebula's core. Several hundred years after this, the wave returning from the boundary of the spherical nebula reached the protosun, 3 concentrated in the centre, and then began to propagate towards the periphery again: a rapid pressure increase resulted in a dramatic rise in temperature at the centre of the protosun, generating a new hydrogen explosion, much weaker than the first one but still strong enough to give extra energy to the wave reflected from the central region. The reflected wave travelled all the way from the centre to the periphery of the spherical nebula, was reflected from the gravitational pause and then returned to the centre again, causing another explosion. This process, repeated several times, eventually established regular oscillations: the wave propagated from the centre to the edge, was reflected from the boundary of the spherical nebula, and returned, causing another explosion that compensated for wave energy loss. Thus, the wave caused explosions while acquiring the energy it needed, establishing a self-sustaining process whose period, defined by the free oscillations of the spherical nebula boundaries, equalled many tens of years. The acoustic radiation pressure prevented gravitational contraction of the pre-solar nebula and compensated for deviations from the spherical shape3 that might result, for example, from the gas turbulence. Waves travelling from the centre and from the periphery interfered with each other to form a giant spherical acoustic cavity resonator the size of the modern Solar System, including the Kuiper belt and the scattered-disc (Figure 2). This resonator contained a standing sound wave with nodes and antinodes, the number of which would have been at least 10 for the pre-solar system: 8 planets + the asteroid belt + the scattered-disc region. Figure 2: The standing wave in the pre-solar nebula: (1) nodes, (2) antinodes (gas compression and expansion), (3) protosun, (4) real-scale wave. Nodes and antinodes are evenly spaced for illustration purposes; in reality, separations would increase dramatically with increasing distance from the centre. This article does not include a calculation of exact distances between the antinodes of the 3Here and below, for simplicity, we are talking about a spherical shape. Needless to say, a rotating nebula takes the form of an ellipsoid of revolution; it can be shown that all the derivations for spherical sound waves retain also for the rotating nebula in the form of an ellipsoid. 4 standing wave; we may simply state that the wavelength increases with increasing distance from the centre of attraction. Section [3.4] outlines some regularities that allow us to make some observations about the physical reasons underlying planetary positions in the Solar System. 2.2 Dust concentration in the antinodes In our model, the pre-solar system can keep 'sounding' for many millions of years, as the periodic central explosions significantly retard the gravitational contraction of the protosun and the acoustic radiation pressure stabilizes the gas nebula, preventing its collapse. The dust present in the pre-solar nebula gradually concentrates in the antinodes of the acoustic oscillating system. Apart from gas viscosity, the process of solid particle concentration in the standing wave also relies on attraction from the large gas masses that periodically emerge in the antinodes of the standing wave. This attractive force makes the dust particles move towards the gas clusters in the antinodes and collide with each other, causing redistribution of their velocity vectors in such a way that the dust particles come to rest in the centre of the antinode (Figure 3). Figure 3: Dust concentration in the antinodes of the standing wave during four successive phases of compression and expansion. Dark areas stand for gas compression in the antinodes, light areas for expansion. During the gas expansion phase, dust particles in each antinode are gravitationally attracted towards two neighbouring antinodes. The two gravitational sources in these antinodes largely cancel each other out and, moreover, are remote, so that the dust particles in the gas expansion phase mostly remain in the same place. The next phase of compression attracts new dust particles to the antinode, which also come to rest due to collisions with each other and to the viscosity of the compressed gas. 2.3 Migration of solid matter and formation of rings A certain period of time after the standing wave was established, most of the dust would have been concentrated in the regions of spherical antinodes. Due to increased dust concentration, the number of particle collisions significantly increased, causing their coalescence and increase 5 in size and mass. The gravitational field within the antinodes and the centrifugal force together caused dust to concentrate and form equatorial rings, with radii corresponding to the antinodes of the standing sound wave (Figure 4). Figure 4: Migration of solid matter in the spherical nebula. Fg represents the gravitational pull toward the centre and Fc represents the centrifugal force. (1) spherical dust clusters in the antinodes, (2) dust clusters in the equatorial rings, (3) protosun, (4) nebula rotational axis, (5) direction of dust migration in the antinodes, (6) direction of massive dust cluster migration across the antinodes under the influence of the resultant force F. The model predicts that the increase in mass of the solid particles clusters also makes it increasingly difficult for them to stay at the antinodes of the standing sound wave as at the same time that they are affected by attraction from the central gas masses, principally the protosun. The largest clusters of dust fall toward the centre; this results in migration of solid matter from the outer regions of the pre-solar nebula towards the internal zones, while the centrifugal force does not allow the clusters to fall directly onto the protosun. The gravitational attraction of the gas masses in the antinodes directs the matter falling towards the spinning rings forming in the equatorial plane, where the solid matter comes to rest. A significant amount of solid matter collected in the region of the orbits of Venus and Earth, as it was concentrated here from most of the pre-solar nebula volume. This dust, gathered in compact rings, is thousands of times more concentrated here than in the primary gas and dust nebula, greatly accelerating the coalescence of particles and leading to the emergence of increasingly massive dust clusters. Agglomerated clusters of dust in narrow spinning rings move in almost identical circular orbits around the protosun and have very small relative velocities. Since their collisions do not result in fragmentation, they form massive planetesimals relatively quickly, which in turn agglomerate to form planets and their satellites. In this model, are no catastrophic collisions 6 between the nascent planets, their satellites or other massive objects; the planets keep the circular orbits and planes of the equatorial dust rings. 2.4 Birth of the Sun Gravitational contraction of the protosun, which had been significantly slowed as a result of regular explosions in the centre, still continued, and millions of years after the establishment of the standing sound wave gas temperature at the centre of the protosun became sufficiently high for thermonuclear fusion of hydrogen into helium to continue between explosions, resulting in the birth of a new star -- the Sun. The newborn Sun stopped augmenting the sound wave by periodic explosions, and the standing wave diminished. The gas shell previously supported by acoustic pressure began to shrink, forming gas-giant planets around already existing solid nuclei. Jupiter acquired the greatest share of mass, as it was located in the region of highest gas density. All the other gas planets were situated in a lower density environment and thus obtained smaller masses. Jupiter also received large amounts of gas from the inner-planet region, blown outwards by the strong solar wind in the first few millions of years after the birth of the Sun. 2.5 The Oort cloud The expanding shell that separated from the spherical part of the nebula during the passage of the wave front from the first explosion (see Figure 1) moved faster than the escape velocity νe and could not return to the centre, as the pull of gravity from the central masses was too weak at such distance. With its expansion, this shell accumulated increasing quantities of highly rarefied gas from the primary nebula at its edge, while its expansion rate gradually slowed and eventually stopped at a distance of about 1 light year from the centre. This formed a giant spherical region containing dust, ice and frozen gases particles in addition to gaseous hydrogen and helium. Over time, the gas component of the shell dissipated in space, while solid particles were concentrated in increasingly large chunks of ice and dust -- the cometary bodies -- to form the Oort cloud [19] which has a weak gravitational connection with the central part of the system. 2.6 Neptune, the Kuiper belt and scattered-disc objects After expansion of the gas shell ceased in the region of the Oort cloud, a weak reverse wave formed within the shell and moved towards the central spherical nebula of gas, where planetary nuclei had already emerged. Hundreds of thousands of years later the reverse wave from the Oort cloud collided with the outer boundary of the central spherical nebula, causing a redistribution of matter on the edge of the Solar System. Large gas and ice masses from the outer antinodes of the standing sound wave were shifted, and the planet Neptune formed a little closer to the Sun relative the original position of the 9th antinode, while its mass increased several times by capture of gas and ice from the reverse wave from the Oort cloud and from the outer antinodes. Scattered-disc objects such as the minor planet Eris acquired highly elongated orbits, as there was a long period under the influence of gravitational pull 7 from the gas masses transported by the reverse wave from the Oort cloud. Minor planets in the Kuiper belt also gained significant eccentricity. The orbits of the low-mass objects in the scattered disc region and in the Kuiper belt have been changed so dramatically, that they had become to reach Neptune and Uranus, and under the influence of their gravity even began to go into the central region of the solar system and possibly caused the late heavy bombardment of the inner planets, in which was also attended cometary bodies from the Oort cloud, that have had unstable orbits in the first time after the formation. The periphery processes were very slow, developing over many millions of years, and were relatively weak in their effect on the central region of the pre-solar system. The reverse wave from the Oort cloud only caused the formation of the Kuiper belt, an offset of Neptune's formation and its mass increase. The rest of the Solar System was and is still affected by the region of the Oort cloud only through comets. 3 Discussion 3.1 Terrestrial and giant planets The difference in chemical composition of the inner and outer planets is explained in our model by the migration of solid matter from the outer spherical dust shells to the internal ones (see Figure 4). Before concentrating in stable spinning dust rings, the heavy chemical elements, such as iron and silicate travelled farther towards the centre of the system and were incorporated by the inner planets, with Earth and Venus having more matter, and Mercury and Mars having less. The composition of the closest to the Sun Mercury at the same time has obtained a significant amount of iron. The region between Mars and Jupiter had insufficient solid matter left for a proper planet, so only the minor planets of the asteroid belt were formed. Lighter and more volatile chemical compounds, such as water, methane or ammonia remained in significant quantities in the colder regions beyond the asteroid belt, and became the compositional basis of the giant planets. 3.2 Angular momentum distribution and dust structure If we assume that the first powerful explosion in the centre of the protosun was not absolutely symmetrical, slight asymmetry in the explosion led to a redistribution of the angular momentum in the pre-solar nebula, with the less massive peripheral zone beginning to rotate faster. In addition, an initial asymmetry could produce the observed 7◦ equatorial plane tilt relative to the Sun's plane of rotation. Currently the Sun has more than 99.8% of the entire Solar System mass but only about 0.5% of the total angular momentum. If the rotational kinetic energy were evenly distributed over the original pre-solar nebula, the resulting rotation (taking into account the pirouette effect) would be thousands or millions of times slower than the current rotation of the planets. The angular momentum was initially distributed evenly in the pre-solar nebula, and this means that such weak rotation cannot result in the process of dust disc formation as described by the nebular hypothesis. 8 Figure 5: An image of a protoplanetary dust nebula around HL Tauri received by ALMA (The Atacama Large Millimeter/submillimeter Array). Credit: ALMA (ESO/NAOJ/NRAO). However, these discs are visible in images of some young stars (Figure 5), and in some cases their internal structure can be identified. If a spherical dust cluster such as described in this model were observed from one side, it would be impossible to observe any individual spheres as they would overlap in the line of sight. Figure 6: The averaged in the line of sight distribution of dust in the spherical antinodes. Closer to centre the dust has already reached the equatorial plane, on the periphery it need to pass a much greater way, so formed a cup-shaped dust distribution structure around the protostar. The stage at which a significant amount of dust has already accumulated in the equatorial plane would give such a system the appearance of a disc (Figure 6), despite the fact that there are significant gaps between individual dust clusters. Figure 7 shows images from the NASA Hubble Space Telescope, showing the complex three- dimensional structure of the dust around young stars that are similar to the structure, which should be observed according to the proposed model (for more images see also Schneider et al. [20]). On these images we see the light reflected from the inner part the near side of the cup-shaped dust structure. In contrast to the images in visible light the image of HL Tauri on the Figure 5 received in submillimetre radiation can show the ring structures of dust around 9 Figure 7: The images from a NASA Hubble Space Telescope visible-light survey of the architecture of dust systems around young stars. Credit: STScI/Hubble, NASA & ESA. the less than 100000 years young star. 3.3 Exoplanetary systems Generally speaking, the explosion in a protostellar nebula centre according to our model can either accelerate the gas shell rotation around the future star, slow it down or even give it a spin in the opposite direction. The latter case has been observed in some exoplanetary systems, e.g. [21]-[27], and the standard nebular hypothesis cannot explain it, since it implies that the central star and its planets should always rotate and revolve in the same direction, following the rotation of the protostellar nebula. In this model, an asymmetry in the initial explosion in the centre of the protostar can also lead to a very strong tilt of the equatorial plane of the planetary system, with tilts of 45◦ or 10 even 90◦ not impossible, e.g. [28]-[31]. Based on the data from the Kepler mission McQuillan et al. have measured the periods of the rotation of stars that have planets [32] and stars without planets [33]. Figure 8 shows a comparison of the distributions of rotation of these stars. Figure 8: The normalized distribution of rotation of stars with planets (blue) and without planets (orange). The horizontal axis represents the rotational period in days and the vertical axis the number of stars with periods in the range of one day, divided by the total number of stars (762 stars with planets and 34030 stars without planets). The inset shows the smoothed difference of these distributions. One can see that the stars with planets rotate slower; according to our model these stars have slowed down its rotation as a result of an asymmetric explosion, which led to the formation of the planetary system. If the central protostar after the explosion accelerate its rotation and the rotation of the protostellar nebula slows down significantly or stops completely, the formation of planets is impossible and only a lone star would be formed from this protostellar nebula. It can also turn out that the nebula begins to rotate in the opposite direction, and in this case, the planets will have retrograde orbits. In their work McQuillan et al. [32] showed that planets with a short orbital period are available only around slowly rotating stars and noted that the reason for this is unclear. The proposed model explains this fact by saying that closed to their stars planets, e.g. the so- called Hot Jupiters, were formed after a very strong asymmetric explosion, so the separation of the protostellar nebula (see Figure 1) has occurred in the immediate vicinity of the central protostar: the spherical nebula with a standing sound wave began spinning very rapidly and the central protostar at the same time began to rotate slowly. 3.4 Background for a physical and mathematical model of the Solar System Oscillations occur in an environment where gravitational field intensity and gas pressure change substantially within a wavelength. It can be assumed that the wavelength is inversely proportional to the product of average values of gravitational acceleration g, which is inversely proportional to the square of the distance to the centre of attraction, and gas pressure p, which 11 in turn depends directly on g: or, supposing that p ∼ g: λ ∼ 1 g · p λ ∼ r4, (2) (3) where λ is the acoustic wavelength, g and p are the average values of gravitational acceleration and gas pressure as function of r, the distance from the centre. Figure 9: Distribution of planetary distances from the Sun. The vertical axis shows the fourth root of distance in kilometres. Figure 9 shows the distribution of planetary distances from the Sun. The planet Neptune is not shown on the figure for the reasons outlined in Section [2.6], its place being taken by Pluto, the main asteroid belt and the scattered disk are represented by the minor planets Ceres and Eris. The planets originated in the antinodes, so the figure also shows the distribution of the standing wave in the pre-solar gas nebula. We can see that the points on the graph may be approximated by two straight lines, supporting our hypothesis that locations of the standing wave antinodes depend on r4. Closer to the Sun, the constant of proportionality determining the wavelength increases, which may be explained by the fact that the speed of the sound wave λ of the increases with rising gas temperature: c ∼ √ T ; consequently, with the frequency f = c sound wave constant, λ(T ) ∼ √ T , (4) where c is the speed of the sound wave, λ is its wavelength and T is the temperature. The dependence hypothesis λ ∼ r4 provides a background for developing a more complex physical and mathematical model for the Solar Systems origin, as well as for planetary systems around other stars. 12 4 Conclusions According to the proposed model of planetary origin from the gas and dust nebula in the field of a standing sound wave, only single stars or multiple stars with large distances between individual components can have planetary systems. Systems with closely located multiple star components cannot sustain the reverse wave after the explosion in the centre of one of the stars causes further periodic explosions due to the motion of the stars, so that it will fade out, while a gas-dust nebula that is not stabilized by the field of a standing sound wave will relatively swiftly accrete to the central stars without any planet formation. However, we do not claim that all the planetary systems must originate on the proposed model. In some rare cases, the planets can be formed, for example, as a result of a close passage of two or more stars and gas-dust nebulae, so may occur planetary systems around closely located multiple stars, e.g. Orosz et al. [34]. Jensen and Akeson [35] describe the case of a binary young stars with components separated by about 386 AU, each of which is surrounded by a cloud of dust, and the plane of rotation of these clouds are strongly tilted relative to each other. The classical theory cannot explain this tilt, the proposed model explains this because the explosions in the centre of each of protostars occur independently of each other, and the dust clouds get their individual tilts. The classical theory suggests that rotation of the protostellar cloud is essential for the for- mation of planets around the central star. The proposed model does not require such an assumption. If a gas-dust nebula does not rotate, after a first asymmetric explosion the central protostar and the peripheral gas masses will gain equal angular momentum, but opposite in sign. The gas-dust nebula will start spinning and at the same time a standing sound wave will emerge, beginning the process of planet formation. The resulting planetary system will have the central star and planets rotating in opposite directions, planetary orbits will be retrograde. The model predicts that the planetary systems with retrograde orbits can also be formed from a rotating protoplanetary nebula, when after an asymmetric explosion (i)the spherical nebula or (ii)the central protostar starts to rotate in opposite direction. In the first case, the planets will have wide orbits around rapidly rotating stars. In the second case, will be formed planets with short orbital periods, such as Hot Jupiters, around slowly rotating stars. The Titius-Bode relation is not based on any physical laws and is only a mathematical formula to a series of numbers. Despite this, the fact that a simple formula with good accuracy describes the actual distances of the planets from the sun, suggesting, that behind it are standing certain physical regularities. Bovaird and al. [36, 37] have analysed a large number of exoplanet systems, consisting of at least from 4 planets, and found that in many of them distances of the planets from the centre correspond to the generalized Titius-Bode relation. The authors predicted the existence of 141 new planets in these systems. Huang and Bakos [38] performed an extensive search in the Kepler data for 97 of the predicted planets in 56 systems and have confirmed five of predictions. Our model provides a physical explanation of the distances of the planets from the central stars and we are confident that it will be able to more accurately predict the new planet in the already open exoplanetary systems. For this purpose, is necessary to develop a mathematical model of a standing wave in the protoplanetary 13 nebula, which is our future target. A theory of the formation of the planetary systems must explain all or at least most of the known facts, but many of the current models consider only some of the facts, leaving all other ignored. There is, for example, abundant recent literature concerning transport of mass and angular momentum in turbulent, magnetized, differentially rotating disks. Measurements made in November 2014 by Rosetta spacecraft and Philae lander on comet 67P/Churyumov- Gerasimenko shown that the comets nucleus is not magnetised (e.g. Auster et al. [39]. This suggests that the magnetic field was absent at the stage of growth of the cometary nucleus and magnetism do not have played a role in the formation of the Solar System. From measurements of the gravitational field Patzold et al. [40] found that the comet 67P/Churyumov-Gerasimenko is low-density, highly porous (72-74 per cent) and homogeneous dusty body without large internal cavities. Massironi et al. [41] showed that the comet was merged from two kilometer- sized objects collided at very low relative speeds. This implies that the comet was formed from particles, coalesced at very low relative velocities, and never experienced deforming its structure collisions with other bodies. This in turn means that the formation of the solar system did not occur under the conditions of turbulence and chaos, as described in many models. The proposed scenario of the planetary system origin answers many open questions related to the origin and evolution of the Solar System. Several known facts and their explanations within the framework of this model are listed below: 4.1 Planetary distances to the Sun are not random -- there are certain regularities (Titius- Bode law). -- Planets are formed in the antinodes of a giant standing sound wave, emerging after a powerful thermonuclear explosion in the centre of the protosun and repeated passage of forward and backward sound waves through the spherical protoplanetary gas-dust nebula. 4.2 There are internal silicate planets and outer gas giant planets. The hypothesis of a rotating protoplanetary disk cannot explain such distribution, as the rotating disk has the whole mass of dust influenced by the centrifugal force, which prevents migration of the matter. -- Distinction between inner and outer planets is explained by migration of the solid matter from the spherical dust concentration zones in the antinodes of the standing wave. 4.3 The Sun contains 99.8% of the mass, but only 0.5% of the angular momentum of the Solar System. -- Asymmetry of the first explosion in the protosun centre resulted to redistribution of angular momentum: the rotation of the peripheral portion of the gas-dust nebula was significantly accelerated, acquiring an increment of angular momentum from the large gas mass emitted from the protosun during explosion. 4.4 There is a 7◦ tilt of the Sun equatorial plane in relation to the average plane of the planetary orbits. 14 -- The equatorial plane of the planetary system tilted during the first powerful explosion in the centre of the pre-solar nebula because of the small asymmetry of the explosion. 4.5 The Oort cloud is a source of comets visiting the inner Solar System. Its existence is not confirmed by direct observations, but is very likely. -- The expanding shell that separated from the interior part of the pre-solar nebula during the passage of the shock wave from the first explosion concentrates large masses of rarefied gas in front of it and stops at a distance of about 1 light year from the Sun, forming a spherical region where comets are formed the Oort cloud. 4.6 Neptune is closer to the Sun than what is implied by the Titius-Bode distribution. -- Neptune was shifted towards the Sun by a backward wave from the Oort cloud. 4.7 The mass of Neptune (17.5 M⊕) is significantly greater than what could be expected based on the decreasing sequence of giant planet masses: Jupiter (318 M⊕), Saturn (95.3 M⊕) and Uranus (14.5 M⊕). -- Neptune gained a significant (a number of times) mass increase from the backward wave of the Oort cloud. 4.8 The Kuiper Belt contains unexpectedly high quantities of wide binary objects (e.g. Petit et al. [42], Parker et al. [43]). -- The growing clumps of matter in rotating rings of dust had very small relative velocities and formed of them in the Kuiper belt large objects approaching at low speeds form gravitationally bound wide binary systems. In our model, the planets retain their orbits unchanged throughout all time of existence solar system; therefore binaries in the Kuiper belt were not destroyed by any approaching massive bodies. Acknowledgments: This research has made use of the NASA Exoplanet Archive, which is operated by the California Institute of Technology, under contract with the National Aeronau- tics and Space Administration under the Exoplanet Exploration Program. Appendix A Currently there are no detailed models of the evolution of protostars in the period from the beginning of thermonuclear reactions to sustainable hydrogen burning and the formation of convective energy transfer zone. When in the process of gravitational contraction temperature in the center of the protostar reaches 10 million Kelvin, start the fusion reactions of hydrogen into helium. These reactions, however, are very slow with the characteristic time millions of years, so a relatively small amount of released energy cannot stop the gravitational contraction of a protostar and the temperature in the central region continues to rise. With an increase in 15 temperature increases the rate of fusion reactions and the amount of released energy, and at some point the contraction of the protostars centre stops. The central region of the protostar at this time is in a very unstable state; the hydrogen burning begins to occur explosively and from the centre to the periphery begins to spread the shock wave. After the first explosion the central region expands and cools, the rapid hydrogen burning stops, then the compression process is repeated and the next explosion occurs. If a standing sound wave arises in the protostellar nebula, the formation of a planetary system begins, as described in our model. If the standing wave does not appear, the explosions would continue regularly or irregularly until a stable burning of hydrogen be established, at the same time from the surface of the protostar can occur a significant ejection of matter, which in some cases can be observed as a jets, outflows or outbursts, e.g. [10]-[18]. During the passage of the shock wave from the first explosion through the central region of protostar are produced so large pressures and temperature that at the wave front can begin nucleosynthesis of elements heavier than helium, including the elements heavier than iron. Authors, who examined the live in the early solar system short-lived isotopes such as, for example, 26Al, 41Ca (e.g. [44, 45]) or 247Cm [46], presume that they were injected in the solar system by supernova explosions within a very short, not more than a few millions of years, time before the formation of the solar system. This, in turn, requires the existence of some particular conditions for the emergence of the solar system. Our model eliminates the need for any special preconditions and does not put the solar system in a unique position among other planetary systems, because the short-lived isotopes can be produced during the formation of the sun. References [1] [2] [3] [4] [5] [6] [7] Kant, I. Allgemeine Naturgeschichte des Himmels. Johann Friderich Petersen, Knigs- berg und Leipzig, 1755. (English translation: W. Hastie, 1968, in Kants Cosmology, Universal Natural History and Theories of the Heavens. Greenwood Publishing, New York). de Laplace, P.S.M. Expositions du systme du monde (Paris, Imprimerie du Cercle Social ), 1796. Chamberlin, T.C. On a Possible Function of Disruptive Approach in the Formation of Meteorites, Comets, and Nebulae. ApJ 1901, 14, 17-40. Moulton, F.R. On the Evolution of the Solar System. ApJ 1905, 22, 165. Schmidt, O.Y. Meteoritic theory of the origin of the Earth and planets. Dokl. Akad. Nauk SSSR 1944, 45 (6), 245-249. von Weizsacker, C. F. Uber die Entstehung des Planetensystems. Z. Astrophys. 1944, 22, 319. Translation, "On the origin of the planetary system" Report No. RSIC-138 [=AD-4432290] Huntsville AL: Redstone Scientific Information Center. McCrea, W.H. The origin of the solar system. Proc. Roy. Soc. London 1960, Ser. A 256, 245-266. 16 [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] Woolfson, M.M. A capture theory of the origin of the solar system. Proc. Roy. Soc. London 1964, Ser. A 282, 485-507. Safronov, V.S. Evolution of the Protoplanetary Cloud and Formation of the Earth and Planets. Evolution of the protoplanetary cloud and formation of the earth and planets., by Safronov, VS. Translated from Russian. Jerusalem(Israel): Israel Program for Scientific Translations, Keter Publishing House 1972, 212 p. Bally, J.; Devine, D.; Fesen, R.A. & Lane, A.P. Twin Herbig-Haro jets and molecular outows in L1228. ApJ 1995, 454, 345-360. Bally, J.; Reipurth, B. & Davis, C.J. Observations of Jets and Outows from Young Stars. In Protostars and Planets V ; Reipurth, B., Jewitt, D. and Keil, K. (eds.), University of Arizona Press, Tucson, 2007, 215-230. Eisloffel, J. Parsec-scale molecular H2 outflows from young stars. Astronomy and As- trophysics 2000, 354, 236-246. Frank, A. et al. Jets and outflows from star to cloud: Observations confront theory. In Protostars and Planets VI ; Beuther, H., Klessen, R. S., Dullemond, C. P., Henning, University of Arizona Press, Tucson, 2014, 451-474. Busquet, G. et al. The CHESS survey of the L1157-B1 bow-shock: high and low exci- tation water vapor. A&A 2014, 561, A120. Nisini, B. et al. Mapping water in protostellar outflows with Herschel-PACS and HIFI observations of L1448-C. A&A 2013, 549, A16. Santangelo, G. et al. Herschel-PACS observations of shocked gas associated with the jets of L1448 and L1157. A&A 2013, 557, A22. Safron, E.J. et al. HOPS 383: An Outbursting Class 0 Protostar in Orion. ApJ 2015, 800, L5. Stanke, T. Observations of molecular jets in Orion A. Astrophysics and Space Science 2003, 287, 149-160. Oort, J.H. The Structure of the Cloud of Comets Surrounding the Solar System and a Hypothesis Concerning its Origin. Bull. Astron. Inst. Neth. 1950, 11, 91-110. Schneider, G. et al. Probing for Exoplanets Hiding in Dusty Debris Disks: Disk Imag- ing, Characterization, and Exploration with HST/STIS Multi-Roll Coronagraphy. The Astronomical Journal 2014, 148, 59. Narita, N. et al. First evidence of a retrograde orbit of a transiting exoplanet HAT-P-7b. PASJ 2009, 61, L35-L40. [22] Winn, J.N. et al. HAT-P-7: a retrograde or polar orbit, and a third body. ApJ 2009, 703, L99. 17 [23] Winn, J.N. et al. ORBITAL ORIENTATIONS OF EXOPLANETS: HAT-P-4b IS PROGRADE AND HAT-P-14b IS RETROGRADE. The Astronomical Journal 2011, 141, 63. [24] [25] [26] [27] [28] [29] Triaud, A. et al. Spin-orbit angle measurements for six southern transiting planets-New insights into the dynamical origins of hot Jupiters. A&A 2010, 524, A25. Queloz, D. et al. WASP-8b: a retrograde transiting planet in a multiple system. A&A 2010, 517, L1. Bayliss, D. et al. Confirmation of a retrograde orbit for exoplanet WASP-17b. ApJ 2010, 722, L224. H´ebrard, G. et al. The retrograde orbit of the HAT-P-6b exoplanet. A&A 2011, 527, L11. Pont, F. et al. The spin-orbit angle of the transiting hot Jupiter CoRoT-1b. MNRAS 2010, 402, L1-L5. Simpson, E.K. et al. The spin-orbit angles of the transiting exoplanets WASP- 1b,WASP-24b,WASP-38b and HAT-P-8b from Rossiter-McLaughlin observations. MN- RAS 2011, 414, 3023-3035. [30] Winn, J.N. et al. The oblique orbit of the super-Neptune HAT-P-11b. ApJ 2010, 723, L223. [31] [32] [33] [34] [35] [36] [37] [38] Hirano, T. et al. A Possible Tilted Orbit of the Super-Neptune HAT-P-11b. PASJ 2011, 63, 531-536. McQuillan, A., Mazeh, T., & Aigrain, S. Stellar rotation periods of the Kepler objects of interest: A dearth of close-in planets around fast rotators. The Astrophysical Journal Letters 2013, 775(1), L11. McQuillan, A., Mazeh, T., & Aigrain, S. Rotation periods of 34,030 Kepler main- sequence stars: the full autocorrelation sample. The Astrophysical Journal Supplement Series 2014, 211(2), 24. Orosz, J.A., et al. Kepler-47: a transiting circumbinary multiplanet system. Science 2012, 337.6101 1511-1514. Jensen, E.L. & Akeson, R. Misaligned protoplanetary disks in a young binary star system. Nature 2014, 511, 567-569. Bovaird, T. and Lineweaver, C.H. Exoplanet predictions based on the generalized Titius-Bode relation. MNRAS 2013, 435: 11261138. Bovaird, T.; Lineweaver, C.H. and Jacobsen, S.K. Using the inclinations of Kepler systems to prioritize new Titius-Bode-based exoplanet predictions. MNRAS 2015, 448: 36083627 Huang, C.X. & Bakos, G.. Testing the TitiusBode law predictions for Kepler multi- planet systems. MNRAS 2014, 442(1), 674-681. 18 [39] [40] [41] [42] [43] [44] [45] [46] Auster, H-U. et al. The nonmagnetic nucleus of comet 67P/Churyumov-Gerasimenko. Science 2015, 349.6247, aaa5102, doi:10.1126/science.aaa5102. Patzold, M. et al. A homogeneous nucleus for comet 67P/ChuryumovGerasimenko from its gravity field. Nature 2016, 530, 6365, doi:10.1038/nature16535. Massironi, M. et al. Two independent and primitive envelopes of the bilobate nucleus of comet 67P. Nature 2015, 526(7573), 402-405, doi:10.1038/nature15511. Petit, J-M. et al. The extreme Kuiper belt binary 2001 QW322. Science 2008, 322.5900, 432-434. Parker, A.H. and Kavelaars, J.J. Destruction of binary minor planets during Neptune scattering. The Astrophysical Journal Letters 2010, 722.2, L204. Boss, A.P., & Foster, P.N. Injection of short-lived isotopes into the presolar cloud. The Astrophysical Journal Letters 1998, 494(1), L103. Meyer, B.S., & Clayton, D.D. Short-lived radioactivities and the birth of the Sun. In From Dust to Terrestrial Planets. Springer Netherlands, 2000, 133-152. Tissot, F.L., Dauphas, N., & Grossman, L. Origin of uranium isotope varia- tions in early solar nebula condensates. Science Advances 2016, 2(3), e1501400, doi:10.1126/sciadv.1501400. 19
1306.3741
1
1306
2013-06-17T05:24:28
The Dust Tail of Asteroid (3200) Phaethon
[ "astro-ph.EP" ]
We report the discovery of a comet-like tail on asteroid (3200) Phaethon when imaged at optical wavelengths near perihelion. In both 2009 and 2012, the tail appears >=350" (2.5x10^8 m) in length and extends approximately in the projected anti-solar direction. We interpret the tail as being caused by dust particles accelerated by solar radiation pressure. The sudden appearance and the morphology of the tail indicate that the dust particles are small, with an effective radius ~1 micrometer and a combined mass ~3x10^5 kg. These particles are likely products of thermal fracture and/or desiccation cracking under the very high surface temperatures (~1000 K) experienced by Phaethon at perihelion. The existence of the tail confirms earlier inferences about activity in this body based on the detection of anomalous brightening. Phaethon, the presumed source of the Geminid meteoroids, is still active.
astro-ph.EP
astro-ph
Revised 2013-June-11 The Dust Tail of Asteroid (3200) Phaethon David Jewitt1,2, Jing Li1 and Jessica Agarwal3 1Department of Earth and Space Sciences, University of California at Los Angeles, 595 Charles Young Drive East, Los Angeles, CA 90095-1567 2Department of Physics and Astronomy, University of California at Los Angeles, 430 Portola Plaza, Box 951547, Los Angeles, CA 90095-1547 3 Max Planck Institute for Solar System Research, Max-Planck-Str. 2, 37191 Katlenburg-Lindau, Germany [email protected] ABSTRACT We report the discovery of a comet-like tail on asteroid (3200) Phaethon when imaged at optical wavelengths near perihelion. In both 2009 and 2012, the tail appears (cid:38)350(cid:48)(cid:48) (2.5×108 m) in length and extends approximately in the projected anti-solar direction. We interpret the tail as being caused by dust particles accelerated by solar radiation pressure. The sudden appearance and the morphology of the tail indicate that the dust particles are small, with an effective radius ∼1 µm and a combined mass ∼3×105 kg. These particles are likely products of thermal fracture and/or desiccation cracking under the very high surface temperatures (∼1000 K) experienced by Phaethon at perihelion. The existence of the tail confirms earlier inferences about activity in this body based on the detection of anomalous brightening. Phaethon, the presumed source of the Geminid meteoroids, is still active. Subject headings: minor planets, asteroids: general - minor planets, asteroids: individual (3200 Phaethon) - comets: general - meteorites, meteors, mete- oroids 1. Introduction Asteroid (3200) Phaethon is a ∼5 km diameter body dynamically associated with the Geminid meteoroid stream (Whipple 1983) and with several kilometer-scale asteroids collec- tively known as the Phaethon-Geminid complex (PGC; Ohtsuka et al. 2009, Kasuga 2009). – 2 – Most meteoroid streams have cometary parents (Jenniskens 2008) from which mass loss is driven by the sublimation of near-surface ice. However, Phaethon is dynamically an asteroid (semimajor axis 1.271 AU, eccentricity 0.89, inclination 22.2, Tisserand parameter relative to Jupiter, TJ = 4.54), raising questions about the mechanisms by which it loses mass. The formation of the PGC could be ancient (Ohtsuka et al. 2009), but the short dy- namical lifetime of the Geminid meteoroid stream (∼103 yr; Gustafson 1989; Ryabova 2007), opens the possibility that Phaethon might still be active. However, attempts over two decades to detect activity in Phaethon have proved negative (Cochran & Barker 1984; Chamberlin et al. 1996; Hsieh & Jewitt 2005; Wiegert et al. 2008). The first evidence for continuing activity was obtained only recently. Photometry in both 2009 (Jewitt & Li 2010) and 2012 (Li & Jewitt 2013) showed anomalous perihelion brightening, in which the apparent bright- ness increased suddenly at large phase angles, opposite to the fading trend expected from the phase function of a solid body. Numerous mechanisms (thermal emission, glints, fluorescence stimulated by the impact of the solar wind, sublimation of embedded ice, prompt emission from forbidden transitions in atomic oxygen) were considered and found incapable of pro- ducing the anomalous brightening (Li and Jewitt 2013). In particular, near-surface water ice is thermodynamically unstable on Phaethon as a result of its high surface temperature. Deeply-buried water ice would be thermally insulated and phase-lagged from the surface heat, leaving no explanation for the coincidence between activity and perihelion (Jewitt & Li 2010) and (Li & Jewitt 2013). Instead, the release of dust from the nucleus is able to explain the data in a plausible way. The process responsible for forming and ejecting the dust is presumed to relate to the high (∼1000 K) temperatures attained by the surface of Phaethon when at perihelion (q = 0.14 AU). Thermal fracture and cracking due to desic- cation shrinkage of hydrated silicates are two processes capable of both producing the dust and ejecting it from the surface (Jewitt & Li 2010; Jewitt 2012; Li & Jewitt 2013). Observations at perihelion are extremely challenging, because the solar elongation then is small (<8◦) and Phaethon must be viewed against the bright, structured and changing background of the solar corona. Here, we use data from the NASA-STEREO coronal imaging spacecraft to perform a search for spatially-resolved evidence of activity at perihelion. 2. Observations We used the Heliospheric Imagers (HI) from the Sun Earth Connection Coronal and Heliospheric Investigation (SECCHI) package (Howard et al. 2008; Eyles et al. 2009) on the NASA STEREO spacecraft. Our observations exclusively employed the STEREO-A HI-1 camera, having a field center offset from the solar center by 14◦ and with a square field of – 3 – view 20◦ in width. The 2048×2048 pixel charge-coupled device (CCD) detectors are binned 2×2 before transmission to Earth. The resulting angular size of each pixel is 70(cid:48)(cid:48). Each 1024×1024 pixel-image is compiled from a set of 30 integrations each of 40 s, and taken at 1 minute intervals. A single downloaded image therefore has an effective exposure time of 1200 s (20 minutes). One such image is obtained every 40 minutes. The onboard combination of multiple short-exposure images permits the rejection of cosmic rays and other artifacts, and avoids saturation of the background corona that would otherwise occur owing to the large pixels in HI-1. The quantum efficiency of the camera is practically uniform across the 6300 to 7300 A wavelength passband (Eyles et al. 2009). We searched for extended emission in the HI-1 images used in our earlier work (Li & Jewitt 2013). Our procedure removed large angular scale structures in the coronal back- ground, but left small scale and rapidly varying features, as well as background stars. We used NASA's HORIZONS software to compute the position of Phaethon as seen from the STEREO spacecraft and calculated the expected location on the CCD in pixel coordinates. When displayed in rapid succession as a movie, the images from 2009 hint at the presence of a tail on Phaethon, but fluctuations in the surface brightness of the coronal and sky background from image to image are much larger than the surface brightness of the tail itself. Simple median stacks show that the tail appears concurrently with the anomalous brightening but is otherwise absent. Unlike the background fluctuations, the statistical significance of the tail grows as more images are combined. To test the possibility that the tail might be an artifact produced by only a fraction of the data, we separately combined subsets of the images (0.5 to 0.8 days at a time). The subsets all showed the tail but, as expected, at lower significance owing to the smaller number of images in the subsets. Next, since the projected antisolar direction, θ(cid:12), changes rapidly in the period of interest, we re-combined the images including a correction for the changing θ(cid:12). We removed field stars from the images by hand prior to computing the median of an image stack. The resulting images improve the apparent brightness of the tail and show that it is aligned with the projected sun-Phaethon line (Figure 1). To test the possibility that the improvement in the de-rotated images might be a result of chance in noisy data, we repeated the procedure but for a wide range of unphysical rotations. In these unphysical image combinations, the Phaethon tail became washed out or invisible, as expected if the tail is real. To test the possibility that the tail might be caused by a peculiar asymmetry or astigmatism in the images from the HI-1 camera, we examined the images of comparably bright field stars located close to the path of Phaethon in the period of interest. The stars showed no asymmetry and no evidence for a Phaethon-like tail (see the inset images of field stars in each panel of the 2009 data in Fig. 1). Lastly, we note that no tail was detected in the image composite having start-time – 4 – UT 2009 June 19d 06h 49m (i.e. one day pre-perihelion, left panel of Figure 1). The tail was detected only on the two subsequent days, coinciding with the anomalous brightening reported in Jewitt & Li (2010) and Li & Jewitt (2013). There are no useful later data from the STEREO spacecraft. The entire procedure was repeated using the data from 2012, with the same result (Figure 2). In both years Phaethon shows a faint, approximately antisolar tail that becomes brighter when the blurring effects of differential image rotation in the image sequence are correctly removed and fainter, to the point of disappearing, when they are not. The tail appears only on the two days for which Phaethon showed anomalous brightening (Li & Jewitt 2013). We measured θP h, the position angle of the Phaethon tail and present the results in Table (1), along with the geometric circumstances of Phaethon in each year. The uncertainties on θP h, determined from azimuthal surface brightness profiles centered on Phaethon, reflect the large pixel scale, the faintness of Phaethon's tail and the complexity of the sky background. The measured θP h are also plotted in Figure (3), where it may be seen that θP h and θ(cid:12) are identical within the uncertainties of measurement. 3. Discussion Thermal emission, specular reflection "glints", fluorescent excitation by the solar wind and prompt emission from the excited 1D level of [OI] (produced by the photo-destruction of water) were all considered and rejected as sources of the anomalous perihelion brightening (Jewitt & Li 2010; Li & Jewitt 2013). The first three would all produce brightening only of the nucleus (i.e. the central pixel in our data) and hence are additionally inconsistent with the detection of a resolved tail. Sodium is depleted in Geminid meteors (Kasuga 2009) and might be baked-out from the nucleus of Phaethon. The Na D-lines at ∼5890A, however, fall outside the 6300 to 7300 A passband of the HI-1 camera and so cannot contribute to the tail. Prompt emission from the forbidden lines of oxygen at 6300A and 6363A falls within the instrumental passband but would require a large production rate of ∼1030 s−1 to match the observed brightening (Li & Jewitt 2013). Furthermore, the photodissociation lifetime of water in sunlight at 0.14 AU is only τd = 0.5 hr (Huebner et al. 1992). It is unlikely that water molecules could travel the length of the tail in such a short time. Our preferred interpretation is that the tail of Phaethon is a dust tail. The key observ- ables are the time of appearance of the tail, the length of the tail and the position angle of the tail in the plane of the sky (Table 1). We computed the running median of ∼30 images having a range of start-times around perihelion. In both years, the emergence of the tail cor- responded with times of perihelia (2009 June 20 07:22 and 2012 May 02 07:49). The length – 5 – of the tail in the plane of the sky was estimated at (cid:96) ∼ 250,000 km in both 2009 and 2012. Since our ability to identify the end of the tail is limited by signal-to-noise considerations, the measured (cid:96) constitutes only a lower limit to the true length. First, we estimate the dust properties from these measurements. The length and rise- time of the tail, τ ∼ 1 day, imply an average speed V = (cid:96)/τ ∼ 3 km s−1, which can be produced at a constant acceleration a = 2(cid:96)/τ 2 ∼ 0.07 m s−2. For comparison, the solar gravitational acceleration at 0.14 AU is g(cid:12) = 0.3 m s−2, giving a ratio β = a/g(cid:12) ∼ 0.2. The ratio of accelerations for a particle moving under the action of radiation pressure can be written in terms of particle properties as β = 3QprF(cid:12)R2 1 4GM(cid:12)cρr . (1) Here, Qpr is the dimensionless radiation pressure factor, F(cid:12) = 1360 W m−2 is the Solar constant, R1 = 1.5×1011 m is the number of meters in 1 AU, G = 6.6×10−11 N kg−2 m2 is the gravitational constant, M(cid:12) = 2×1030 kg is the mass of the Sun and c = 3×108 m s−1 is the speed of light. Particle quantities ρ and r are the density and radius, respectively. We assume Qpr = 1, ρ = 3000 kg m−3 and substitute β = 0.2 into Equation (1) to obtain r ∼ 1 µm. In other words, the sudden emergence of the long tail implies a large β, corresponding to particles about 1 µm in radius. Particles of this size have scattering parameter x = 2πr/λ ∼ 9 at the λ = 0.7 µm wavelength of observation, near the peak scattering efficiency (Bohren & Huffman 1983). Presumably, much smaller particles exist but contribute weakly to the effective cross-section because they are inefficient scatterers (x (cid:28) 1) while larger particles (x (cid:29) 1) may exist and scatter efficiently, but are relatively rare and slow-moving (and would be confined to the vicinity of the nucleus in our data). We next computed syndyne (particles with a single β ejected over a range of times) and synchrone (particles with a range of β ejected at one time) models of the trajectories of dust particles. The models take account of orbital motion and projection into the plane of the sky as viewed from STEREO-A. In making these trajectory calculations we assume that the particles are ejected from Phaethon with negligible initial velocity. Model results for 2009 June are plotted in Figure (4). Equivalent calculations for 2012 May give comparable results and are not shown. The syndynes for large particles (small β) would occupy a tail having a position angle inconsistent with the one observed, as would synchrones for particles older than a few days (see Figure 4). The latter result is further consistent with the onset of the anomalous perihelion brightening within a day of our first tail detection. The position angle of the tail and the sudden growth of the tail are both consistent with the action of radiation pressure on small particles. – 6 – Estimates of the dust mass from the photometry taken in 2009 and 2012 (Li & Jewitt 2013), give Md = 2.5×105rµm and Md = 4×105rµm, respectively, where rµm is the particle radius expressed in microns. We take the average, Md ∼ 3×105 rµm, and substitute rµm = 1 to estimate the mass of efficient scatterers in the tail as Md ∼ 3×105 kg. This mass is miniscule compared to the estimated nucleus mass (2×1014 kg) and Geminid stream mass (1012 to 1013 kg; Hughes & McBride 1989; Jenniskens 1994). If ejected uniformly over τ ∼ 1 day, the average mass loss rate from Phaethon would be dM/dt ∼ Md/τ ∼ 3 kg s−1. At this rate, the timescale for replenishment of the Geminid stream mass (∼1012 to 1013 s or 2×104 to 2×105 yr) is far longer than the ∼103 yr dynamical stream age (Gustafson 1989; Ryabova 2007). It thus seems unlikely that the particles contributing to the optical tail in Figures (1) and (2) are sufficient to supply the Geminid meteoroid stream. This conclusion is reinforced by the syndyne/synchrone models, which further show that particles with β > 0.07, like those dominating the optical tail, are gravitationally unbound to the solar system. Such particles cannot be a significant source of Geminid stream meteoroids. On the other hand, much larger, slower, potentially mass-dominant particles could exist while contributing little to the scattering cross-section at optical wavelengths. Such particles would be subject to smaller acceleration by radiation pressure (c.f. Equation 1) and would remain in the unresolved vicinity of the nucleus in our data. Furthermore, we see no reason to assume that mass loss, even at perihelion, should occur in a steady state. It is entirely possible, for instance, that the perihelion mass loss rate varies stochastically (perhaps by orders of magnitude) from orbit to orbit, analogous to the way in which steady erosion of a coastal headland by ocean waves leads to rare but mass-dominant landslides. Thus, we can conclude that the inferred mass loss rate is too small to supply the Geminids in steady state, but we cannot use the new data to rule-out the possibility that Phaethon continues to actively supply its own meteoroid stream. – 7 – 4. Summary We have discovered a tail on Geminid-parent asteroid (3200) Phaethon at perihelion. The tail unambiguously establishes the presence of on-going mass-loss, confirming our prior inferences based on unresolved photometry alone. The key features and conclusions from this tail are: 1. The tail grows to full length ((cid:38) 250,000 km) within a single day, implying acceleration from the nucleus at 0.07 m s−2 or greater (about 0.2 times the local solar gravitational acceleration). This large acceleration is consistent with the action of radiation pressure on spherical dust grains ∼ 1 µm in radius. 2. Taken together, the photometry and the inferred grain size indicate a tail mass ∼3×105 kg and a mass production rate ∼3 kg s−1. 3. Most particles in the optical tail follow gravitationally unbound orbits and thus do not contribute to the Geminid meteoroid stream. Much larger, slower, potentially mass- dominant and gravitationally-bound particles could be simultaneously ejected from Phaethon but would escape detection in our data. 4. Previously suggested mechanisms of thermal fracture and desiccation cracking of hy- drated minerals remain plausible sources of Phaethon's tail. We thank Toshi Kasuga and Pedro Lacerda for reading the paper and Michael A'Hearn for his review. This work was supported by a grant to DCJ from NASA's Planetary As- tronomy program. The Heliospheric Imager instrument was developed by a collaboration that included the University of Birmingham and the Rutherford Appleton Laboratory, both in the UK, the Centre Spatial de Liege (CSL), Belgium, and the U.S. Naval Research Lab- oratory (NRL), Washington DC, USA. The STEREO/SECCHI project is an international collaboration. – 8 – REFERENCES Bohren, C. F., and Huffman, D. R. 1983, Absorption and Scattering of Light by Small Particles (New York: Wiley) Chamberlin, A. B., McFadden, L.-A., Schulz, R., Schleicher, D. G., & Bus, S. J. 1996, Icarus, 119, 173 Cochran, A. L., & Barker, E. S. 1984, Icarus, 59, 296 Tsiganis, K., Morbidelli, A., & Licandro, J. 2010, A&A, 513, A26 Eyles, C. J., Harrison, R. A., Davis, C. J., et al. 2009, Sol. Phys., 254, 387 Gustafson, B. A. S. 1989, A&A, 225, 533 Howard, R. A., Moses, J. D., Vourlidas, A., et al. 2008, Space Sci. Rev., 136, 67 Hsieh, H. H., & Jewitt, D. 2005, ApJ, 624, 1093 Huebner, W. F., Keady, J. J., & Lyon, S. P. 1992, Ap&SS, 195, 1 Hughes, D. W., & McBride, N. 1989, MNRAS, 240, 73 Jenniskens, P. 1994, A&A, 287, 990 Jenniskens, P. 2008, Earth Moon and Planets, 102, 505 Jewitt, D. 2012, AJ, 143, 66 Jewitt, D., & Li, J. 2010, AJ, 140, 1519 Kasuga, T. 2009, Earth Moon and Planets, 105, 321 Li, J., & Jewitt, D. 2013, AJ, 145, 154 Ohtsuka, K., Nakato, A., Nakamura, T., et al. 2009, PASJ, 61, 1375 Ryabova, G. O. 2007, MNRAS, 375, 1371 Yang, B., & Jewitt, D. 2010, AJ, 140, 692 Whipple, F. L. 1983, IAU Circ., 3881, 1 Wiegert, P. A., Houde, M., & Peng, R. 2008, Icarus, 194, 843 This preprint was prepared with the AAS LATEX macros v5.2. – 9 – Table 1. Observing Geometry and Tail Position Angles UT Date and Timea Rb ∆c αd e θP h θ(cid:12)[deg]f 2009 June 19d 06h 49m 0.147 20d 06h 49m 0.140 21d 06h 49m 0.146 2012 May 01d 08h 09m 0.146 02d 08h 09m 0.140 03d 08h 09m 0.146 1.027 0.972 0.916 1.052 1.003 0.950 57 79 102 46 67 88 - 128±23 108±16 - 124±17 110±18 131 124 116 137 126 117 aStart time of the image composite from which the tail properties were measured. Each composite consists of images taken over a period of one day with corrections for rotation of the projected antisolar direction applied. Other parameters in the table all refer to the start time. bHeliocentric distance, in AU cPhaethon to STEREO A distance, in AU dPhase angle, in degrees eMeasured position angle of the tail and estimated 1-σ uncer- tainty, in degrees fPosition angle of the projected anti-Solar direction, in degrees – 10 – Fig. 1.- Composite images of (3200) Phaethon in 2009 compared with the projected sun- Insets are 490(cid:48)(cid:48) square comet line (white). The Sun is to the upper right in each panel. and show field stars near to Phaethon to demonstrate the point spread function of the data. Each panel has North to the top, East to the left and shows the median of ∼30 images taken over a 1 day period starting at the times listed in Table 1. Anomalous brightening peaks were reported on June 20 and 21 in Jewitt and Li (2010). – 11 – Fig. 2.- Same as Figure (1) but for data from 2012 (see Table 1). Anomalous brightening peaks were detected on May 02 and 03 in Li and Jewitt (2013). – 12 – Fig. 3.- Measured position angle of the tail in 2009 (red circles) and 2012 (blue circles), compared with the projected antisolar direction (red and blue lines). Error bars on the position angle measurements show the estimated 1-σ uncertainties from composite images formed by averaging the data over 1-day intervals, as described in the text. – 13 – Fig. 4.- Phaethon on UT 2009 June 20, 06:49 compared with dust models. Synchrones (green) correspond to ejection at 100, 50, 30, 20, 10, 5.3, 4.3, 3.3, 2.3, 1.8, 1.3, 0.8 days before the date of the image. Syndynes (red) correspond to β = 0.002, 0.005, 0.01, 0.02. 0.04, 0.07, 0.1, 0.2, 0.4, 0.9, as marked.
1210.5585
1
1210
2012-10-20T07:17:09
Spitzer 3.6 micron and 4.5 micron full-orbit lightcurves of WASP-18
[ "astro-ph.EP" ]
We present new lightcurves of the massive hot Jupiter system WASP-18 obtained with the Spitzer spacecraft covering the entire orbit at 3.6 micron and 4.5 micron. These lightcurves are used to measure the amplitude, shape and phase of the thermal phase effect for WASP-18b. We find that our results for the thermal phase effect are limited to an accuracy of about 0.01% by systematic noise sources of unknown origin. At this level of accuracy we find that the thermal phase effect has a peak-to-peak amplitude approximately equal to the secondary eclipse depth, has a sinusoidal shape and that the maximum brightness occurs at the same phase as mid-occultation to within about 5 degrees at 3.6 micron and to within about 10 degrees at 4.5 micron. The shape and amplitude of the thermal phase curve imply very low levels of heat redistribution within the atmosphere of the planet. We also perform a separate analysis to determine the system geometry by fitting a lightcurve model to the data covering the occultation and the transit. The secondary eclipse depths we measure at 3.6 micron and 4.5 micron are in good agreement with previous measurements and imply a very low albedo for WASP-18b. The parameters of the system (masses, radii, etc.) derived from our analysis are in also good agreement with those from previous studies, but with improved precision. We use new high-resolution imaging and published limits on the rate of change of the mean radial velocity to check for the presence of any faint companion stars that may affect our results. We find that there is unlikely to be any significant contribution to the flux at Spitzer wavelengths from a stellar companion to WASP-18. We find that there is no evidence for variations in the times of eclipse from a linear ephemeris greater than about 100 seconds over 3 years.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 17 (2008) Printed 23 March 2021 (MN LaTEX style file v2.2) Spitzer 3.6µm and 4.5µm full-orbit lightcurves of WASP-18 P.F.L. Maxted1, D.R. Anderson1, A. P. Doyle1, M. Gillon2, J. Harrington3, N. Iro1, E. Jehin2, D. Lafreni`ere4, B. Smalley1, J. Southworth1 1Astrophysics Group, Keele University, Keele, Staffordshire ST5 5BG 2 Institut d'Astrophysique et de G´eophysique, Universit´e de Li`ege, All´ee du 6 Aout 17, Bat. B5C, 4000 Li`ege, Belgium 3Planetary Sciences Group, Department of Physics, University of Central Florida, Orlando, FL 32816-2385, USA 4D´epartement de physique, Universit´e de Montr´eal, C.P. 6128 Succ. Centre-Ville, Montr´eal, QC, H3C 3J7, Canada Submitted 2012 ABSTRACT We present new lightcurves of the massive hot Jupiter system WASP-18 obtained with the Spitzer spacecraft covering the entire orbit at 3.6 µm and 4.5 µm. These lightcurves are used to measure the amplitude, shape and phase of the thermal phase effect for WASP-18 b. We find that our results for the thermal phase effect are limited to an accuracy of about 0.01 per cent by systematic noise sources of unknown origin. At this level of accuracy we find that the thermal phase effect has a peak-to-peak amplitude approximately equal to the secondary eclipse depth, has a sinusoidal shape and that the maximum brightness occurs at the same phase as mid-occultation to within about 5 degrees at 3.6 µm and to within about 10 degrees at 4.5 µm. The shape and amplitude of the thermal phase curve imply very low levels of heat redistribution within the atmosphere of the planet. We also perform a separate analysis to determine the system geometry by fitting a lightcurve model to the data covering the occultation and the transit. The secondary eclipse depths we measure at 3.6 µm and 4.5 µm are in good agreement with previous measurements and imply a very low albedo for WASP- 18 b. The parameters of the system (masses, radii, etc.) derived from our analysis are in also good agreement with those from previous studies, but with improved precision. We use new high-resolution imaging and published limits on the rate of change of the mean radial velocity to check for the presence of any faint companion stars that may affect our results. We find that there is unlikely to be any significant contribution to the flux at Spitzer wavelengths from a stellar companion to WASP-18. We find that there is no evidence for variations in the times of eclipse from a linear ephemeris greater than about 100 seconds over 3 years. Key words: stars: individual: WASP-18; planetary systems 2 1 0 2 t c O 0 2 . ] P E h p - o r t s a [ 1 v 5 8 5 5 . 0 1 2 1 : v i X r a 1 INTRODUCTION Hot Jupiters are currently at the forefront of observa- tional studies that can provide meaningful tests for mod- els of exoplanet atmospheres. The atmospheric tempera- tures for a typical hot Jupiter orbiting a solar-type star with a period of 3 days can be up to 1500 K. For tran- siting hot Jupiters this makes it feasible to measure the planet-star flux ratio directly from the depth of secondary eclipse in the lightcurve due to the occultation of the exo- planet by the host star. Early results with the Spitzer Space Telescope confirmed the existence of secondary eclipses in the lightcurves of HD 209458 (Deming et al. 2005) and TrES-1 (Charbonneau et al. 2005) with the expected depth ∼ 0.5 per cent at mid-infrared wavelengths. The secondary eclipse depth has now been measured using Spitzer for more than 20 hot Jupiters (Cowan & Agol 2011). Comparison of these observations with atmospheric models has been used to reveal the diversity of hot Jupiter atmospheres with re- gard to their composition (Madhusudhan et al. 2011), the presence or absence of a temperature inversion in their atmo- spheres, (Knutson et al. 2010), and their albedos and heat recirculation efficiencies (Cowan & Agol 2011). The secondary eclipse depth for a hot Jupiter at in- frared wavelengths measures the brightness temperature of the hemisphere facing the star -- the "day-side" -- integrated over the visible hemisphere. This brightness temperature, Tday, will depend on the pattern of emission over the day side, the spectral energy distribution (SED) of this emission, the Bond Albedo, Ab, and the efficiency with which heat is redistributed to the night-side of the planet. Observations at several wavelengths, particularly near-infrared observations c(cid:13) 2008 RAS 2 P.F.L. Maxted et al. near the peak of the day-side SED, reduce the extent to which we must rely on models to account for the conversion from brightness temperature to effective temperature when interpreting these observations (Madhusudhan et al. 2011). In general, it is not possible to disentangle the degeneracy between Bond albedo and heat recirculation efficiencies from secondary eclipse observations alone. One exception is the case of HD 189733, in which very high quality Spitzer ob- servations of the secondary eclipse reveal asymmetries that can be inverted to produce a map of the brightness temper- ature on the day side of HD 189733 b (Majeau et al. 2012; de Wit et al. 2012). Apart from this exceptional case, it is currently only possible to obtain information on the redistri- bution of heat in the planet's atmosphere by observing the thermal phase effect -- the variation in infrared brightness of the system as a function of orbital phase. There are several different ways to parametrize the re- distribution of heat from the day-side to the night-side of a planet (Spiegel & Burrows 2010). Here we use the parameter Pn, the fraction of the incident energy that is transported to the night side of the planet. Plausible values of this pa- rameter vary from 0 up to 0.5. A value of Pn = 0 would imply a night-side brightness temperature Tnight ≪ Tday, whereas Pn = 0.5 implies Tnight ≈ Tday.1 If Pn > 0 then this suggests that winds at some level in the atmosphere move heat around the planet. In practise, very high efficiencies for strongly irradiated planets are unlikely because the winds that transport heat to the night-side will dissipate some of their energy through turbulence or shocks (Goodman 2009). Nevertheless, some redistribution of energy from the day- side to the night-side is likely, and may lead to significant offsets between the sub-stellar point and the hottest regions of the atmosphere (Cooper & Showman 2005). This will be observed in the thermal phase effect as an offset in the phase of maximum brightness from opposition. Cowan et al. (2007) used 8 separate observations with Spitzer spread throughout the orbit of the three hot Jupiter systems HD 209458, HD 179949 and 51 Peg to measure their thermal phase effect. They were able to place useful upper limits on the phase variation in 51 Peg and HD 209458 and to detect a variation with a peak-to-trough amplitude of 0.14 per cent in HD 179949. HD 179949 is a non-transiting hot Jupiter, so the inclination of the orbit and the radius of the planet are unknown, but even allowing for this uncer- tainty, the observed amplitude of the phase variation pro- vided an upper limit of Pn < 0.21 and shows that the hottest point is near the sub-stellar point. HD 209458 is a transit- ing system, so the inclination of the orbit and the radius of the exoplanet are known. This allowed Cowan et al. to translate the upper limit on the amplitude of the thermal phase into a lower limit Pn > 0.32, and thus establish that apparently different hot Jupiters are likely to have a variety of Pn values. Harrington et al. (2006) detected the phase variation of the planet υ And b using the MIPS instrument on Spitzer at 24µm. With additional data and an improved understand- ing of the systematic noise sources in MIPS, they were able to refine their estimate of the amplitude of the phase vari- 1 The brightness temperatures are not necessarily equal since the SED from the day-side and night-side may be different. ation and show that there is a large (∼ 80◦) phase off- set between the time of maximum brightness and oppo- sition (Crossfield et al. 2010). The inclination and radius of υ And b are not known accurately because it is a non- transiting exoplanet. This can result in a large uncertainty in the value of Pn inferred from the amplitude of the phase curve (Burrows et al. 2008). However, the large phase off- set observed in υ And b implies a large Pn despite the large amplitude of the phase variation. The thermal phase curve of HD189733 b has been ob- served using Spitzer at 3.6µm and 4.5µm (Knutson et al. 2012), at 8µm (Knutson et al. 2007) and at 24µm (Knutson et al. 2009). There are also multiple observa- tions of the transits and eclipses at 8µm for this system (Agol et al. 2010). The combined analysis of these results by Knutson et al. (2012) shows that heat recirculation from the day-side to the night-side is efficient for this relatively cool hot Jupiter (Tday ≈ 1200 K) and that this recircula- tion leads to a peak in the thermal phase effect that occurs ∼ 25 degrees before opposition. Cowan et al. (2012) obtained Warm Spitzer photom- etry covering the complete 26 hour orbit of the very hot Jupiter WASP-12 b at 3.6 µm and 4.5 µm. They found that their interpretation of the lightcurves depends on the as- sumptions made about the nature of the systematic noise in the lightcurves and that red noise is the dominant source of uncertainty in their analysis. Nevertheless, they were able to show that the thermal phase variation in WASP- 12 b is large, indicative of poor day-to-night heat redistri- bution (Pn . 0.1). The small offset they observe between the phase of maximum brightness and secondary eclipse (16 ± 4 degrees) in the 4.5 µm data is consistent with this interpretation. The phase offset at 3.6 µm could not be de- termined unambiguously from their data. Although thermal phase curves are only currently avail- able for a few systems, there does appear to be a pattern of weak recirculation for the hottest planets. Cowan & Agol (2011) have looked for trends in the value of Tday/Tsub, where Tsub is the equilibrium temperature of the sub-stellar point, in a sample of 24 transiting exoplanets with secondary eclipse measurements. This quantity will depend on both the albedo of the planet and the recirculation efficiency, but a large value can only be obtain if both the albedo and the re- circulation efficiency are low. Cowan & Agol found that this is the case for all of the 6 hottest planets (Tday > 2400 K) in their sample and point out that this is, in general terms, the expected behviour given that the radiative timescale scales as T −3 whereas the advective timescale (which they assume to be of-order the local sound speed) scales as T −0.5. This simple scaling argument does not explain the apparent transition in behaviour at Tday ≈ 2400 K but Perna et al. (2012) do observe a transition at about this temperature in their suite of three-dimensional circulation models for hot Jupiters. This is mainly due to the change in the ratio of the radiative and advective timescales, with the presence of an atmospheric inversion playing a lesser role in determining the recirculation efficiency. This transition may also be re- lated to the onset of ionisation of alkali metals in the planet's atmosphere, leading to severe magnetic drag (Perna et al. 2010). There is an ongoing debate as to whether the result- ing Ohmic dissipation can transport energy into the interior of the planet and so explain the very large radii observed for c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 some hot Jupiters (Laughlin et al. 2011; Rauscher & Menou 2012; Perna et al. 2012; Huang & Cumming 2012). Here we present Warm Spitzer photometry covering the complete orbit of the very hot Jupiter WASP-18 b. This ex- oplanet is unusual for its combination of short orbital pe- riod (0.945 d) and high mass (10 MJup), which results in strong tidal interactions between the planet and the star (Hellier et al. 2009). Southworth et al. (2009) derived ac- curate masses and radii for the star and planet in the WASP-18 system based on high quality optical photometry of the transit and the spectroscopic orbit from Hellier et al.. Nymeyer et al. (2011) used Spitzer photometry in all four IRAC bands covering the secondary eclipse of WASP-18 to measure the brightness temperature of the day side from 3.6 µm to 8.0 µm. The high brightness temperatures derived (∼ 3200 K) imply that WASP-18 b has near-zero albedo and almost no redistribution of energy from the day side to the night side of the planet. Our primary aim is to use our Warm Spitzer photome- try at 3.6 µm and 4.5 µm to measure the amplitude, phase and shape of the thermal phase effect. We also use these data to re-measure the secondary eclipse depths at 3.6 µm and 4.5 µm for comparison with the results of Nymeyer et al. (2011). We measure the times of eclipse and transit from our data and from new optical photometry of several transits and use these together with published times of mid-eclipse to re-measure the eclipse ephemeris and to look for possi- ble variations in the period. We consider the likelihood that WASP-18 has a stellar companion based on published radial velocity data and new high-resolution imaging at H-band and K-band. The contamination of the lightcurve for a hot Jupiter system by a companion star has the potential to bias the results obtained if not properly accounted for. Compan- ions stars may also play a role in the formation and evo- lution of some hot Jupiter systems (Fabrycky & Tremaine 2007; Mardling 2007). Our analysis also provides an accu- rate characterisation of the primary eclipse (transit) which can be used in combination with other data to re-measure the mass and radius of the star and planet. 2 OBSERVATIONS 2.1 Spitzer photometry We were awarded Spitzer General Observer time during Cy- cle 62 to observe two complete orbits of WASP-18 with IRAC (Fazio et al. 1998), one orbit with each of the two IRAC channels operating during the warm mission. Observations with channel 1 (3.6µm) were obtained on 2010 January 23, observations with channel 2 (4.5µm) were obtained on 2010 August 23. On both dates, 243 200 images with an expo- sure time of 0.36s were obtained in sub-array mode. The total duration of each sequence of observations is 29 hours. In addition, a sequence of 64 sub-array images also with an exposure time of 0.36s were obtained immediately after the observations of WASP-18 at a slightly offset position. These were used to check for hot pixels or other image artifacts on Spitzer lightcurves of WASP-18 3 the detector. We used Basic Calibrated Data (BCD) pro- cessed with version S18.18 of the Spitzer IRAC pipeline for our analysis. 2.2 AO imaging We obtained adaptive optics high resolution H- and K-band images of WASP-18 using the NICI instrument at Gemini South. The observations were obtained on the night of 2010 December 27 under good seeing (0.5 -- 0.6 arcsec). The in- strument was configured with the CLEAR focal plane mask and the H 50/50 beam-splitter, and we used the narrow band filters Kcont (2.2718 µm) and FeII (1.644 µm) in the red and blue channels, respectively. We observed the target at five dither positions corresponding to the center and corners of a square of side 6 arcsec. At each dither position we obtained three images consisting of the co-addition of 3 exposures of 1.5 s. The full-width at half-maximum (FWHM) of the point spread function (PSF) in these images was 0.065 arcsec in the FeII filter and 0.073 arcsec in the Kcont filter. Data reduction consisted of subtracting a sky image, di- viding by the flat field, and fixing bad pixels by interpolating from neighbouring pixels. The sky image was created from the median combination of the images at the different dither positions after masking out the star signal in each image. The reduced images were registered to a common position and field orientation and then combined using the median value of each pixel. No other point source was detected in the resulting images. The sensitivity of our AO imaging to detect faint companions was determined by first computing the median absolute deviation of the pixel values within an- nuli of various radii and width equal to one PSF FWHM. The resulting contrast curve was then properly scaled, and verified to be adequate, by adding and recovering (by vi- sual inspection) fake companions in the images at various separation and with various contrasts. In doing this last ex- ercise we used both the K- and H-band images to differenti- ate speckles from true companions, which display a different chromatic behaviour. Using this approach, we estimate the detection limits in difference of magnitudes to be 4.0 mag at >0.2′′, 5.4 mag at >0.4′′ and 6.0 mag at >0.5′′. Finally, we note the presence of a faint ghost in the image at 13 pixel (0.23′′) separation and contrast of ∼4.1 mag in the Kcont filter and ∼5.9 mag in the FeII filter. 2.3 TRAPPIST photometry Five transits of WASP-18 b were observed with the 60 cm robotic telescope TRAPPIST3 (Jehin et al. 2011; Gillon et al. 2012) located at ESO La Silla Observatory (Chile). TRAPPIST is equipped with a thermo-electrically- cooled 2k × 2k CCD camera with a field of view of 22′ × 22′ (pixel scale = 0.65 arcsec pixel−1). The first four transits were observed in an Astrodon 'I + z' filter that has a trans- mittance > 90 per cent from 750 nm to beyond 1100 nm, the red end of the effective bandpass being defined by the spec- tral response of the CCD. The last transit was observed in the Sloan z' filter. For all transits, the telescope was slightly defocused to minimize pixel-to-pixel effects and to optimize 2 PI: P. Maxted, program ID 60185 3 http://www.ati.ulg.ac.be/TRAPPIST c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 4 P.F.L. Maxted et al. Table 1. Details of the transit lightcurves obtained with TRAP- PIST for WASP-18. For each lightcurve we list the observation date, the filter used, the number of measurements, and the expo- sure time. Date Filter Np Texp [s] 2010 Sep 30 2010 Oct 02 2010 Dec 23 2011 Jan 08 2011 Nov 10 I+z I+z I+z I+z z' 712 977 688 648 624 12 8 6 6 10 the observational efficiency. For each run, the stellar images were kept on the same pixels, thanks to 'software guiding' system deriving regularly astrometric solutions on the sci- ence images and sending pointing corrections to the mount if needed. Table 1 gives the logs of these TRAPPIST obser- vations. After a standard pre-reduction (bias, dark, flatfield cor- rection), the stellar fluxes were extracted from the images using the iraf/daophot4 aperture photometry software (Stetson 1987). For each transit, several sets of reduction parameters were tested, and we kept the one giving the most precise photometry for the stars of similar brightness as WASP-18. After a careful selection of reference stars, dif- ferential photometry was obtained. 3 SPITZER DATA ANALYSIS 3.1 Conversion to flux and noise model We converted the BCD images from units of MJy/steradian to mJy using the values for the pixel size at the centre of the subarray provided in the image headers (1.225′′ × 1.236′′ for channel 1, 1.205′′ × 1.228′′ for channel 2). We used the raw images together with the values of the gain and readout noise for each channel to calculate the noise level in each pixel assuming Poisson counting statistics. 3.2 Image times BCD data in sub-array mode are delivered as FITS files con- taining a data cube of 64 images of 32 × 32 pixels per file. We used the FITS header keyword BMJD OBS to assign a Barycentric UTC modified Julian date (BMJD) to the start of the exposure for the first image in the data cube. The BMJD of the mid-exposure time for each image in the data cube was then calculated using the values for the start and end times of the integration from the FITS header (AINT- BEG and ATIMEEND) to calculate the time taken to ob- tain the 64 images and assuming that these images were uniformly spaced in time. Long observing sequences such as the ones we have used 4 iraf is distributed by the National Optical Astronomy Obser- vatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. for WASP-18 cannot be executed using standard observing modes so multiple instrument engineering requests (IERs) are used to obtain the data. The images obtained in the sec- ond of the two IERs used for our WASP-18 observations do not have the coordinates of the target in the FITS header and so the light-time correction to the solar system barycen- tre are incorrect for these images. We calculated the light- time correction for the images obtained before the interrup- tion from the difference in the keyword values BMJD OBS − MJD OBS. We then use a linear extrapolation of this light- time correction as a function of MJD OBS to calculate the BMJD of the images obtained after the interruption based on their MJD OBS values. The uncertainty in the exposure time introduced by this procedure is negligible. 3.3 Outlier rejection We compared each image to the other 63 images in the same data cube in order to identify discrepant data points in the images. We are particularly concerned here with identifying discrepant pixel values that may affect the photometry of the target. As the target moves during the sequence of 64 images, we use a robust linear fit (least absolute deviation) to the 64 pixel values from each file for each pixel to predict the expected value for each pixel value in each image. We then flag the pixels in each image that deviate from their expected value by more than 5 times their standard error. We find that the number of pixels flagged using this method is much larger than expected given the known incidence of cosmic ray hits on the IRAC detectors. This discrepancy is due to a few pixels well away from the target position that are noisier than predicted by our noise model. As these pixels have no effect on our photometry and a negligible effect on the estimate of the background level we ignore this discrepancy. We also flagged any pixels in our images that are flagged as bad pixels in the "Imask" file provided for each BCD file by the Spitzer IRAC pipeline. 3.4 Sky background estimate We use the mean of the image pixel values excluding those within 10 pixels of the target position to estimate the back- ground value in each image. Values more than 4 standard deviations from the mean and flagged pixels were ignored in the calculation. We used a Gaussian fit to a histogram of these pixel values to estimate the standard deviation of the background pixel values, σbg. The number of points used to estimate the background was ≈ 700. Typical values of σbg are 0.0033 mJy/pixel for channel 1 and 0.0025 mJy/pixel for channel 2. 3.5 Aperture photometry We tried three different methods to measure the location of the star on the detector, the daophot cntrd and gcntrd algorithms and a least-squares fit of a bivariate Gaussian distribution to an 11×11 sub-image centred on the nomi- nal star position. We refer to this latter algorithm as the gauss2d method. We used a fixed value for the full-width at half-maximum of FWHM= 1.25 pixels for both axes of c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 5 diffuse than the stellar images and look more like the loga- rithm of the point spread function, as described in the IRAC instrument handbook (Version 2.0.1, p. 116) . The image ar- tifact in the channel 1 offset sky images has up to about 0.6 mJy per pixel in the channel 1 image and a total of about 6 mJy within an aperture with a radius of 5 pixels. For com- parison, a typical channel 1 image of WASP-18 has a peak flux of 50 -- 70 mJy/pixel and a total flux of 163 ± 3 mJy within an aperture of the same radius, so the image artifact affects the photometry of WASP-18 by a few per cent. For the channel 2 data the corresponding figures are a total of about 1 mJy/pixel in the artifact compared to 42 mJy/pixel in the peak and a total flux of 103 ± 1 mJy in the images of WASP-18 so for this channel image persistence affects the photometry by about 1 per cent. The IRAC instrument handbook describes the be- haviour of image persistence artifacts in the IRAC arrays during the warm mission. In channel 1 the artifacts decay exponentially with a timescale of about 4.5 hours. Channel 2 residuals start out as positive, but then become negative with a decay timescale of a few minutes. For channel 1 data we make the assumption that the image artifact will be approximately constant after some time during the exposure sequence comparable to the de- cay timescale. To correct for the effect of image persistence we create a "master offset image" from the median of the 64 offset sky images, subtract the background value from this image and then subtract the result from the images of WASP-18. This correction will be inaccurate for some frac- tion of the data at the start of the observing sequence while the image persistence builds-up. We discuss this point fur- ther below. For the channel 2 data it is not clear how the image ar- tifact affects the photometry of WASP-18. The interval be- tween the end of the observing sequence for WASP-18 and the start of the offset sky image is 49 s, which is comparable to the decay timescale for the artifact. The exact form and timescale for the decay of the artifact is not known so it is not even possible to make a precise estimate of contribution of the image artifact to the measured flux for WASP-18. How- ever, it is likely that the image artifact contributes less than 2 per cent given the decay timescale for this feature is a few minutes. We did attempt to measure the decay timescale from the data taken subsequent to our own observations, but the artifact was not detectable in those images. For the channel 2 data we treat the contribution of the image persis- tence artifact as an additional source of uncertainty in our analysis. 3.7 Initial assessment of the data The flux of WASP-18 measured with an aperture radius of 4.5 pixels is shown for both channels in Fig. 2. Also shown in this figure are the positions of the star on the array cal- culated using the gauss2d method. The coordinates x and y are measured relative to the centre of a corner pixel in the sub-array. The form of the variations in the x,y posi- tions measured using the gcntrd and cntrd algorithms are similar, but the amplitudes of the variations are less and there is an offset between these values and the results of the gauss2d method. For example, the y positions measured for the channel 1 data using the gcntrd method have me- Figure 1. Images obtained before and after our WASP-18 obser- vations. All images are linearly scaled (inverse grey scale) between 0 and 10 MJy/sr. The "before" images are 30 s exposures in the region of IC 2560. The "after" images are the median of the 64 offset sky images with an exposure time of 30 s each. the Gaussian profile in the gauss2d method based on the re- sults of fitting the images with the FWHM as a free param- eter. The cntrd algorithm determines the position where the derivatives of the image values go to zero. The gcntrd algorithm fits a Gaussian profile to the marginal x and y dis- tributions of the image values. We set the input parameters to cntrd and gcntrd to run on a sub-image of 5× 5 pixels around the target position. We compare the performance of these different algorithms below. We used the IDL Astronomy Users library5 implemen- tation of the daophot aper procedure (Stetson 1987) to perform synthetic aperture photometry on our images. We used the 2006 November version of this procedure which al- lowed us to use the option to use an exact calculation of the intersection between a circular aperture and square pixel for correct weighting of pixels at the edge of the aperture. We used 13 aperture radii uniformly distributed from 1.5 pixels to 4.5 pixels. The results we obtained for larger aper- ture radii were not useful because the lightcurves have much lower signal-to-noise due to the additional background noise. Fluxes measured from images containing any flagged pixels in the aperture were rejected from further analysis, although much less than 1 per cent of images were affected in this way. The median number of rejected pixels per image is 2. 3.6 Image persistence Our IRAC images are affected by image persistence, partic- ularly the channel 1 images. This can be seen in the offset sky images obtained immediately after our WASP-18 ob- servations (Fig. 1). The resulting image artifacts are more 5 http://idlastro.gsfc.nasa.gov/ c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 6 P.F.L. Maxted et al. dian value of 14.95 with 98 per cent of the data in the range y = 14.78 -- 14.95, cf. a median value of 14.88 and range y = 14.65 -- 14.99 for the gauss2d method. The feature that stands out from Fig. 2 is the well- known correlation between the measured flux and the posi- tion of the star on the detector, particularly in channel 1. This position-dependent sensitivity variation (PDSV) makes it difficult to see the transit and secondary eclipse in these "raw" aperture flux measurements. The channel 2 data ap- pear to be less affected by PDSV, so the transit and sec- ondary eclipse can be seen in the raw flux measurements. PDSV is a combination of the "pixel phase effect" described in the IRAC instrument handbook (Version 2.0.1, p. 45) and pixelation noise. The pixel phase effect is a variation in the sensitivity of each detector pixel that depends on the dis- tance of the stellar image from the centre of the pixel. The motion in the x and y directions for our channel 2 data result in a smaller variation in the distance of the star from the cen- tre of the pixel compared to the channel 1 data, which may partly explain why the data quality is better in this chan- nel (Anderson et al. 2011). Pixelation noise affects synthetic aperture photometry with small aperture radii because for a pixel at the edge of the aperture, the fraction of the flux falling within that pixel is not the same as the fraction of the aperture within the pixel. approximately the same period, then it would become im- possible to determine whether any variation in the measured flux with P ≈ 0.94 is due to the flux variation of WASP-18 or due to the PDSV. Understanding the correlations between the lightcurve model parameters and the correction for PDSV is prob- lematic in the case of our WASP-18 data because of the large number of parameters required to model the complete lightcurve plus the large number of parameters that may be required to characterise the complex structure ("corru- gation") in the PDSV. One method we experimented with was to use the pixel sensitivity map method of Ballard et al. (2010) applied to the residuals to a least-squares fit of a lightcurve model. This approach can be applied iteratively until the solution and pixel sensitivity map converge. The problem with this approach is that it becomes difficult to identify correlations between the lightcurve model parame- ters and the pixel sensitivity map. We avoided this problem by excluding the data during the transit and occultation from our analysis of the thermal phase effect. The main ad- vantage of this approach is that fitting a model to the data between the eclipses can be reduced to a linear least-squares problem. This makes is straightforward to find the best so- lution of the problem and to investigate the correlations be- tween the free parameters of the model. 3.8 Analysis of the thermal phase effect 3.8.2 Model for PDSV and the thermal phase effect 3.8.1 Correction for PDSV The usual method developed to correct for PDSV in IRAC data for observations of the secondary eclipses of hot Jupiters is to include parameters in the least-squares fit of an eclipse model to the data to represent the PDSV. This is usu- ally a simple linear or quadratic relation between sensitivity and each of the coordinates x and y (e.g., Beerer et al. 2011; Anderson et al. 2011). Ballard et al. (2010) have developed an alternative method to correct for PDSV in their Warm Spitzer 4.5µm observations of GJ 436. They created a pixel sensitivity map from the data themselves. This approach was straightforward in the case of GJ 436 because the lightcurve of the target is expected to be constant apart from the possi- ble presence of transits affecting a small fraction of the data. The pixel sensitivity map generated by Ballard et al. for the IRAC channel 2 shows complex structure that they describe as "corrugation . . . low-level sinusoidal-like variations with a separation of approximately 5/100ths of a pixel between peaks". A similar concept based on bi-linear interpolation rather than a smoothed look-up table has been developed by Stevenson et al. (2012) and applied to Spitzer photome- try of HD 149026 b. For our WASP-18 data we are interested in character- ising the amplitude and shape of the phase variation as well as measuring the shape and depths of the transit and sec- ondary eclipse. The phase variation has a period comparable to the length of the observing sequence so it is important to understand any correlations between the correction for the PDSV and the parameters of the lightcurve model. The phase variation of WASP-18 can be modelled approximately as a sinusoidal variation in flux with P=0.94 d. For exam- ple, in the worst case scenario, if the position of the target on the detector also varied approximately sinusoidally with Our model for the measured magnitude of the system be- tween the eclipse and transit is j=1 aj cos(jφi) +PNsin k=1 bk sin(kφi) (1) mi = c0,0 +PNcos λ=1PNxy +PNx +PNxy ι=1 cι,0pι(x′ i) +PNy κ=1 c0,κpκ(y ′ i) i)pµ(y ′ i), µ=1 cλ,µpλ(x′ where mi is the magnitude of WASP-18 at time ti; φi = 2π(ti − T0)/P is the orbital phase relative to the time of mid-transit, T0; pn is a Legendre polynomial of order n; x′ i = (xi − ¯x)/(xmax − xmin) and similarly for y ′ i (xmin is the minimum value of xi, etc.). We use the values T0 = BMJD 54220.98163 for the time of mid-transit and P = 0.94145299 d for the orbital period from Hellier et al. (2009) throughout this paper unless otherwise stated. By using Leg- endre polynomials and normalized coordinates (x′ i) we find that we can use singular value decomposition to find solutions of this least-squares problem for Legendre polyno- mials up to at least 12th order. This is sufficient to model the corrugations with a scale of 0.05 pixels seen by Ballard et al. if they are present in our data. i, y ′ 3.8.3 Linear decorrelation against position In Fig. 3 we show the result of using the simplest reasonable model for our data, in which the magnitude of the phase variation varies sinusoidally and the PDSV is linear in x and y, i.e., mi = c0,0 + a1 cos(φi) + b1 sin(φi) + c1,0p1(x′ i) + c0,1p1(y ′ i). Note that our calculations are done using magnitudes, but we plot the results as fluxes and quote parameter values in unit of per cent. The least-squares fit of this model to c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 7 Figure 2. The flux of WASP-18 measured in IRAC channel 1 (left panel) and channel 2 (right panel) measured with a circular aperture of radius of 4.5 pixels.. The position of the star on the array measured using the gauss2d method is shown below each panel. For clarity, we have only plotted a random selection of 1 per cent of the data here. Dashed lines indicate the start and end times of the transit and secondary eclipse assuming a duration of 0.08 d for each and assuming that the secondary eclipse occurs at phase 0.5. the unbinned aperture photometry outside of eclipse and transit is used to determine the coefficients of the model for the PDSV. We then apply this correction to all the data. The results in Fig. 3 are for an aperture radius of 2.5 pixels for both apertures and target positions measured using the gauss2d method. This is the combination of aperture radius and positions that gave the lowest RMS residual for the data between the transit and eclipse. Results for other apertures and for cntrd and gcntrd methods are similar. It is clear that a linear correction is insufficient to fully remove the effect of the PDSV, but this simple model does show clearly some features of our data. Firstly, we see that the eclipse and transit are clearly visible in both channels. Two transits are visible in the channel 2 data but the first transit is not seen clearly in the channel 1 data because there is a large "ramp" affecting the first few hours of the data. There is a cosine- like variation in flux observed in both channels with the maximum flux occurring near phase 0.5 (secondary eclipse). Part of this signal is the phase variation we wish to measure. However, there must also be some instrumental component or other systematic effect that contributes to this variation because the phase variation due to the planet cannot have an amplitude larger than the secondary eclipse depth. c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 The obvious suspect for the systematic noise source in the channel 1 data is the image artifact shown in Fig. 1. The "ramp" is the right size (≈ 3 per cent) and builds up over the same sort of timescale as the known decay timescale of this artifact. Our interpretation of this lightcurve is that the image artifact builds up over the first 6 -- 8 hours before reaching an approximate equilibrium between the arrival of new photons from WASP-18 and its own decay timescale. We have tried several methods to account for this ramp- like feature in the data but none of these methods is any better than the more pragmatic approach of simply exclud- ing the first 6 -- 8 hours of the channel 1 data. Without go- ing into the details of these various methods, we can state here that we almost always found that the amplitude of the phase variation measures in channel 1 was similar to the depth of the secondary eclipse and often was slightly larger. It is possible to create models for the systematic noise in the channel 1 lightcurves that achieve more physically real- istic (lower) values for the amplitude of the phase variation, but these models are not based on any physical model of the instrumental noise, i.e., they are arbitrary, and they re- quire several additional free parameters that are often not well constrained by the data or any physical understand- ing of what these parameters represent. The overall quality 8 P.F.L. Maxted et al. Figure 3. The flux of WASP-18 measured in IRAC channel 1 (left panel) and channel 2 (right panel) after a linear correction for position dependent sensitivity variations and a simple sinusoidal model for the phase variation. The zero-point of the flux scale is set from the mean flux during secondary eclipse. Observations obtained during the transit and secondary eclipse (small points) were excluded from the calculation of the coefficients for the decorrelation. The data have been binned into 0.0025 d bins for display purposes only. of the decorrelated lightcurve obtained with these arbitrary and complex models of the the instrumental noise is also not much better than the best results presented below for the partial lightcurve. Clearly, a more complete understanding of the instrumental noise in IRAC for warm mission obser- vations would be a great help for the interpretation of our data, but in the absence of this we present the results for the partial lightcurve and make an attempt to quantify the extent to which instrumental noise introduces systematic er- rors in our results. 3.8.4 Optimum decorrelation against position. We used the model given in equation (1) to fit the data for channel 1 and channel 2 excluding data within 0.05 d of mid-transit and mid-eclipse and also excluding the first 60 000 observations (7.3 hours of data) for channel 1. We used Ncos = 2 and Nsin = 1 to model the phase variation of WASP-18. The sine term allows for a phase shift from phase 0.5 for the time of maximum brightness and the first harmonic of the cosine variations (coefficient a2) allows for some optimisation of the shape of this phase variation. To model the PDSV we tried Nx = Nxy = Ny/2 = 1, 2, . . . 6. We use Ny = 2Nx because there is a larger range of motion in the y direction. We fit these models to the lightcurves for all combinations of aperture size and position measurement methods. We used the Bayesian information criterion (BIC) to identify the combination of (Nx, Ny, Nxy) that provides the best compromise between quality of fit and number of free parameters for a given lightcurve. We calculated the BIC using the expression BIC = χ2 + Npar loge(N ), where Npar is the number of free parameters and N is the number of observations. We used the RMS of the residuals to identify the aperture size and position measurement method that give the best lightcurves. For both channels we find the best results are obtained for (Nx, Ny, Nxy) = (5, 10, 5) with positions measured using the cntrd method and an aper- ture radius of 2.25 pixels. These models and lightcurves are shown in Fig. 4 and the parameters of interest are given in Table 2. The standard error estimates given in Table 2 ac- count for the correlations between parameters (Press et al. 1992) although the correlation coefficients between the pa- rameters in this table and all other parameters in the model are small (< 0.3). In Fig. 5 we show how the parameters a1, a2 and b1 obtained for (Nx, Ny, Nxy) = (5, 10, 5) vary as a function of c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Table 2. Results of linear least squares fits to the phase variation between eclipses using the model given in equation (1). These results are for (Nx, Ny, Nxy) = (5, 10, 5) with positions measured using the cntrd method and an aperture radius of 2.25 pixels. N is the number of points included in the fit and BIC is the Bayesian information criterion as defined in the text. A is the amplitude of the thermal phase effect and φmax is the phase of maximum brightness relative to phase 0.5. Random and systematic errors are given for each quantity in that order. Parameter Channel 1 Channel 2 a1 [%] a2 [%] b1 [%] A [%] φmax N χ2 BIC RMS [%] 0.183 ± 0.004 ± 0.01 0.148 ± 0.005 ± 0.01 0.023 ± 0.005 ± 0.01 0.003 ± 0.005 ± 0.01 0.001 ± 0.003 ± 0.01 −0.006 ± 0.003 ± 0.01 0.296 ± 0.009 ± 0.02 0.366 ± 0.007 ± 0.02 0.001 ± 0.003 ± 0.01 −0.010 ± 0.006 ± 0.02 179851 216717.3 217259.0 0.717 133124 142398.6 142928.2 0.539 aperture radius and position measurement method for our various lightcurves. Also plotted in Fig. 5 are the ampli- tude of the phase variation and the offset from phase 0.5 to the time of maximum brightness in phase units. There is some dependence on aperture radius for these results, e.g., the value of a1 for both channels show a trend to- wards smaller values with increasing radius. We also see that there is worse agreement between the results for dif- ferent position measurement methods for smaller apertures as a result of the increased sensitivity of the pixelation noise to small differences in the assumed position. For all of the coefficients in both channels we see that the results vary by about ±0.01 per cent as a function of aperture radius. We therefore assume that systematic noise limits the accuracy of these results to ±0.01 per cent. Despite the limit of ±0.01 per cent in the accuracy of these results, we are able to draw some definite conclusions about the thermal phase effect in WASP-18. Firstly, the am- plitude of the thermal phase effect is very similar to the depth of the secondary eclipse. This can be seen in Fig. 4 and by comparing the values for the amplitude in Table 2 to the eclipse depths given in Table 3. Secondly, the off- set between phase 0.5 and the time of maximum brightness due to the thermal phase effect is consistent with 0 to within about 0.01 phase units for the channel 1 data and 0.02 phase units for the channel 2 data. Thirdly, the parameter a2 is also consistent with the value 0 so the shape of the thermal phase variation is sinusoidal to within the limits sets by the systematic noise. 3.9 Eclipse model We tried several different methods to model the entire lightcurve for each channel including both eclipses, the phase variation and PDSV, but were not able to find any method that gave reliable results. We suspect that there is some fac- tor other than position on the detector that introduces sys- tematic noise at the level of ∼0.01 per cent with a timescale of ∼day. This can be seen in Fig. 4, where there are clear c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 9 systematic errors remaining in the lightcurve at this level. These systematic errors are not removed by increasing the complexity of the model used for the PDSV. It will be diffi- cult to identify this additional factor given that little infor- mation about the shape of the point spread function can be measured from the undersampled IRAC images. There are many published studies that have used IRAC photometry obtained over ∼ 5 hours of observation to suc- cessfully model hot Jupiter eclipses, so we decided to only model the data within 0.1 phase units of the primary and secondary eclipses. We fit these data simultaneously using a single model to account for the true flux variations of the system. We then account for systematic errors in the mea- sured flux independently for the data around each eclipse. We used have the NDE lightcurve model (Nelson & Davis 1972; Popper & Etzel 1981; Etzel 1981) to model the primary and secondary eclipses in our lightcurves. This model uses biaxial ellipsoids to approximate to pro- jected area of the star/planet. Gim´enez (2006) has shown that this model used with an integration ring size of one degree (as we have done) can be used to model planetary transits with a precision of ∼ 4 × 10−5, which is sufficient for our purposes. From inspection of our model lightcurves we find that ∼ 1 4 of the model data points during primary eclipse are affected by numerical noise at this level and that there is no numerical noise during secondary eclipse. We created a double-precision version of the NDE model that has negligible numerical noise, but that runs appreciably slower than the original single-precision code. We used our double-precision version to verify that the numerical noise in the single-precision version has a negligible effect on our results, so all the results presented here are based on the single-precision version. We use the NDE model to calculate ℓs(φ) and ℓp(φ), the contribution of the star and planet, respectively, to the total apparent flux at any given phase, φ, including the effects of tidal distortion and eclipses. Note that ℓs and ℓp include the effects of the eclipses and transits and the ellipsoidal variation of both star and planet. To model the variation in magnitude due to the phase effect of the planet we use the harmonic series h(φ) = a1 cos(φ) + b1 sin(φ) + a2 cos(2φ) − hmax, where the values of a1, b1 and a2 are taken from the least- squares fit to the data between transit and eclipse for the same aperture size and position measurement method and hmax is chosen such that the maximum value (corresponding to minimum flux) of h(φ) is 0. The apparent magnitude of the system is then given by mi = m0 − 2.5 log [ℓs(φi) + ℓp(φi)] + ℓp(φi)h(φi)/ℓp,max, where ℓp,max is a normalization factor. Our calculations are done using magnitudes but we present the results in flux units or as percentages. In addition, we model the PDSV independently for the data around primary and secondary eclipse using Legendre polynomial functions of the x and y position plus an optional linear function of time. For each set of lightcurve model parameters we calculate the optimum values of the PDSV model parameters using singular value decomposition to fit the residuals from the lightcurve model. The parameters of the NDE lightcurve model of rele- 10 P.F.L. Maxted et al. Figure 4. The flux of WASP-18 measured in IRAC channel 1 (left panel) and channel 2 (right panel) in an aperture of radius 2.25 pixels after correction for position dependent sensitivity variations (PDSV) for the parameter sets (Ncos, Nsin, Nx, Ny, Nxy) = (2, 1, 5, 10, 5) and positions measured with the cntrd method. Data are plotted averaged in 200 s bins for clarity and the best-fit sinusoidal model is also shown. The mean value in secondary eclipse is indicated with a dotted line. Note that data in eclipse (small points) are not included in the fit. The PDSV model is shown as a function of time in the middle panels and as a function of position as a grey-scale plot in the lower panels. The grey-scale is linear between ±4 per cent for channel 1 and 1 per cent for channel 2 with positive values being white. vance to our study are: J, the surface brightness of the planet in units of the central surface brightness of the star exclud- ing the thermal phase contribution; r1 = Rstar/a, the radius of the star in units of the semi-major axis; r2 = Rplanet/a, the radius of the planet in units of the semi-major axis; i, the inclination, u⋆, the linear limb-darkening coefficient for the star; e cos(ω) and e sin(ω), where e is the orbital ec- centricity and ω is the longitude of periastron. We fix the mass ratio of the system at the value mplanet/mstar = 0.01. We did not use this combination of parameters directly as free parameters in our least-squares fitting because there are significant correlations between them. Instead, we introduce the following parameters which are more directly related to the observed features of the lightcurve: ∆mtr, ∆moc, W , S. The parameters of the NDE lightcurve model are then calculated as follows: r2 r1 π k = r1 = 2.5 ∆mtr; = r ln(10) 2√kpW 2(1 − S2); r2 = kr1; b = r (1 − k)2 − S2(1 + k)2 1 − S2 ; c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 11 Figure 5. Coefficients of the sinusoidal model for the phase variation in WASP-18 as a function of aperture radius. The parameter set (Nx, Ny, Nxy) = (5, 10, 5) was used for the correction for position dependent sensitivity. Plotting symbols are as follows: cntrd -- squares; gcntrd -- diamonds; gauss2d -- filled circles. The solid line in the upper panels shows the semi-amplitude of the phase variation for the cntrd results. The solid line in the lower panel is the the phase offset from phase 0.5 for time of maximum brightness for the thermal phase effect derived from the cntrd results. Points have been offset horizontally by ±0.05 for clarity. i = cos−1 (br1) ; J = 1 − u⋆/3 ln(10)∆moc − 1(cid:17) 2.5 . k2(cid:16) are adapted parameters from These Seager & Mall´en-Ornelas (2003) so that, for a circular orbit, ∆mtr is the depth of the primary eclipse in mag- nitudes, ∆moc is magnitude difference between the flux duration occultation and the minimum of the thermal phase curve; W is the width of the transit in phase units and S is the duration of the ingress phase of the transit in units of W . The intermediate variables used here are k, the radius ratio and b, the impact parameter. We also include a correction to the time of mid-transit compared to the ephemeris of Hellier et al., ∆T0. We are careful here to define what we mean by the depth of the secondary eclipse because the variation in flux due to the thermal phase effect on the timescale of the eclipse is comparable to the precision with which we can measure the depth from our photometry (∼ 0.01 per cent) and the maximum of the thermal phase effect may not occur at mid-eclipse. For ease of calculation, interpretation and com- c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 parison with other measurements, we simply measure the mean flux during occultation (excluding ingress and egress phases), fin, and the mean flux either side of the eclipse in a region ± 0.1 phase units around the time of mid-eclipse, fout and define the secondary eclipse depth to be D = fout−f in . The fluxes are measured from the lightcurve corrected for PDSV. fout There are some second-order effects not accounted for by our model. We do not account for the brightness distri- bution on the day-side of the planet, but this will have a negligible effect on our results given that the thermal phase effect is not strongly peaked and is symmetrical about phase 0.5, so this distribution will be approximately uniform and symmetrical. Doppler boosting is negligible compared to our signal-to-noise (. 0.001 per cent). We make a small correc- tion to the results for the effects of image persistence in the channel 2 data by assuming a dilution of the eclipses for 1±1 per cent. We have applied a correction to the apparent times of secondary eclipse for the light travel time across the or- bit of 2a/c ≈ 20 s so that the times and phases quoted here are the true time of mid-occultation relative to the apparent time of mid-transit. We also apply a correction to the values of e cos(ω) quoted below for this light travel time. 12 P.F.L. Maxted et al. Figure 6. Upper panel: Raw photometry for an aperture radius of 3 pixels (filled symbols -- channel 1, open symbols -- channel 2) together with the correction for PDSV based on positions measured using the cntrd method (lines). Lower panel: Photometry corrected for PDSV (points) and models fit by least-squares (lines). The channel 2 data have been vertically offset by 0.03 in the upper panel and 0.01 in the lower panel. In both panels the data and models are plotted in 60 s bins. We use the simplex algorithm of Nelder & Mead (1965) to optimise the least-squares fit of our model to the lightcurves. The simplex algorithm is a simple way to op- timize a least-squares solution given an initial set of param- eters, but it is not guaranteed to find the global minimum value of χ2 in the parameter space. In this case we are able to estimate accurate initial values for the most important parameters and so any solution will not be very far from the global minimum. However, we do find that numerical noise prevents us from using this algorithm by itself to find the optimum solution. We work around this problem by test- ing many initial starting values. We found that the solution with the lowest value of χ2 sometimes has parameters that are slightly biased when compared to other solutions with similar values of χ2 as a result of the numerical noise. We avoid this problem by taking the median value of each pa- rameter for all solutions with χ2 within 5 of the minimum as our best estimate of the parameter. The results for the depths of the eclipses and the RMS of the residuals for each data set are shown in Fig. 7. The depth of the primary eclipse (transit) can vary slightly with wavelength because the apparent radius of the planet will be larger at wavelengths where the atmosphere has a large opacity (Seager & Sasselov 2000). The size of this effect is approximately 2HRplanet/R2 star where H is the atmospheric scale height. In practice, for WASP-18 b the size of this ef- fect is negligible (≈ 0.001 per cent) because the large surface gravity of this massive planet makes the scale height of the atmosphere (∼ 40 km) much smaller than the size of the star. It can be seen from Fig. 7 that the solutions with the lowest RMS occur for an aperture radius 2 pixels, but the transit depths for channel 1 and 2 disagree by about 0.01 per cent for these data sets. The transit depths measured in channels 1 and 2 are consistent with each other for an aperture radius of 3 pixels and lie near the centre of the range of values obtained. The best fit to the lightcurves for an aperture radius of 3 pixels using the positions from the gcntrd method are shown in Fig. 6 and the parameters for the model used are given in Table 3. It can be seen that there is some residual corre- lated noise in the lightcurves after removal of the model for the PDSV. We quantified this residual correlated noise by calculating the RMS of the residuals after binning for a range of bin sizes (Pont et al. 2006). The results are shown in Fig. 8 and compared to the expectation for pure photon noise. At the timescale of the eclipse it can be seen that the channel 2 data are only weakly affected by correlated noise (. 0.005 per cent) but the channel 1 data are affected corre- lated noise at a level of 0.005 -- 0.01 per cent, particularly for the data covering occultation. Given that there is significant correlated noise in our data, we decided to calculate the random error on our model c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 13 Figure 8. RMS of the residuals after binning as a function of bin size (solid line) compared to the predicted photon noise (dotted line). The vertical, dashed lined in each panel shows the duration of eclipse and the duration of ingress/egress. Figure 9. Parameter correlation plots from our residual permuta- tion error analysis for channel 1 (blue crosses) and channel 2 (red diamonds). Our adopted values for each channel are indicated using dotted and dashed lines for channel 1 and 2, respectively. of the star. The observed value of Vrot sin I = 11 km s−1 (Hellier et al. 2009) together with the stellar radius imply a rotation period of about 6 d for WASP-18. We used the sine- wave fitting method described in Maxted et al. (2011) to cal- culate periodograms over 4096 uniformly spaced frequencies from 0 to 1.5 cycles/day. The false alarm probability (FAP) for the strongest peak in these periodograms was calculated using a boot-strap Monte Carlo method also described in Maxted et al. (2011). Variability due to star spots is not ex- pected to be coherent on long timescales as a consequence Figure 7. Depths of eclipses for transit and occultation mea- sured by fitting the eclipses. Channel 1 data are shown with filled symbols, channel 2 data with open symbols. Different symbols denote different position measurement methods. The RMS of the residuals of the fits are also shown using the same symbols. parameters using the "prayer-bead" method (Pont et al. 2006). This uses a circular permutation of the residuals by a random number of steps to create mock data sets. We applied the circular permutation to the residuals of the pri- mary and secondary eclipses independently and then used the simplex algorithm to fit models to 1024 mock data sets. The standard deviation of parameters from the fits is used to calculate the random errors for the model parameters given in Table 3 based on the analysis of the lightcurves for an aperture radius of 3 pixels using the positions from the gcntrd method. The random errors quoted include the ef- fect of the uncertainty in correcting for image persistence in the data. We use the range of values from the different apertures and position measurement methods to estimate the systematic errors on each parameter. The distribution of the parameters for the mock data sets is shown for some parameters of interest in Fig. 9. As can be seen from this fig- ure, the eclipse depths derived from the mock data sets can be biased by up to ∼ 0.005 per cent from the actual value. This is an consequence of the correlated noise in the resid- uals. We also used the results from these mock data sets to calculate the Pearson correlation coefficient, r, for all pairs of free parameters used in the least-squares fit. There is a weak anti-correlation between W and S (r ≈ −0.5), and a weak correlation between W and e sin(ω) (r ≈ 0.5), but the other free parameters are uncorrelated, as expected. 4 OPTICAL VARIABILITY The interpretation of our data would be considerably com- plicated by any intrinsic variability of the star WASP-18. We have analysed the WASP lightcurves of WASP-18 to determine whether they show periodic modulation due to the combination of magnetic activity and the rotation c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 14 P.F.L. Maxted et al. Table 3. Results of least squares fits to the primary and secondary eclipses. J ′ = J/(1 − u⋆/3) is the ratio of the integrated surface brightness of the star and the day-side of the planet. Other symbols are defined in the text. Random and systematic errors are given for each parameter in that order. Parameters that can be derived from the analysis of the optical TRAPPIST lightcurves are also given in the final column. Parameter Channel 1 Channel 2 TRAPPIST ∆mtr [%] ∆moc [%] D [%] W S u⋆ e cos(ω) e sin(ω) ∆T0 [s] r1 r2 k b i J ′ Phase of mid-occultation 0.969 ± 0.013 ± 0.007 0.015 ± 0.014 ± 0.009 0.304 ± 0.017 ± 0.009 0.0936 ± 0.0005 ± 0.0003 0.792 ± 0.009 ± 0.0007 0.06 ± 0.03 ± 0.06 0.0002 ± 0.0004 ± 0.0003 −0.003 ± 0.006 ± 0.004 −109 ± 8 ± 0 0.287 ± 0.006 ± 0.004 0.0281 ± 0.0006 ± 0.0003 0.0982 ± 0.0007 ± 0.0004 0.39 ± 0.06 ± 0.04 83.6 ± 1.0 ± 0.7 0.33 ± 0.02 ± 0.02 0.5003 ± 0.0006 ± 0.0004 0.979 ± 0.013 ± 0.009 0.028 ± 0.006 ± 0.017 0.379 ± 0.008 ± 0.013 0.0942 ± 0.0003 ± 0.0002 0.802 ± 0.008 ± 0.007 0.07 ± 0.03 ± 0.04 0.0001 ± 0.0002 ± 0.0003 −0.001 ± 0.003 ± 0.002 −108 ± 8 ± 0 0.282 ± 0.005 ± 0.004 0.0278 ± 0.0005 ± 0.0005 0.0987 ± 0.0006 ± 0.0004 0.32 ± 0.06 ± 0.04 84.8 ± 1.1 ± 0.8 0.40 ± 0.01 ± 0.02 0.5002 ± 0.0003 ± 0.0005 0.965 ± 0.056 0.0946 ± 0.0011 0.291 ± 0.017 0.0286 ± 0.0026 0.0983 ± 0.0030 0.41 ± 0.15 83 ± 3 Ntransit Noccultation χ2 RMS [%] 33996 36847 74213.9 0.55 35261 34355 90739.5 0.75 3649 2768.5 0.40 of the finite lifetime of star-spots and differential rotation in the photosphere and so we analysed the data from each observing season independently. We removed the transit sig- nal from the data prior to calculating the periodograms by subtracting a simple transit model from the lightcurve. In addition to the 2 seasons of data from Hellier et al. (2009) we also analyse 6041 observations obtained during the 2012 Jun -- Dec observing season. This date range covers the time of our Spitzer channel 2 observations. We did not find any significant periodic signals (FAP< 0.05) for WASP-18 apart from frequencies near 1 cycle/day due to instrumental effects. We examined the distribution of amplitudes for the most significant frequency in each Monte Carlo trial and used these results to estimate a 95 per cent upper confidence limit of 0.1 per cent for the am- plitude of any periodic signal in these WASP lightcurves. Beaulieu et al. (2010) have shown that the amplitude of the modulation at IRAC wavelengths due to star spots in solar- type stars is an order-of-magnitude smaller than at optical wavelengths. We conclude that any intrinsic variability of WASP-18 due to star spots has a negligible impact on our analysis. being constrained by normal priors based on the ephemeris presented by Nymeyer et al. (2011). The details of this anal- ysis are similar to the ones described in Gillon et al. (2012). The parameters derived from the least-squares fit to the 5 lightcurves are shown in Table 3. It can be seen that there is very good agreement between the parameters of the system dervied from the optical and infrared lightcurves. We have also measured a new time of mid-transit by using a least-squares fit of the NDE lightcurve model to the 2010 season of WASP data. The time of mid-transit quoted is close to the mid-point of dates of observation for these data and the standard error on the time of minimum is cal- culated using the prayer-bead method. All these times of mid-eclipse are given in Table 4 together with other pub- lished times of mid-eclipse. We used a least squares fit with a single value of the period and the times of mid-transit and mid-occultation as free parameters to determine a following linear ephemeris. TDB(mid − transit) = 2455265.5525(1) + 0.9414523(3) · E TDB(mid − occult.) = 2455266.0234(3) + 0.9414523(3) · E 5 ECLIPSE EPHEMERIS The analysis above provided two new, precise measurements of the time of mid-occultation and mid-transit. In addition we have 5 new times of mid-transit from our TRAPPIST observations. A global analysis of the five lightcurves was performed with the MCMC software described by Gillon et al. (2012). In addition to the baseline model and to the tran- sit ephemeris and shape parameters, the timings of the tran- sits were included as free parameters, the transit ephemeris The χ2 value for this fit was 21.8 with 11 degrees-of- freedom so the standard errors quoted in the final digits here have been scaled by p21.8/11. We also tried a quadratic ephemeris fit to the same data but found that this did not significantly improve the fit. The residuals from this O − C diagram for this linear ephemeris is shown in Fig. 10. It can be seen that the times of eclipse for WASP-18 have not varied by more than about 100 s over 3 years. c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Table 4. Apparent Barycentric Dynamical Time (TDB) of mid- transits (tr) and mid-occultation (oc) for WASP-18. The cycle number is calculated from our updated linear ephemeris and O−C is the residual from this ephemeris. Times of mid-occultation have been corrected for light-travel time across the orbit. BJDTDB − 2450000 Type Cycle O − C Source 4664.9061 ±0.0002 4820.7168 ±0.0007 4824.4815 ±0.0006 5220.8337 ±0.0006 5221.3042 ±0.0001 5392.6474 ±0.0004 5419.0083 ±0.0012 5431.7191 ±0.0003 5432.1897 ±0.0001 5470.7885 ±0.0004 5473.6144 ±0.0009 5554.5786 ±0.0005 5570.5842 ±0.0006 5876.5559 ±0.0013 tr oc oc oc tr tr tr oc tr tr tr tr tr tr −576 0.00013 −411 0.00019 −407 −0.00097 14 −0.00006 0.00017 15 197 0.00014 225 0.00015 238 −0.00114 239 −0.00087 280 −0.00057 0.00098 283 369 0.00026 0.00118 386 711 0.00097 1 2 2 a a 3 b a a c c c c c Triaud et al. 1: 3: http://var.astro.cz/ETD/; a: Spitzer c: TRAPPIST. (2010); 2: Nymeyer et al. (2011); IRAC; b: WASP; Figure 10. Residuals from our best-fit linear ephemeris for ob- served times of mid-transit (circles) and mid-occultation (squares) for WASP-18. The difference between our ephemeris and the ephemeris of Hellier et al. is also shown (dashed line). 6 WASP-18 STELLAR PARAMETERS A total of 21 individual HARPS spectra of WASP-18 were co-added to produce a single spectrum with a typical S/N of around 200:1. The analysis was performed using the meth- ods given in Doyle et al. (2012). The Hα and Hβ lines were used to give an initial estimate of the effective temperature (Teff ). The surface gravity (log g) was determined from the Ca i lines at 6162A and 6439A (Bruntt et al. 2010), along with the Na i D lines. Additional Teff and log g diagnostics were performed using the Fe lines. An ionisation balance between Fe i and Fe ii was required, along with a null de- pendence of the abundance on either equivalent width or excitation potential (Bruntt et al. 2008). This null depen- dence was also required to determine the micro-turbulence (ξt). The parameters obtained from the analysis are listed c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 15 Table 5. Stellar parameters of WASP-18 from our spectroscopic analysis. Parameter Value Teff log g ξt v sin i [Fe/H]a Massb Radiusb Sp. Typec Distance [K] [km s−1] [km s−1] [M⊙] [R⊙] [pc] 6400 ± 75 4.29 ± 0.10 1.20 ± 0.08 12.1 ± 0.5 0.10 ± 0.08 1.26 ± 0.09 1.25 ± 0.15 F6 130 ± 20 a[Fe/H] is relative to the solar value obtained by Asplund et al. (2009). b Mass and radius estimated using the Torres et al. (2010) calibration. c Spectral Type estimated from Teff using the table in Gray (2008). in Table 5. The value of [Fe/H] was determined from equiv- alent width measurements of several unblended lines, and additional least squares fitting of lines was performed when required. The quoted error estimates include that given by the uncertainties in Teff , log g, and ξt, as well as the scatter due to measurement and atomic data uncertainties. The projected stellar rotation velocity (v sin i) was de- termined by fitting the profiles of several unblended Fe i lines. A value for macroturbulence (vmac) of 4.6 ± 0.3 km s−1 was assumed, based on the calibration by Bruntt et al. (2010). An instrumental FWHM of 0.07 ± 0.01 A was deter- mined from the telluric lines around 6300A. A best fitting value of v sin i = 10.9 ± 0.7 km s−1 was obtained. The rotation rate (P = 5.8 ± 0.8 d) implied by the v sin i gives a gyrochronological age of ∼ 1.1+4.7 0.6 Gyr us- ing the Barnes (2007) relation. The value of Teff derived from our spectroscopic analysis agrees well with the value 6455±70 K derived by Maxted et al. (2011) from optical and near-infrared photometry using the infrared flux method. The distance derived here assuming that WASP-18 is a main-sequence star and quoted in Table 5 is consistent with the value 100 ± 10 pc derived from the Hipparcos parallax (van Leeuwen 2007). 7 PHYSICAL PARAMETERS Our new photometric and spectroscopic results allow for an improved determination of the physical properties of the WASP-18 system. We performed this analysis follow- ing the method of Southworth (2009), which requires as its input parameters measured from the lightcurves and spectra, plus tabulated predictions of theoretical models. From the lightcurves we adopted r1 = 0.284 ± 0.005, r2 = 0.0280± 0.0005 and i = 84◦ ± 1◦. The stellar Teff and [Fe/H] were taken from the spectroscopic determination in the pre- vious section, and the star's velocity amplitude was taken to be K1 = 1816.7 ± 1.9 m s−1 (Triaud et al. 2010). An initial value of the velocity amplitude of the planet, K2, was used to calculate the physical properties of the sys- tem with the physical constants listed by Southworth (2011). The mass and [Fe/H] value of the star were then used to ob- tain the expected Teff and radius, by interpolation within 16 P.F.L. Maxted et al. Table 6. Derived physical properties of the WASP-18 system. Parameter values are shown with random and, where appropriate, systematic errors, respectively. Parameter Value MA (M⊙) RA (R⊙) log gA (cgs) ρA (ρ⊙) Mb (MJup) Rb (RJup) gb (m s−2) ρb (ρJup) Teq (K) a (AU) Age (Gyr) 1.295 ± 0.052 ± 0.027 1.255 ± 0.027 ± 0.009 4.353 ± 0.017 ± 0.003 0.655 ± 0.035 10.52 ± 0.28 ± 0.15 1.204 ± 0.027 ± 0.008 179.9 ± 6.4 5.64 ± 0.31 ± 0.04 2411 ± 35 0.02055 ± 0.00028 ± 0.00014 0.4 +0.8 −0.9 +0.5 −0.3 one set of tabulated predictions from stellar theory. K2 was refined in order to find the best agreement between the ob- served and expected Teff , and the measured r1 and expected R1 a . This was performed for ages ranging from the zero-age main sequence to when the star was significantly evolved (log g < 3.5), in steps of 0.01 Gyr. The overall best fit was found, yielding estimates of the system parameters and also the stellar age. for details) plus a calibration of This procedure was performed separately using five dif- ferent sets of stellar theoretical models (see Southworth 2010, stellar prop- erties based on well-studied eclipsing binary star sys- tems (Enoch et al. 2010), with calibration coefficients from Southworth (2011). The results are given in Table 6, where we quote the mean value for each parameter, the random er- ror and an estimate of the systematic error from the range of values derived from the different stellar models, where ap- propriate. It can be seen from Table 6 that the results from different models are consistent to within the random errors on each parameter. In comparison to previous work, we have derived more precise radii, surface gravities and densities, for both com- ponents. We constrain the age of the star to be less than 1.7 Gyr, consistent with the gyrochronological age derived above. 8 POSSIBILITY OF CONTAMINATION BY A COMPANION STAR We have estimated the probability that our Spitzer photom- etry of WASP-18 is contaminated by the "third-light" from a companion star. It is not possible to detect a modest amount of third-light contamination directly from the lightcurve it- self because its only effect is to reduce the depths of the eclipses. It would be possible to find a good fit to a lightcurve affected by third-light contamination, but the parameters of the lightcurve model would be biased, e.g., k would be too small. Our calculation is based on the upper limit from our AO observations of 4.0 magnitudes for the brightness of any companion between 0.2 -- 2 arcseconds from WASP-18 and the upper limit of 43 m s−1 y−1 over a baseline of 500 days to the variation in the mean radial velocity of WASP-18 from Triaud et al. (2010). We assume that the probability distribution for the mass, eccentricity and period of the hy- pothesised companion is the same as the distributions for companions to solar-type stars from Raghavan et al. (2010). We approximated the distribution of companion masses us- ing a uniform distribution from 0.2 M⊙ to 0.8 M⊙ and used a uniform eccentricity distribution from 0 to 1. We then created a set of 65536 simulated binary stars with randomly selected periods, masses and eccentricities according to these distributions and randomly orientated orbits. We found that of these simulated binary stars, approximately 55 per cent would have been resolved by our AO imaging at the dis- tance of WASP-18; 20 per cent would have orbital periods less than 500 days and a semi-amplitude of 43 m s−1 or more and 25 per cent would show a change in radial velocity of 43 m s−1 or more over 500 days. This leaves only 5 per cent of hypothesised binaries as not detectable given our AO imaging and the published radial velocity data. The proba- bility that WASP-18 has a stellar companion is further re- duced because the overall binary fraction observed for planet hosting stars is approximately 25 per cent (Raghavan et al. 2010). An M-dwarf at the same distance as WASP-18 just below out detection limit of 4.0 magnitudes in the K-band at 0.2 arcsec would contribute no more than 5 per cent of the light at 4.5 µm. The more stringent limit of 6.0 magnitudes in the K-band that applies for separations of 0.5 -- 2.0 arcsec corresponds to an M-dwarf that contributes no more than 1 per cent at 4.5 µm. Of the simulated binary stars approxi- mately 45 per cent would be detected at this resolution. In conclusion, our AO imaging and the published ra- dial velocity data show that it is unlikely that WASP-18 has a stellar companion that significantly contaminates our Spitzer photometry. 9 DISCUSSION The values of D in Table 3 are in very good agreement the values 0.31 ± 0.02 at 3.6 µm and 0.38±0.02 per cent at 4.5 µm measured independently by Nymeyer et al. (2011). Their analysis of the secondary eclipse depths in 4 IRAC passbands suggests that the day-side atmosphere of WASP- 18 is likely to feature a temperature inversion. For zero albedo and zero redistribution of heat to the night side of the planet the integrated brightness temperature for the day- side is Tε=0 = (cid:0) 2 T0 = 3110 ± 35 K (Cowan & Agol 2011). For black-body emission this implies eclipse depths of 0.329 ± 0.005 at 3.6 µm and 0.379 ± 0.011 per cent at 4.5 µm, both in good agreement with the observed values. Zero-redistribution of heat within the atmosphere is also consistent with our observation that the peak of the ther- mal phase curve occurs close to the time of mid-occultation. Little can be said about the chemical composition of the day-side atmosphere at this stage because no strong molec- ular absorption or emission features have been detected from these secondary eclipse depth measurements. 3(cid:1)1/4 The amplitude of the thermal phase curve we have measured and the lack of a significant offset between the maximum in this curve and the time of mid-eclipse are both consistent with the conclusion based on the secondary eclipse depths that the albedo and recirculation efficiency for WASP-18 are both very low. This is consistent with c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 the hypothesis that very hot Jupiters have weak recircula- tion based mainly on secondary eclipse depth measurements alone (Cowan & Agol 2011). The good agreement between the recirculation efficiency inferred from the eclipse depths and measured from the thermal phase curve for WASP-18 strengthens this conclusion. The stellar limb darkening at infrared wavelengths is lower than at optical wavelengths and so the tran- sit produces a more "box-shaped" eclipse. This, combined with the precise photometry that is possible with Spitzer IRAC data, results in more precise estimates for param- eters such as Rstar/a, Rplanet/a and k = Rplanet/Rstar. Our results for these parameters agree well with the re- sults of Southworth et al. (2009). The agreement with the results of Triaud et al. (2010) is less good mainly because they find a larger stellar radius than our study (Rstar/a = 0.313 ± 0.010). The values of e cos(ω) and e sin(ω) derived from our analysis agree well with those of Triaud et al., but the value of e sin(ω) = 0.0085 ± 0.0009 they derive from their high quality radial velocity data is much more precise than ours and points to a small but significantly non-zero eccentricity. Arras et al. (2012) have argued that the small value of this apparent eccentricity combined with longitude of periastron very close to ω = 90◦ is exactly the signal ex- pected due to surface flows induced by tides on the planet. Their conclusion that the orbital eccentricity of WASP-18 b is less than 0.009 is consistent with the results of our analy- sis, although we are not able to confirm whether e ≪ 0.009 as they suggest. The measurement of the thermal phase effect for hot Jupiters using a continuous set of observations over an or- bital cycle with Warm Spitzer is not a well-established tech- nique, so it is useful to compare our experience of observing WASP-18 with the results using a similar observing strat- egy for WASP-12 obtained by Cowan et al. (2012) and for HD 189733 by Knutson et al. (2012). We find that system- atic errors of unknown origin limit the accuracy with which we can measure the amplitude of signals with time-scale comparable to the orbital period to about ±0.01 per cent. The main difficulty that Cowan et al. report in their WASP- 12 analysis is a signal on twice the orbital frequency in the 4.5 µm data that they tentatively attribute to the el- lipsoidal modulation of WASP-12 b. However, as they make clear, this signal is not seen in their 3.6 µm data and may be due to "uncorrected systematic noise". The ampli- tude of this "cos(2φ)" signal in their 4.5 µm data is about ±0.1 per cent, ten times larger than the level of systematic noise we find on these timescales. We do not see any sig- nal for the cos(2φ) harmonic in our 4.5 µm data greater than about 0.02 per cent. Knutson et al. use the wavelet- based method of Carter & Winn (2009) to account for sys- tematic noise in their full-orbit lightcurves of HD 189733 by assuming that this noise has a power spectral density vary- ing as 1/frequency. They find that the systematic noise con- tributes 0.0162 per cent of the total scatter in their channel 1 data -- comparable to the level seen in our data -- but only 0.0017 per cent in channel 2 -- much less than we see in our data. A systematic application of the wavelet-based method to archival Spitzer data may be a useful way to better understand the systematic noise sources in this in- strument and perhaps identify observing strategies that can reduce systematic noise levels. c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17 Spitzer lightcurves of WASP-18 17 10 CONCLUSIONS The amplitude, shape and phase of the thermal phase ef- fect we have measured from our Warm Spitzer lightcurves of WASP-18 are consistent with a sinusoidal variation with the same amplitude as and symmetric about the secondary eclipse, to within an accuracy ≈ 0.01 per cent set by some unknown source of systematic error. One contribution to this systematic error is likely to be the image persistence we observe from the offset images obtained immediately after our observations of WASP-18. This leads to the conclusion that WASP-18 b has a low albedo and that heat transport to the night-side of the planet is inefficient. This is the same conclusion reached by Nymeyer et al. (2011) based on the eclipse depths at 3.6 µm, 4.5 µm, 5.8 µm and 8.0 µm. The eclipse depths we measure at 3.6 µm and 4.5 µm are consis- tent with the previous measurements. ACKNOWLEDGEMENTS This work is based on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a con- tract with NASA. Support for this work was provided by NASA through an award issued by JPL/Caltech. PM would like to thank Felipe Menanteau for providing his proprietary data to us for the analysis of the image persistence artifacts. We thank Bryce Croll and Heather Knutson for enabling the AO observations of WASP-18 to be obtained and included in this manuscript. REFERENCES Agol E., Cowan N. B., Knutson H. A., Deming D., Steffen J. H., Henry G. W., Charbonneau D., 2010, ApJ, 721, 1861 Anderson D. R., Smith A. M. S., Lanotte A. A., Barman T. S., Collier Cameron A., Campo C. J., Gillon M., Har- rington J., Hellier C., Maxted P. F. L., Queloz D., Triaud A. H. M. J., Wheatley P. J., 2011, MNRAS, 416, 2108 Arras P., Burkart J., Quataert E., Weinberg N. N., 2012, MNRAS, 422, 1761 Asplund M., Grevesse N., Sauval A. J., Scott P., 2009, ARA&A, 47, 481 Ballard S., Charbonneau D., Deming D., Knutson H. A., Christiansen J. L., Holman M. J., Fabrycky D., Seager S., A'Hearn M. F., 2010, PASP, 122, 1341 Barnes S. A., 2007, ApJ, 669, 1167 Beaulieu J. P., Kipping D. M., Batista V., Tinetti G., Ribas I., Carey S., Noriega-Crespo J. A., Griffith C. A., Cam- panella G., Dong S., Tennyson J., Barber R. J., Deroo P., Fossey S. J., Liang D., Swain M. R., Yung Y., Allard N., 2010, MNRAS, 409, 963 Beerer I. M., Knutson H. A., Burrows A., Fortney J. J., Agol E., Charbonneau D., Cowan N. B., Deming D., Desert J., Langton J., Laughlin G., Lewis N. K., Showman A. P., 2011, ApJ, 727, 23 Bruntt H., Bedding T. R., Quirion P.-O., Lo Curto G., Carrier F., Smalley B., Dall T. H., Arentoft T., Bazot M., Butler R. P., 2010, MNRAS, 405, 1907 Bruntt H., De Cat P., Aerts C., 2008, A&A, 478, 487 18 P.F.L. Maxted et al. Bruntt H., Deleuil M., Fridlund M., Alonso R., Bouchy F., Hatzes A., Mayor M., Moutou C., Queloz D., 2010, A&A, 519, A51 Burrows A., Budaj J., Hubeny I., 2008, ApJ, 678, 1436 Carter J. A., Winn J. N., 2009, ApJ, 704, 51 Charbonneau D., Allen L. E., Megeath S. T., Torres G., Alonso R., Brown T. M., Gilliland R. L., Latham D. W., Mandushev G., O'Donovan F. T., Sozzetti A., 2005, ApJ, 626, 523 Cooper C. S., Showman A. P., 2005, ApJ, 629, L45 Cowan N. B., Agol E., 2011, ApJ, 729, 54 Cowan N. B., Agol E., Charbonneau D., 2007, MNRAS, 379, 641 Cowan N. B., Machalek P., Croll B., Shekhtman L. M., Burrows A., Deming D., Greene T., Hora J. L., 2012, ApJ, 747, 82 Crossfield I. J. M., Hansen B. M. S., Harrington J., Cho J. Y.-K., Deming D., Menou K., Seager S., 2010, ApJ, 723, 1436 de Wit J., Gillon M., Demory B.-O., Seager S., 2012, arxiv:1202.3829 Deming D., Seager S., Richardson L. J., Harrington J., 2005, Nature, 434, 740 Doyle A. P., Smalley B., Maxted P. F. L., Anderson D. R., Collier-Cameron A., Gillon M., et al. 2012, MNRAS, sub- mitted. Enoch B., Collier Cameron A., Parley N. R., Hebb L., 2010, A&A, 516, A33+ Etzel P. B., 1981, in E. B. Carling & Z. Kopal ed., Photo- metric and Spectroscopic Binary Systems A Simple Syn- thesis Method for Solving the Elements of Well-Detached Eclipsing Systems. p. 111 Fabrycky D., Tremaine S., 2007, ApJ, 669, 1298 Fazio G. G., Hora J. L., Willner S. P., Stauffer J. R., Ashby M. L., Wang Z., Tollestrup E. V., Pipher J. L., Forrest W. J., McCreight C. R., Moseley S. H., Hoffmann W. F., Eisenhardt P., Wright E. L., 1998, in A. M. Fowler ed., So- ciety of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 3354 of Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, In- frared array camera (IRAC) for the Space Infrared Tele- scope Facility (SIRTF). pp 1024 -- 1031 Gillon M., Triaud A. H. M. J., Fortney J. J., Demory B.- O., Jehin E., Lendl M., Magain P., Kabath P., Queloz D., Alonso R., Anderson D. R., Collier Cameron A., Fumel A., Hebb L., Hellier C., Lanotte A., Maxted P. F. L., Mowlavi N., Smalley B., 2012, A&A, 542, A4 Gim´enez A., 2006, A&A, 450, 1231 Goodman J., 2009, ApJ, 693, 1645 Gray D. F., 2008, The Observation and Analysis of Stellar Photospheres. Cambridge University Press Harrington J., Hansen B. M., Luszcz S. H., Seager S., Dem- ing D., Menou K., Cho J. Y.-K., Richardson L. J., 2006, Science, 314, 623 Hellier C., Anderson D. R., Cameron A. C., Gillon M., Hebb L., Maxted P. F. L., et al. 2009, Nature, 460, 1098 Huang X., Cumming A., 2012, ApJ, 757, 47 Jehin E., Gillon M., Queloz D., Magain P., Manfroid J., Chantry V., Lendl M., Hutsem´ekers D., Udry S., 2011, The Messenger, 145, 2 Knutson H. A., Charbonneau D., Allen L. E., Fortney J. J., Agol E., Cowan N. B., Showman A. P., Cooper C. S., Megeath S. T., 2007, Nature, 447, 183 Knutson H. A., Charbonneau D., Cowan N. B., Fortney J. J., Showman A. P., Agol E., Henry G. W., Everett M. E., Allen L. E., 2009, ApJ, 690, 822 Knutson H. A., Howard A. W., Isaacson H., 2010, ApJ, 720, 1569 Knutson H. A., Lewis N., Fortney J. J., Burrows A., Show- man A. P., Cowan N. B., Agol E., Aigrain S., Charbon- neau D., Deming D., D´esert J.-M., Henry G. W., Langton J., Laughlin G., 2012, ApJ, 754, 22 Laughlin G., Crismani M., Adams F. C., 2011, ApJ, 729, L7 Madhusudhan N., Harrington J., Stevenson K. B., Nymeyer S., Campo C. J., et al. 2011, Nature, 469, 64 Majeau C., Agol E., Cowan N. B., 2012, ApJ, 747, L20 Mardling R. A., 2007, MNRAS, 382, 1768 Maxted P. F. L., Anderson D. R., Collier Cameron A., Hellier C., Queloz D., Smalley B., Street R. A., Triaud A. H. M. J., West R. G., Gillon M., Lister T. A., Pepe F., Pollacco D., S´egransan D., Smith A. M. S., Udry S., 2011, PASP, 123, 547 Maxted P. F. L., Koen C., Smalley B., 2011, MNRAS, 418, 1039 Nelder J. A., Mead R., 1965, The Computer Journal, 7, 308 Nelson B., Davis W. D., 1972, ApJ, 174, 617 Nymeyer S., Harrington J., Hardy R. A., Stevenson K. B., Campo C. J., et al. 2011, ApJ, 742, 35 Perna R., Heng K., Pont F., 2012, ApJ, 751, 59 Perna R., Menou K., Rauscher E., 2010, ApJ, 724, 313 Pont F., Zucker S., Queloz D., 2006, MNRAS, 373, 231 Popper D. M., Etzel P. B., 1981, AJ, 86, 102 Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P., 1992, Numerical recipes in FORTRAN. The art of scientific computing. Cambridge University Press Raghavan D., McAlister H. A., Henry T. J., Latham D. W., Marcy G. W., Mason B. D., Gies D. R., White R. J., ten Brummelaar T. A., 2010, ApJS, 190, 1 Rauscher E., Menou K., 2012, ApJ, 750, 96 Seager S., Mall´en-Ornelas G., 2003, ApJ, 585, 1038 Seager S., Sasselov D. D., 2000, ApJ, 537, 916 Southworth J., 2009, MNRAS, 394, 272 Southworth J., 2010, MNRAS, 408, 1689 Southworth J., 2011, MNRAS, 417, 2166 Southworth J., Hinse T. C., Dominik M., Glitrup M., Jørgensen U. G., et al. 2009, ApJ, 707, 167 Spiegel D. S., Burrows A., 2010, ApJ, 722, 871 Stetson P. B., 1987, PASP, 99, 191 Stevenson K. B., Harrington J., Fortney J. J., Loredo T. J., Hardy R. A., Nymeyer S., Bowman W. C., Cubillos P., Bowman M. O., Hardin M., 2012, ApJ, 754, 136 Torres G., Andersen J., Gim´enez A., 2010, A&A Rev., 18, 67 Triaud A. H. M. J., Collier Cameron A., Queloz D., An- derson D. R., Gillon M., Hebb L., Hellier C., Loeillet B., Maxted P. F. L., Mayor M., Pepe F., Pollacco D., S´egransan D., Smalley B., Udry S., West R. G., Wheatley P. J., 2010, A&A, 524, A25+ van Leeuwen F., 2007, A&A, 474, 653 c(cid:13) 2008 RAS, MNRAS 000, 1 -- 17
1708.02669
1
1708
2017-08-08T22:44:32
A differential Least Squares Deconvolution method for high precision spectroscopy of stars and exoplanets I. Application to obliquity measurements of HARPS observations of HD189733b
[ "astro-ph.EP" ]
High precision measurements of stellar spectroscopic line profiles and their changes over time contain very valuable information about the physics of the stellar photosphere (stellar activity) and can be used to characterize extrasolar planets via the Rossiter-McLaughlin effect or from reflected light from the planet. In this paper we present a new method for measuring small changes in the mean line profile of a spectrum by performing what we call differential Least Squares Deconvolution (dLSD). The method consists in finding the convolution function (or kernel) required to transform a high signal-to-noise ratio template of the star into each observed spectrum. Compared to similar techniques, the method presented here does not require any assumptions on the template spectrum (eg. no line-list or cross-correlation mask required). We show that our implementation of dLSD is able to perform -at least- as good as other techniques by applying it to star-planet obliquity measurements of exoplanet HD183799 during its transit. Among other things, the method should enable model independent detection of light reflected by an exoplanet.
astro-ph.EP
astro-ph
MNRAS 000, 1–8 (2015) Preprint 10 August 2017 Compiled using MNRAS LATEX style file v3.0 A differential Least Squares Deconvolution method for high precision spectroscopy of stars and exoplanets I. Application to obliquity measurements of HARPS observations of HD189733b John B. P. Strachan,1(cid:63) Guillem Anglada-Escud´e.1 1School of Physics and Astronomy, Queen Mary University of London, 327 Mile End Rd., London, E1 4NS, UK Accepted XXX. Received YYY; in original form ZZZ ABSTRACT High precision measurements of stellar spectroscopic line profiles and their changes over time contain very valuable information about the physics of the stellar photo- sphere (stellar activity) and can be used to characterize extrasolar planets via the Rossiter-McLaughlin effect or from reflected light from the planet. In this paper we present a new method for measuring small changes in the mean line profile of a spectrum by performing what we call differential Least Squares Decon- volution (dLSD). The method consists in finding the convolution function (or kernel) required to transform a high signal-to-noise ratio template of the star into each ob- served spectrum. Compared to similar techniques, the method presented here does not require any assumptions on the template spectrum (eg. no line-list or cross-correlation mask required). We show that our implementation of dLSD is able to perform -at least- as good as other techniques by applying it to star-planet obliquity measurements of exoplanet HD183799 during its transit. Among other things, the method should enable model independent detection of light reflected by an exoplanet. Key words: techniques: spectroscopic – planet-star interactions – stars: rotation 1 INTRODUCTION Since the discovery of the first exoplanet 51 Peg b over 20 years ago using the Doppler method (Mayor & Queloz 1995) over 3500 exoplanets have been discovered 692 of which have been discovered with the Doppler method according to the exoplanet encyclopedia (Schneider et al. 2011). The Doppler method indirectly infers the existence of an exoplanet from the radial velocity (RV) shifts in the spectra of the parent star caused by the reflex motion of the star due to the grav- itational pull on it from the exoplanet. Several high resolution spectrometers used for Doppler velocity measurements such as HARPS (Pepe et al. 2000) operate in the visible wavelength range. More recently spec- trometers such as CARMENES (Quirrenbach et al. 2010) have been built to operate in the near infra-red as well as the visible wavelength range and are in the process of carry- (cid:63) E-mail: [email protected] © 2015 The Authors ing out radial velocity surveys on M dwarf stars in the solar neighbourhood. If an exoplanet happens to transit its parent star then the Rossiter-McLaughlin (RM) effect, (Rossiter (1924) and McLaughlin (1924)) can be observed from high resolution spectra. The RM effect was initially observed in spectra of eclipsing binary stars. As the eclipsing star passed in front of the other rotating star the spectral lines were shifted due to asymmetry in blue shifted or red shifted light of the eclipsed star being blocked. The first RM effect observed for an ex- oplanet was observed over fifteen years ago for HD204958 (Bundy & Marcy (2000) and Queloz et al. (2000)). The RM effect depends on the projected spin orbit mis- alignment angle of the system and the projected rotational velocity of the star. The spin orbit misalignment angle is the angle projected on the sky between the rotation axis of the star and the normal to the orbital plane of the transiting exoplanet. Three different techniques have been used to de- termine these parameters for transiting exoplanet systems using high resolution spectroscopy. 2 John B. P. Strachan, Guillem Anglada-Escude. The first method relies on retrieving these parameters based on the radial velocity measurements e.g., Bundy & Marcy (2000), Queloz et al. (2000) and Triaud et al. (2009). Systematic errors can occur in this method arising from the time-variable asymmetry of the stellar spectral lines during transit and solutions identified can be degenerate (Collier Cameron et al. 2010). Least Squares Deconvolution (LSD) was introduced by Donati et al. (1997) in order to detect magnetic fields in stars using spectropolarimetric observations. Observing the magnetic field signature in the spectropolarimetric observa- tions was not possible to do directly due to the low signal to noise ratio (SNR) of the observation. LSD which involved deconvolving the observation with a template based on an atomic line list enabled the signal in each spectral line to be added and the resulting SNR to be increased by a factor of approximately the square route of the number of lines in an observation. The disadvantages with this technique is that the list of atomic lines in the template has to be complete and their weight have to be known. Collier Cameron et al. (2010) introduced the use of the Cross Correlation Function (CCF) in order to track the shadow of the transiting exoplanet HD189733a as it passed across its parent star. The CCF is produced as described in Baranne et al. (1996) and Pepe et al. (2002) using a tem- plate for the spectral type of the star which has a set of box-shaped emission lines corresponding to the lines in the spectrum of the star. The template is then correlated with the spectrum for the star resulting in the CCF which is a single high SNR spectral line which can be fitted to a Gaus- sian. As the exoplanet passes in front of the star the bump corresponding to the light blocked by the exoplanet can be seen moving across the CCF. An alternative method using CCFs described in Cegla et al. (2016) has the advantage that it does not assume a particular Gaussian function for the CCF due to the subtraction of out and in-transit CCFs. Again disadvantages with these techniques include ensur- ing all lines are correctly identified in the template and also catering for blended lines both of which are problems for late type stars. In order to cater for the limitations here of the above techniques a new method called differential Least Squares Deconvolution (dLSD) has been developed. Instead of build- ing templates by identifying lines from atomic linelists or from the star itself this method uses a high SNR template which is just a combination of spectra from the star. The method consists in finding the convolution kernel that needs to match the template to the observation in a least squares sense. As the light blocked by the exoplanet as it passes in front of the star is to first order the same as the spectra of the star though, Doppler shifted and inverted, the Kernel function will contain a sharply peaked function representing the blocked light. The Kernel function can then be fitted to a forward model. The fitting procedure consists of 1) generating the ex- pected spectrum using the template and a parameterized physical model, 2) applying the deconvolution procedure to generate a synthetic kernel, and 3) apply the same deconvo- lution procedure to the observed spectrum to generate the kernel of the observation and 4) compare the two using a Bayesian procedure. The parameterized physical model in- cludes among other things a limb darkening function, planet- Kr(vi) ≈ n j=1 star radius ratio on the obliquity of the orbit and the impact parameter (distance in plane of sky of closest approach of the exoplanet to the centre of the parent star). In this paper, we will only assume two free parameters (projected spin-orbit misalignment angle λ and projected stellar rotational veloc- ity veqsini) for simplicity because these are the only ones that cannot be determined with photometry. We describe in Section 2 the details of the dLSD algo- rithm and the forward model used to retrieve the projected spin orbit misalignment angle and projected rotational ve- locity parameters for a given system. In Section 3 we report on the performance of the dLSD algorithm based on test data and then for HARPS observations for HD189733. Fi- nally the conclusions for the paper are given in Section 4. 2 DESCRIPTION OF THE DLSD ALGORITHM High resolution spectra from systems such as HARPS are first processed using the HARPS-TERRA software (Anglada-Escud´e & Butler 2012) and the option to remove the blaze from the spectra is selected. The HARPS-TERRA software creates a high signal to noise (SNR) template spec- trum T and provides the observation spectra O which have been velocity shifted to take account of the earth's rotation and also the measured Doppler shift of the star. Each spectrum is itself composed of a number of diffrac- tion orders N. Thus for instance T = {T1, T2, .Tr, .., TN} where Tr represents the rth diffraction order. The observations and template are interpolated using bicubic splines so that they have the same sampling and are moved to velocity space where vi represents the velocity of the ith sampled element. The spectra are processed order by order. For the rth order the residuals Rr to be fitted using least squares are obtained by subtracting the convolved spectrum Cr from the observation: Rr(vi) = Or(vi) − Cr(vi), where the convolved spectrum is the convolution of our ker- nel with the template: (1) Cr(vi) = Kr ∗ Tr(vi). (2) In order to obtain Kr a deconvolution has to be per- formed. Firstly we approximate Kr to be a linear combina- tion of n basis functions θ j representing the signal due to the transiting exoplanet plus a Dirac delta function which represents the signal from the star: αj θ j(vi) + δ(vi), (3) where αj are free parameters to be fitted. Top hat functions are used for the basis functions which have the property θ j(vi) is 1 when i=j and 0 otherwise. We define a set of basis spectrum functions bj(vi) to be the convolution of the jth basis function with the template: s k=1 bj(vi) = θ j ∗ Tr(vi) = θ j(vk)Tr(vi − vk), (4) MNRAS 000, 1–8 (2015) where s is the number of velocity elements in the template. The residuals can now be expressed in terms of the ob- servation and template spectra and the spectrum basis func- tions as: Rr(vi) = Or(vi) − Tr(vi) − n j=0 αj bj(vi). (5) Normally the number of velocity elements in the tem- plate (s) is significantly greater than the number of elements in the kernel (n) and we have an over-determined system which we fit using the least squares χ2 method: χ2(α0, ..., αn−1) = R(vi)2 σ2 i , (6) where σi is the error/weight for each velocity element. The weight we use assuming Poisson statistics is: s i=1 σi =(cid:112)Tr(vi). The template value is used as opposed to the observa- tion value due to the template not containing outliers from noise and thus producing more reliable weight estimates. To minimise the χ2 we differentiate it with respect to each αj and set the resulting terms equal to zero. Rearrang- ing the resulting equations we get a set of n simultaneous equations which in matrix form is: AKr = u, (8) where Kr = {α1, α2, ..., αn} are our kernel coefficients and where each element in the n-column vector u is: s s i=1 i=1 u[m] = bm(vi)Rr(vi) , σ2 i and each element in the n×n matrix A is: A[m][j] = bm(vi)bj(vi) σ2 i , (7) (9) (10) Deconvolution is an example of a Fredholm integral equation of the first kind and the solution to these equa- tions are well known to often be ill-conditioned (Groetsch 2007). In addition our matrix A will be rank-deficient as we would expect that the majority of the elements in K will be effectively 0 (within the noise level). In order to deal with these issues we use Tikhonov regularisation (Tikhonov (1977)) where the solution to the set of simultaneous equa- tions is taken to be the minimisation of the following func- tional composed of an accuracy term and a penalty term: M(u, κ) = in fKr ∈F{AKr − u2 + κIKr2}, where I is the identify matrix, F is the domain of Kr and κ is a free parameter commonly called the Tikhonov pa- rameter expressing the relative weight of the penalty term to the accuracy term. The Tikhonov parameter has to be carefully chosen as picking a value too large will result in (11) MNRAS 000, 1–8 (2015) N r=1 Differential Least Squares Deconvolution 3 over-smoothing of the solution and if its too small then the noise will end up swamping the signal There are several methods which can be used to se- lect the value of the Tikhonov parameter including: Dis- crepancy principle, L-Curve criterion, General Cross Val- idation (GCV) and Normalized Cumulative Periodogram (NCP)(Hansen (2010). An implementation of these methods is available in Regtools (Hansen 1994) and were tested. The NCP was selected as it worked well and did not have the dis- advantages inherent in the other methods. The Discrepancy Method required manual selection of a safety parameter and as such is not ideal to use for automated software. The choice of this parameter was also sensitive to the resulting value determined for the Tikhonov parameter. The L-Curve cri- terion resulted in a dramatic over-smoothed solution due to the cross-over between dominance of the accuracy term and penalty term not being sharp. The GCV suffered from a known issue of sometimes under-smoothing the solution. Once solutions for the kernel for all N differential orders of the residuals have been determined we then combine these kernels using a simple mean: K(vi) = Kr(vi) N . (12) An additional detail to the algorithm was that a mask M(vi) was used to identify velocity elements vi which were to be included in the calculation of the χ2 statistic (having value 1) or not (having value 0). The mask was initialised so all its elements had value 1. After calculating the residual spectrum R(vi) in equation (1) clipping is performed to a user configurable level. All clipped velocity elements have their corresponding mask entry set to 0. In order to avoid problems with telluric contamination a telluric mask TM(vi) was used and Doppler shifted to move it to the same frame as the observations. Any velocity elements vi coinciding with the position of telluric lines in the mask had their entry in mask M(vi) set to 0. 2.1 Forward model Our aim is to determine the projected spin-orbit misalign- ment angle of the exoplanet orbit and the projected rota- tional velocity of the star. In order to do this these two parameters are free parameters in our forward model. The forward model uses these free parameters along with a num- ber of system parameters derived from photometry to derive a model kernel. This model kernel can then be compared to the kernel obtained from the in transit observations in order to confirm the validity of the values of the free parame- ters used. The rest of this section describes how the forward model enables us to produce the model kernel. In order to keep the forward model as simple as possible our model does not cover the time when the exoplanet is partially transit- ing the star - we only model from the time of second point of contact to the time of third point of contact during the transit. We assume the star follows a standard quadratic limb darking law: I(µ) = I(0)(1 − 1(1 − µ) − 2(1 − µ)2), (13) 4 John B. P. Strachan, Guillem Anglada-Escude. where 1 and 2 are the limb darkening coefficients and µ is the cosine of the angle between the direction of the cen- tre of the star to the observer and the centre of the star to the transiting exoplanet. 1 and 2 are fixed value input parameters to our forward model. Following Hartman et al. (2015) we take the spectrum of the star out of transit to be: SROOT(v) = S ∗ G(v), where S is the spectrum of the star which has not been broadened due to the rotation of the star and G is the rota- tionally broadening profile of the star. (14) During the transit the spectrum of the star is given by: (15) SRIT(v) = S ∗ (G − D)(v) where D represents the light (which is not rotationally broadened) blocked by the planet. Assuming quadratic limb darkening the analytical functions for G and D are given in the appendix of Hartman et al. (2015) and are not re- peated here except to say that G = G(u,1,2) and D = D(u,1,2, xp, yp) where u is the relative velocity shift the broadening kernel is measured at and xp, yp is the current position of the planet in the plane of the sky. The projected spin-orbit misalignment angle λ, the an- gle between the angular momentum vector of the exoplanet and the axis of rotation vector of the star and the projected rotational velocity of the star veqsinI where veq is the equa- torial velocity of the star and I is the inclination of the spin axis with respect to the observer are the two free parameters in our forward model. Here we specify two forward models to determine xp, yp and u from the free parameters specified in the previous paragraph and a number of fixed parameters derived from observables. For the first model we assume that we do not rely on the orbital parameters for the system. This can occur when only have old or partial radial velocity data for the orbit. In this case we rely on parameters derived from photometry including the second and third times of contact t2 and t3, the ratio of the exoplanet radius to that of the star RP/R∗ and impact parameter b. The impact parameter is the minimum distance in the plane of the sky from the exoplanet to the centre of the star in units of stellar radius. For a transiting exoplanet b has a value between 0 and 1. For this case we assume circular motion of the exoplanet is valid and the exoplanet moves across the star in a straight line at a constant velocity vplanet . The position of the exo- planet in the plane of the sky at any point in time is given by coordinate pair (xp, yp) as detailed in Figure 1 Both of the axis are scaled so that R∗ = 1. Given the assumptions above and values for the pro- jected spin-orbit misalignment angle λ and impact parame- ter b the path of the planet across the star is on the line with yp = b. The distance between the second and third points of contact d in units of stellar radius (see Figure 2) is given by: (cid:113)(1 − RP/R∗)2 − b2. d = 2 (16) Figure 1. Planet during transit in the plane of the sky showing the spin-orbit misalignment angle λ and impact parameter b. The path of the planet is also shown along with the axis used to locate the position (xp, yp ) of the planet. Figure 2. Plane of sky showing position of planet transiting at second and third points of contact and the distance d between them. Providing we know the times of second and third point of contact t2 and t3 from observations then we have that the velocity of the planet is: d t3 − t2 = vplanet = 2(cid:112)(1 − RP/R∗)2 − b2 xp(t) = vplanet(t − t2) −(cid:113)(1 − RP/R∗)2 − b2. planet at time t is: t3 − t2 . (17) (18) Thus the projected on the sky x-coordinate for the The shortest distance from the spin axis to the planet MNRAS 000, 1–8 (2015) sp again scaled in units of R∗ expressed in terms of xp and yp is using simple geometry from Figure 1 : sp(t) = xpcosλ − ypsinλ. The Doppler shift velocity u is: (19) u(t) = spveqsinI, where veq is the observed equatorial rotation velocity of the planet and I is the angle the rotation axis is inclined with respect to the observer. (20) For the case of the forward model if the orbit of the planet is known then the on sky coordinates are given by: xp = rsin(ν + ω − π/2), (21) yp = rcos(ν + ω − π/2)cosi, where ν is the true anomaly, ω is the argument of pericentre and i is the inclination of the orbit to the line of sight of the observer. We can then use equations 19 and 20 to calculate the Doppler shift u. (22) For the purposes of the forward model we determine the non-rotationally broadened spectrum of the star S from equation 14 by deconvolving the high SNR template T we have for the star with the rotational broadening profile G: SROOT,r(v) = Tr = Sr ∗ G(v), where the suffixes r are present as we perform the deconvo- lution order by order. (23) Having determined Sr we can then determine SRIT,r (v) from equation 15 by performing a convolution for each order r: SRIT,r(v) = Sr ∗ (G − D)(v), (24) We can now determine the forward model kernel pro- file Mr as specified in equation 11 using dLSD taking the template Tr in equation 1 as Sr and the observations Or in equation 2 as SRIT,r (v). 2.2 Bayesian model for parameter selection Bayes equation is used to determine the posterior probability distributions P(θD) of the free parameters θ in the forward model based on the data D: P(Dθ)P(θ) P(D) ∝ P(Dθ)P(θ), P(θD) = (25) where P(Dθ) is the likelihood of the data given the param- eters θ, P(θ) is the prior probability distribution of the free parameters and P(D) is the normalization constant which we do not calculate here as only the value proportional to P(θD) is required to be determined. Running dLSD on the spectroscopic data and the for- ward model synthetic data results in two files to be com- pared. We have the processed data file (D) and the forward model file (M) which depends on the free parameters θ. Both MNRAS 000, 1–8 (2015) Differential Least Squares Deconvolution 5 files have the same format with each row containing a Kernel k calculated for one of the spectra. Each row contains several elements e which contains the value of the Kernel function at the velocity denoted by the element. Assuming Gaussian distributions for the uncertainties in the measurements the likelihood function L = P(Dθ) is given by: N n k=1 e=1 (cid:113) L = (cid:16)− 1 2 1 2π(σ2 k ) exp + σ2 ke (D[k, e] − M[k, e])2 σ2 k + σ2 ke , (26) (cid:17) where N is the number of kernels, n is number of elements in a kernel, σk and σke are the errors in the kernel and the ker- nel element respectively, and D[k,e] and M[k,e] correspond to the value of the eth element in the kth kernel for the data file and the model file respectively. In order to avoid correlations between kernel element the kernels were sampled at the res- olution of the spectroscope which for HARPS is 2.5kms−1. We tested that the noise in the kernel was Gaussian using test generated data samples with the Smirnof-Kolmogorov test. We measured the noise for the kernel elements using the standard deviation of the kernel elements in the wings of the kernel where there is no signal. There are two parameters in the model: the observed rotation velocity of the star veqsinI and the projected spin orbit misalignment angle λ. We use flat priors for both of these parameters: P(veqsinI) = where veqsinI ∈ {(veqsinI)min, ..., (veqsinI)max}. (veqsinI)max − (veqsinI)min , 1 (27) 1 4π , whereλ ∈ {−2π, ..., 2π}, P(λ) = (28) and where (veqsinI)min and (veqsinI)max represents the ex- tremes in the range of values for veqsinI and depends on the star. In practice logarithms of values are calculated numeri- cally so from Bayes equation the posterior probability of the free parameters θ becomes: lnP(θD) ∝ lnL + lnP(θ), where logarithm of the likelihood lnL is: lnL = nNln(cid:0) 1√ (cid:1) + N k=1 2π (cid:33) n − N e=1 ln (cid:32) 1(cid:113)(σ2 n k 1 2 k=1 e=1 ) + σ2 ke (D[k, e] − M[k, e])2 (σ2 k + σ2 ke ) (29) . (30) Markov Chain Monte Carlo (MCMC) is used to calcu- late the posterior probabilities. The chain is started using an initial set of free parameter values θ0 and lnLP(θ0) is calculated. A trial set of parameter is then calculated: θi+1 = θi + step ∗ N(0, 1), (31) John B. P. Strachan, Guillem Anglada-Escude. 6 where N (0,1) is a Gaussian distribution with mean 0 and variance 1. The step number is adjusted so that 20% of the trial set of parameters are accepted. The trial set parameters are accepted if: lnLP(θi+1) − lnLP(θi) > 0, and if this condition is not met then the trial set of param- eters can also be accepted if: (32) Random[0, 1] < exp[lnLP(θi+1) − lnLP(θi)]. (33) The chains are normally allowed a burn in of 100 ac- cepted proposals and are then allowed to run for 1000 ac- cepted proposals. 3 PERFORMANCE 3.1 Testing by injecting simulated data Testing of the algorithm was carried out using out of tran- sit data from HARPS high resolution spectroscopic data for HD189733. A high SNR template was created from the spec- tra captured that night and then the algorithm dLSD was run against seven out of transit spectra. The resulting ker- nels are shown in Figure 3 showing a signal comprising of black and white vertical bends around the 0 velocity point of the kernel and which eventually disappears away from it. As flux is conserved in convolution operations the scale for the values of the kernel are such that if we had a kernel with all elements 0 bar one element with value -0.1 then this would correspond to the signal caused by the transiting planet having exactly the same spectrum as that of the star out of transit but with -1/10th of the amplitude. This signal could be due to a number or reasons includ- ing correction for the blaze not being accurate enough and high frequency components in the spectra causing the band- ing due to aliasing from the deconvolution. As the signal is constant across the different Kernels it can be removed by averaging and then subtracting the average from each Kernel. The same seven spectra were used and then injected with a signal to simulate the effect of the planet transiting the star which took the form of adding a copy of the tem- plate spectrum whose amplitude was multiplied by a factor of -0.05 and whose velocity was shifted from -6 to +6 kms−1 in steps of 2 kms−1. The kernels for the seven spectra are shown in Figure 4 and show clearly the synthetic signal for the transiting planet - the black diagonal line moving from left to right. In addition there is ringing. The ringing has maximum amplitude of approximately 16% of the ampli- tude of the main signal. The ringing is due to the Gibbs ef- fect (Wilbraham (1848)) which is well known to be present with Tikhonov regularisation where the kernel function is discontinuous or has sharp edges. In the above the seven observation spectra were not normalised which will be the case for non-synthetic data. Normalizing these spectra gives the results for the kernel in Figure 5 where at mid-transit the signal as weaker and is stronger towards the start and end of the transit. Figure 3. Greyscale of kernel function for seven out of transit spectra for HD189733 showing vertical striped bands near to the central velocities of the kernel. Figure 4. Greyscale of kernel functions for seven out of transit spectra for HD189733 with injected signal from -6 to +6 kms−1 in steps of 2kms−1. Black diagonal line shows the presence of the signal along with alternative white and black banding caused by Gibbs effect. Figure 5. Greyscale of kernel function for seven out of transit spectra for HD189733 with injected signal from -6 to +6 kms−1 in steps of 2kms−1 which has been normalised. 3.2 HD189733b HD189733b is a hot Jupiter transiting its parent star (Bouchy et al. (2005)) and is one of the most studied ex- oplanets. High resolution spectra of HD189733 were ob- tained using HARPS (High Accuracy Radial velocity Planet Searcher) at the 3.6 metre telescope in La Silla, Chile over four nights: July 30th, August 4th and September 8th 2006, and August 29th 2007 under the allocated programme 079.C-0828(A). These spectra are publically available and were obtained by us from the ESO archive. The spectra were processed using HARPS-TERRA and then the dLSD software and the kernel for the spectra for the night of August 4th 2006 are shown in Figure 6. We used the simple forward model not relying on all the orbital parameters to determine the projected spin orbit MNRAS 000, 1–8 (2015) Differential Least Squares Deconvolution 7 Figure 6. Greyscale of kernel function for the 20 spectra of HD189733 on the night of 9th August 2006. The lines in blue represent veq sinI determined for the star and the red dots repre- sent the four points of contact during the transit. Table 1. Parameters used from previous studies of HD189733b. Parameter Value Units Reference TC T4 - T1 T2 - T1 b Rp /R∗ 2453988.80339 1.827 24.6 0.687 0.1581 BJD Triaud et al. (2009) hours Winn et al. (2007) mins Winn et al. (2007) R∗ Triaud et al. (2009) Triaud et al. (2009) misalignment angle and rotation of the star. This was due to the radial velocities (RVs) available from HARPS not having good coverage of the complete orbit as the data came mainly from around the transit times. HARPS-TERRA was used to derive the RVs from the HARPS data and a straight line was fitted to the out of transit measurements in order to determine the RVs of the star in transit. The values we used for all the fixed parameters are de- tailed in Table 1 with references to the literature where the values came from. There were two main sources for the data. The paper from Triaud et al. (2009) who derived the param- eters from both spectroscopic and photometric data. This paper had all the fixed parameters required for the forward model bar times for the second third and contacts. In order to obtain these times we took photometrically derived data from Winn et al. (2007). In order to fit the model kernel to the kernel derived from the observations with our free parameters we use Markov Chain Monte Carlo (MCMC) with the Metropolis Hastings algorithm (Metropolis et al. (1953)) The standard χ2 statistic was used to measure the fit. We allowed a burn of 100 accepted proposals before capturing a chain of 1000 accepted proposals. A correlation diagram for the probabil- ity distributions of the two free parameters λ and veqsinI is shown in Figure 7. This figure shows no significant correla- tion between the parameters. We produced chains for three different fixed values of the limb darkening parameter 1 = 0.7, 0.8 and 0.9 with 2 held at 0 and the results are presented in Table 2. The results show a small amount of correlation between the limb dark- ening parameter and the projected spin orbit misalignment angle. We also calculated quadratic limb darkening parame- MNRAS 000, 1–8 (2015) Figure 7. Correlation diagram showing the probability distri- butions for the parameters λ and veq sinI for the transit of 8th September 2008 of HD189733. Table 2. Mean and one sigma errors for the fitted parameters for the 7th September 2006 HD189733 data for differing values of the limb darkening parameters  . Parameter Value Error units 1 = 0.7, 2 = 0.0 λ veq sinI 1 = 0.8, 2 = 0.0 λ veq sinI 2 = 0.9, 2 = 0.0 λ veq sinI LDTK: 1 = 0.7676 , 2 = 0.0028 λ veq sinI -0.374 2.988 0.475 0.015 degrees kms−1 -0.108 2.998 0.376 0.040 degrees kms−1 0.007 3.021 0.416 0.043 degrees kms−1 -0.069 2.977 0.437 0.037 degrees kms−1 ters using the Limb Darkening Toolkit (LDTK) (Parviainen & Aigrain 2015) which uses the PHOENIX synthetic atmo- spheres and stellar spectra library (Husser et al. 2013). As inputs to the LDTK we used the HD189733 stellar parame- ters: effective temperature 4875 ± 43K, logg 4.56 ±0.03 and metallicity z = -0.03 ± 0.08 from Boyajian et al. (2015). The results of running dLSD against these limb darkening parameters are given in Table 2. We also ran Markov chains where instead of processing the spectra produced from the forward model with dLSD and comparing it with the dLSD output from the observa- tion spectra we just compared the observation spectra di- rectly with spectra generated from the forward model using a least squares fit. The Markov chains did not converge due to the lack of signal to noise in the RM effect using this method. 8 John B. P. Strachan, Guillem Anglada-Escude. 4 DISCUSSION AND CONCLUSIONS Donati J.-F., Semel M., Carter B. D., Rees D. E., Collier Cameron We have developed and implemented the dLSD technique. The spin-orbit misalignment we calculated is within one σ of perfect alignment and is not far from previous published results such as the work of Triaud et al. (2009) who cal- culated λ = 0.87+0.32 −0.28 degrees and Cegla et al. (2016) who calculated λ = −0.4 ± 0.2 degrees. The projected rotational velocity is in broad agreement with previous results whose values range from 2.9 to 3.5 kms−1. The rotation period of the star carried out photometri- cally by Henry & Winn (2008) is 11.953± 0.009 days which if we assume the rotation angle is aligned with us gives sinI is 1 and using an average of the three values we obtained for veqsinI gives us a radius of 0.714R(cid:12). This value is signif- icantly smaller than the values calculated by Triaud et al. (2009) of R∗=0.766±+0.007 −0.013R(cid:12) and of the R∗=0.805±0.016R(cid:12) value using interferometry from Boyajian et al. (2015). How- ever the discrepancy may be due to the differential rotation of the star as explained in Collier Cameron et al. (2010). One current drawback with the dLSD approach is the amount of computer time to perform the calculation in the forward model. For the analysis of HD189733 to calculate one step in the MCMC chain takes 5 minutes on a laptop with a 2.3GHz I7 processor. Given that only 20% of the steps are accepted the time taken to perform all the accepted steps including the burn in is just over 5 days. However there is significant room to de- crease the time of the individual steps by introducing paral- lel processing into the algorithm so that each of the spectral orders could be processed separately. In future work the authors expect to be able to draw quantitative comparisons between the different methods CCFs, LSD and dLSD discussed here for different types of stars. In conclusion we have shown that we have been able to build the dLSD algorithm which can successfully be used to detect the path of an exoplanet as it transits its par- ent star using high resolution spectroscopy. In particular we have shown that the only template we need to use is one which is just built directly from the spectra of the star itself and from this we have been able to estimate the spin orbit misalignment angle for the HD189733 system. ACKNOWLEDGEMENTS J.S. acknowledges support of studentship funded by Queen Mary University of London. This research was also sup- ported by STFC Consolidated Grant ST/P000592/1. REFERENCES Anglada-Escud´e G., Butler R. P., 2012, ApJS, 200, 15 Baranne A., et al., 1996, A&AS, 119, 373 Bouchy F., et al., 2005, A&A, 444, L15 Boyajian T., et al., 2015, MNRAS, 447, 846 Bundy K. A., Marcy G. W., 2000, PASP, 112, 1421 Cegla H. M., Lovis C., Bourrier V., Beeck B., Watson C. A., Pepe F., 2016, A&A, 588, A127 Collier Cameron A., Bruce V. A., Miller G. R. M., Triaud A. H. M. J., Queloz D., 2010, MNRAS, 403, 151 A., 1997, MNRAS, 291, 658 Groetsch C., 2007, J. Phys.: Conf. Ser., 73 Hansen P. C., 1994, Numerical Algorithms Hansen P., 2010, Discrete inverse problems : insight and algo- rithms. SIAM Hartman J. D., et al., 2015, AJ, 150, 197 Henry G. W., Winn J. N., 2008, AJ, 135, 68 Husser T.-O., Wende-von Berg S., Dreizler S., Homeier D., Rein- ers A., Barman T., Hauschildt P. H., 2013, A&A, 553, A6 Mayor M., Queloz D., 1995, Nature, 378, 355 McLaughlin D. B., 1924, Popular Astronomy, 32, 558 Metropolis N., Rosenbluth A. W., Rosenbluth M. N., Teller A. H., Teller E., 1953, J. Chem. Phys., 21, 1087 Parviainen H., Aigrain S., 2015, MNRAS, 453, 3821 Pepe F., et al., 2000, in Iye M., Moorwood A. F., eds, Proc. SPIEVol. 4008, Optical and IR Telescope Instrumen- tation and Detectors. pp 582–592 Pepe F., Mayor M., Galland F., Naef D., Queloz D., Santos N. C., Udry S., Burnet M., 2002, A&A, 388, 632 Queloz D., Eggenberger A., Mayor M., Perrier C., Beuzit J. L., Naef D., Sivan J. P., Udry S., 2000, A&A, 359, L13 Quirrenbach A., et al., 2010, in Ground-based and Air- Instrumentation for Astronomy III. p. 773513, borne doi:10.1117/12.857777 Rossiter R. A., 1924, ApJ, 60 Schneider J., Dedieu C., Le Sidaner P., Savalle R., Zolotukhin I., 2011, A&A, 532, A79 Tikhonov A. N., 1977, Solutions of ill-posed problems. John Wiley and Sons Triaud A. H. M. J., et al., 2009, A&A, 506, 377 Wilbraham H., 1848, The Cambridge and Dublin Mathematical Journal, 3, 198 Winn J. N., et al., 2007, AJ, 133, 1828 This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1–8 (2015)
1909.02031
1
1909
2019-09-04T18:09:45
Probing planet formation and disk substructures in the inner disk of Herbig Ae stars with CO rovibrational emission
[ "astro-ph.EP", "astro-ph.GA", "astro-ph.SR" ]
[abridged]CO rovibrational lines are efficient probes of warm molecular gas and can give unique insights into the inner 10 AU of proto-planetary disks. Recent studies have found a relation between the ratio of lines originating from the second and first vibrationally excited state, denoted as $v2/v1$, and the emitting radius of CO. In disks around Herbig Ae stars the vibrational excitation is low when CO lines come from close to the star, and high when lines only probe gas at large radii (more than 5 AU). We aim to find explanations for the observed trends between CO vibrational ratio, emitting radii, and NIR excess, and identify their implications in terms of the physical and chemical structure of inner disks around Herbig stars. Slab models and full disk thermo chemical models are calculated. Simulated observations from the models are directly compared to the data. Broad CO lines with low vibrational ratios are best explained by a warm (400-1300 K) inner disk surface with gas-to-dust ratios below 1000; no CO is detected within/at the inner dust rim, due to dissociation at high temperatures. In contrast, explaining the narrow lines with high vibrational ratios requires an inner cavity of a least 5 AU in both dust and gas, followed by a cool (100-300 K) molecular gas reservoir with gas-to-dust ratios greater than 10000 at the cavity wall. In all cases the CO gas must be close to thermalization with the dust. The high gas-to-dust ratios needed to explain high $v2/v1$ in narrow CO lines for a subset of group I disks can naturally be interpreted as due to the dust traps that have been proposed to explain millimeter dust cavities. The broad lines seen in most group II objects indicate a very flat disk in addition to inner disk substructures within 10 AU that can be related to the substructures recently observed with ALMA.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. Herbig_rovib September 6, 2019 c(cid:13)ESO 2019 Probing planet formation and disk substructures in the inner disk of Herbig Ae stars with CO rovibrational emission Arthur D. Bosman1, Andrea Banzatti2, 3, Simon Bruderer4, Alexander G. G. M. Tielens1, Geoffrey A. Blake5, and Ewine F. van Dishoeck1, 4 9 1 0 2 p e S 4 . ] P E h p - o r t s a [ 1 v 1 3 0 2 0 . 9 0 9 1 : v i X r a 1 Leiden Observatory, Leiden University, PO Box 9513, 2300 RA Leiden, The Netherlands e-mail: [email protected], 2 Department of Physics, Texas State University, 749 N Comanche Street, San Marcos, TX 78666, USA 3 Department of Planetary Sciences, University of Arizona, 1629 East University Boulevard, Tucson, AZ 85721, USA 4 Max-Planck-Institut für Extraterrestrische Physik, Giessenbachstrasse 1, 85748 Garching, Germany 5 Division of Geological & Planetary Sciences, California Institute of Technology, 1200 E California Blvd, Pasadena, CA 91125 September 6, 2019 ABSTRACT Context. CO rovibrational lines are efficient probes of warm molecular gas and can give unique insights into the inner 10 AU of proto-planetary disks, effectively complementing ALMA observations. Recent studies have found a relation between the ratio of lines originating from the second and first vibrationally excited state, denoted as v2/v1, and the Keplerian velocity or emitting radius of CO. Counterintuitively, in disks around Herbig Ae stars the vibrational excitation is low when CO lines come from close to the star, and high when lines only probe gas at large radii (more than 5 AU). The v2/v1 ratio is also counterintuitively anti-correlated with the near-IR (NIR) excess, which probes hot/warm dust in the inner disk. Aims. We aim to find explanations for the observed trends between CO vibrational ratio, emitting radii, and NIR excess, and identify their implications in terms of the physical and chemical structure of inner disks around Herbig stars. Methods. First, slab model explorations in LTE and non-LTE are used to identify the essential parameter space regions that can produce the observed CO emission. Second, we explore a grid of thermo-chemical models using the DALI code, varying gas-to-dust ratio and inner disk radius. Line flux, line ratios and emitting radii are extracted from the simulated lines in the same way as the observations and directly compared to the data. Results. Broad CO lines with low vibrational ratios are best explained by a warm (400 -- 1300 K) inner disk surface with gas-to-dust ratios below 1000 (NCO < 1018 cm−2); no CO is detected within/at the inner dust rim, due to dissociation at high temperatures. In contrast, explaining the narrow lines with high vibrational ratios requires an inner cavity of a least 5 AU in both dust and gas, followed by a cool (100 -- 300 K) molecular gas reservoir with gas-to-dust ratios greater than 10000 (NCO > 1018 cm−2) at the cavity wall. In all cases the CO gas must be close to thermalization with the dust (Tgas ∼ Tdust). Conclusions. The high gas-to-dust ratios needed to explain high v2/v1 in narrow CO lines for a subset of group I disks can naturally be interpreted as due to the dust traps that have been proposed to explain millimeter dust cavities. The dust trap and the low gas surface density inside the cavity are consistent with the presence of one or more massive planets. The difference between group I disks with low and high NIR excess can be explained by gap opening mechanisms that do or do not create an efficient dust trap, respectively. The broad lines seen in most group II objects indicate a very flat disk in addition to inner disk substructures within 10 AU that can be related to the substructures recently observed with ALMA. We provide simulated ELT-METIS images to directly test these scenarios in the future. Key words. protoplanetary disks -- molecular processes -- astrochemistry -- radiative transfer -- line: formation 1. Introduction Planetary systems are thought to be built within proto-planetary disks of gas and dust around young stars. How these disks tran- sition from the initial gas-rich remnants of star formation to the solid-body dominated debris disks and planetary systems is still an open question. While for most disks it seems that they go through a quick dispersal process (Cieza et al. 2007; Currie & Kenyon 2009), there is a subset of disks that goes through a pro- longed period where dust and gas are (partially) depleted in the inner disk, but where there is a large reservoir of mass at larger radii, the so called transition disks (Maaskant et al. 2013; Garufi et al. 2017; van der Marel et al. 2018). It is thought that in these disks the cavity is formed either through giant planet formation or X-ray/UV photoevaporation (for reviews, see Owen 2016; Er- colano & Pascucci 2017). These processes cause different distri- butions of gas and dust in the inner disk. To distinguish these scenarios, we have to study the inner disk. CO rovibrational lines are good tracers of the inner disk, because they are strong lines that originate only from warm, dense gas. The upper level energies for the first vibrationally ex- cited level are around 3000 K so they will only be excited in en- vironments with high temperatures ((cid:38) 300 K). Furthermore CO is a very stable molecule and is thus expected to survive even in regions where there is little dust to shield the gas from UV pho- tons (e.g. Bruderer 2013). Finally the transitions are strong so small columns of excited CO are needed to produce bright lines making CO columns as low as 1016 cm−2 easily detectable if the excitation conditions are right. Article number, page 1 of 25 A&A proofs: manuscript no. Herbig_rovib Fig. 1. Disk structures as proposed by Banzatti et al. (2018) (left) and CO vibrational ratio v2/v1 and emitting radius from near-infrared CO spectra of Herbig stars used for comparison to the models in this work (right), see details in Section 2). The three groups from Banzatti et al. (2018), are shown in different colors: group II in magenta, high-NIR group I in green, low-NIR group I in blue. Disks where FNIR is not available are marked in black. We mark two regions that will be used for comparison with models: disks that have CO inside 5 AU, which show a low vibrational ratio (group II and high-NIR group I, bottom left corner), and disks that have CO only outside of 5 AU, which show high vibrational ratios (low-NIR group I, top right corner) . Table 1. Median values for Herbig groups from Banzatti et al. (2018), plus notes from imaging studies Herbig group Group II Group I (high-NIR) Group I (low-NIR) FNIR 0.16 0.27 0.08 v2/v1 RCO (AU) 0.12 0.05 0.27 3 5 18 log(Fe/H) Dust structures from imaging Refs no large inner cavities, some substructures < 5 AU (1) 40 -- 100 AU cavities, (misaligned) inner disks < 10 AU (2) 15 -- 50 AU cavities, no significant inner disks (3) -4.4 -4.6 -5.2 Notes. References: (1) e.g. Menu et al. (2015); Huang et al. (2018); Isella et al. (2018) ; (2) e.g. Pinilla et al. (2018); Stolker et al. (2016); Avenhaus et al. (2017); Tang et al. (2017); di Folco et al. (2009); Boehler et al. (2018); Benisty et al. (2017); (3) van der Plas et al. (2017); Fedele et al. (2017); White et al. (2018); Pinilla et al. (2018) Observing the fundamental CO lines around 4.7 µm allows for the simultaneous measurement of rovibrational line fluxes from the first (v1) and second (v2) excited states. The v2/v1 line flux ratio carries information on the excitation conditions of the gas. The high resolving power on spectrographs such as Keck- NIRSPEC (McLean et al. 1998), VLT-CRIRES (Kaeufl et al. 2004), IRTF-CSHELL and now iSHELL (Rayner et al. 2016) make velocity-resolved observations of CO line profiles possible (e.g. Najita et al. 2003; Blake & Boogert 2004; Thi et al. 2005; Brittain et al. 2007; Pontoppidan et al. 2008; Salyk et al. 2011; Brown et al. 2013; Banzatti & Pontoppidan 2015; Brittain et al. 2018). As the emission is expected to come from a Keplerian disk, the width of the line, once coupled with disk inclination and stellar mass, can be used to estimate the CO emitting radii for different gas velocities, thereby obtaining information on the spatial distribution of CO gas in inner disks. For disks around T-Tauri stars, CO rovibrational lines have been used to probe molecular gas within inner disk dust cavi- ties (Pontoppidan et al. 2008, 2011a), and to propose an inside- out clearing scenario for gas and dust (Banzatti & Pontoppidan 2015). In their sample of T-Tauri disks, CO shows a lower vi- brational temperature with decreasing linewidth (and hence in- creasing emitting radius). This fits well with the expectation that the gas temperature decreases as a function of distance from the star. Furthermore the variation in CO emitting radii can be ex- plained by varying inner molecular gas disk radii, indicating that the inner edge of the molecular disk is not set by the sublimation of dust but carved by another process such as planet formation or photoevaporation. Herbig disks, instead, behave very differently and show an inverse relation between linewidth and vibrational excitation (Banzatti & Pontoppidan 2015). This was attributed to UV- fluorescence becoming more important for stars with higher con- tinuum UV fluxes when the thermal excitation becomes less effi- cient at larger disk radii (Brittain et al. 2007; Brown et al. 2013; Banzatti & Pontoppidan 2015). However, full thermo-chemical models suggest that UV fluorescence is not the dominant excita- tion mechanism for the v = 1 and v = 2 levels of CO (Thi et al. 2013; Hein Bertelsen et al. 2014). This is supported by the ob- served rovibrational excitation diagrams that only show a strong difference between rotational and vibrational temperatures for levels v = 3 and higher, indicating that only for the higher vi- brational levels, UV pumping is important (van der Plas et al. 2015). Furthermore, based on their SEDs, Herbigs are divided into two groups. Disks with strong far-infrared emission relative to their mid-infrared emission are classified as group I or "flared" disks. Disks with strong mid-infrared emission in comparison to their far infrared emission are classified as group II or "settled" disks (Meeus et al. 2001). An evolutionary sequence between these groups was inferred with group I disk being the precursors of group II disks (Dullemond & Dominik 2004). However it is Article number, page 2 of 25 inner rim\inner edgeinner disk (outer) radiusdisk gapgap (outer) radiusouter disksolarFe/Hcavityinner edge/ cavity radiuscavity wallouter disk surfacesub-solarFe/HCO rovibnear-IRinner rim\inner edgesolarFe/Hgroup IIhigh-NIR group Ilow-NIR group Iouter diskinner disk surface100101RCO (AU)102101100v2/v1FNIR0.16group IIFNIR>0.22 group IFNIR<0.12 group IR = 5 AUv2/v1=0.2 Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. 2. Selection of stacked CO line profiles from observed spectra (Section 2). The v1 lines are shown in black, v2 lines in blue. Gaps visible in some line profiles are due to telluric absorption. Disk inclinations are between 20 and 50 deg for all these objects. In HD 31648, the RCO is taken for the broad component defined by the line wings. the group I disk that are often observed to have a large (> 10 AU) cavity in either scattered light or sub-mm imaging, implying that they are unlikely the precursors for the group II disks that are generally less massive and do not show any cavity (Maaskant et al. 2013; Garufi et al. 2017). Inner disk ((cid:46) 5 AU) tracers of both gas and dust add interest- ing pieces to this puzzle. Table 1 reports median values for three inner disk tracers (near-infrared excess, FNIR; CO vibrational ra- tio, v2/v1; CO emitting radius, RCO) as well as the median stellar surface Fe abundance (Kama et al. 2015) used to identify three groups of Herbig disks in Banzatti et al. (2018); here we also add notes on the presence of disk structures from imaging stud- ies. Group II disks exhibit a narrow range of intermediate values for the near-infrared excess, FNIR = L1.2−4.5µm/L(cid:63) (Garufi et al. 2017). All of the group II disks show broad CO 4.7 µm rovi- brational lines indicating that CO is emitting from small radii. These tracers together indicate that both molecular gas and dust are present and abundant at small distances from the star ((cid:46) 5 AU; van der Plas et al. 2015; Banzatti et al. 2018). The group I disks, instead, remarkably split into two very distinct groups (Banzatti et al. 2018). Some of them have very high near-infrared excesses (high-NIR), higher than the group II sources, and only moderately broad CO rovibrational lines. The rest of the group I disks have low near-infrared excesses (low- NIR) and the narrowest CO rovibrational lines. Group I disks thus, while all having dust cavities imaged at larger radii, seem to show a marked dichotomy in their inner disks between those that have abundant gas and dust in the inner few AU, and those that do not, without a gradient of situations in between. While these groups do not show any segregation in terms of mass accretion rates (Banzatti et al. 2018), stellar elemental abundances show that the low-NIR group I disks are depleted in Fe compared to all of the other sources (Table 1), suggesting that the stars in low- NIR group I disks accrete gas that is depleted in dust compared to the 100:1 ISM dust ratio, suggesting that dust is trapped at larger radii in the disk (Kama et al. 2015). In this work we focus on CO rovibrational emission, and in particular the observed trends between the radius and exci- tation of CO emission and the NIR excess (Fig. 1), to expand our growing understanding of inner disk structure and evolution in Herbigs. Specifically, we aim to explain the dichotomy be- tween low vibrational ratios coming from gas within < 5 AU and the high vibrational ratios coming from larger radii. The observational dataset from Banzatti et al. (2017, 2018), briefly presented in Sec. 2, is used for comparison and validation of the models. In Sec. 3 the vibrational excitation of CO will be studied through simple slab models. Full thermo-chemical models using Dust And LInes (DALI, Bruderer et al. 2012; Bruderer 2013) for different physical structures will be presented and analysed in Sec. 4. The implications will be discussed in Sec. 5 and our conclusion will be summarized in Sec. 6. 2. Data overview The CO emission lines adopted in this work for comparison to the models are taken from the compilation included in Ban- zatti et al. (2017, 2018), based on spectra originally presented in Pontoppidan et al. (2011b); Brown et al. (2012); Banzatti et al. (2015); van der Plas et al. (2015); Banzatti et al. (2018). The data consist of high resolution (R ∼75,000 -- 100,000) spectra of CO rovibrational emission around 4.7 µm for 20 Herbig Ae stars and 3 F stars, taken with the CRIRES instrument on the Very Large Telescope (VLT) of the European Southern Observatory (ESO; Kaeufl et al. 2004) and iSHELL on the NASA Infrared Telescope Facility (IRTF; Rayner et al. 2016). The spectrum of HD 142666 is taken from a previous survey (Blake & Boogert 2004; Salyk et al. 2011) done with Keck-NIRSPEC (R ∼ 25,000; McLean et al. 1998). The two parameters we focus on in this work, the CO vibrational ratio, v2/v1, and a characteristic emit- ting radius, are measured from stacked line profiles as explained in Banzatti & Pontoppidan (2015). In brief, the vibrational ratio v2/v1 is measured from the line flux ratio between lines around the v2 P(4) line (v(cid:48) = 2, J(cid:48) = 3 → v(cid:48)(cid:48) = 1, J(cid:48)(cid:48) = 4) and around the v1 P(10) line (v(cid:48) = 1, J(cid:48) = 9 → v(cid:48)(cid:48) = 0, J(cid:48)(cid:48) = 10). The choice of these specific lines is driven by the spectral coverage of the observations, and by the need to use unblended lines (see details in Banzatti & Pontoppidan 2015). The vibrational flux ratio between the v2 P(4) and v1 P(10) line is used as a proxy for the vibrational ratio between the v2 and v1 levels. The vibrational ratio depends on the lines that are used in the comparison, even lines of matching J level show a variation of up to 50% in the vibrational ratio. The v2 P(4) and v1 P(10) line ratio lies within the range of values obtained by using match- ing J levels and is thus a good proxy for the vibrational ratio (see Appendix A in Banzatti & Pontoppidan 2015). A characteristic emitting radius is estimated from the half width at half maxi- mum (HWHM) of the line profile, assuming Keplerian rotation and using literature values for the disk inclination and the stellar mass. As better measurements of disk inclinations have become available over time for some disks, estimates of CO radii have changed accordingly; the error-bars in Fig. 1 reflect the uncer- tainties in the disk inclinations. Figure 1 shows these parameters and their trend as discussed above, namely that the vibrational ratio is larger when CO emission comes from larger disk radii. Figure 2 shows a selection of CO line profiles, chosen to span the full range of CO emitting radii and vibrational ratios for the Article number, page 3 of 25 A&A proofs: manuscript no. Herbig_rovib three groups of disks in Fig. 1. The broader lines (i.e. smaller CO emitting radii) have low v2/v1, and can have flat or double- peaked line profiles. HD 31648 is the only exception that clearly shows two velocity components, as commonly found for T-Tauri stars (Bast et al. 2011; Banzatti & Pontoppidan 2015). This com- bination of broad wings and strong peak indicates that the emit- ting area of the CO rovibrational lines spans a large range of radii (see more in Section 4). In this analysis, for HD 31648 we take the CO radius as indicated by the broad component, defined by the broad line wings. The narrower lines (i.e. larger CO emit- ting radii) often show a single peak profile indicative of a more extended emitting area, but in some cases they clearly show a double peak profile, indicative of an emitting region that is con- fined to a narrower ring. In addition, we use CO line fluxes as measured in Banzatti et al. (2017), which we scale to a common distance of 150 pc for comparison with the model. The near-infrared excess is mea- sured between 1.2 and 4.5 µm (Garufi et al. 2017; Banzatti et al. 2018). Table 1 shows the median values for these parameters for the three groups of Herbig disks, as reported in Banzatti et al. (2018). Spectra and individual measurements can be found in the original references reported in this section. 3. Slab modelling of the vibrational ratio To be able to infer the physical conditions of the CO emitting re- gions we have to look at the CO line formation process. To com- pute the strength of a line one needs to know both the chemical and physical state of the gas. Physics and chemistry are strongly intertwined with the temperature, density and radiation field in- fluencing the chemistry and the chemical abundances influenc- ing the heating and cooling of the gas, changing the temperature. Various thermo-chemical models have been developed it solve this coupled problem (e.g. Woitke et al. 2009; Bruderer et al. 2012; Bruderer 2013; Du & Bergin 2014), however, before we dive into the full problem, we will first study the line formation of CO in a more controlled setting. The line formation of CO will be studied using two different types of slab models. First the behaviour of a slab of CO with fixed excitation will be studied analytically; this will reveal the effects of the optical depth and excitation on the vibrational ra- tios. Afterwards RADEX models (van der Tak et al. 2007) will be used to study non-LTE effects. These RADEX models will be used to constrain the physical conditions of the CO rovibrational emitting regions. 3.1. Analytical line ratios 3.1.1. Methods In the case of a mono-thermal slab of CO that is in LTE with an excitation temperature Tex, the continuum subtracted peak sur- face brightness can be computed by: I(u, l) =(cid:0)B(cid:0)(Tex, ν(u,l) (cid:1) − B(cid:0)Tback, ν(u,l) , (1) (cid:1)(cid:1) ×(cid:16) 1 − exp−(cid:104) hν(u,l) 1 − e−τ(u,l)(cid:17) (cid:105)(cid:17) (cid:16) kTex g(l) exp where τ(u, l) = g(u) g(l) c2A(u, l) 8π2ν2(u, l) NCO √ 2πσvZ(Tex) (cid:35) . (cid:34) − E(l) kTex (2) In Eq. 1 I(u, l) is the continuum subtracted line peak intensity, B(T, ν) is the Planck function at temperature T and frequency ν, Article number, page 4 of 25 Fig. 3. CO vibrational ratio, v2/v1, for different temperatures and columns from the analytic model. The green line shows the τ = 1 con- ditions for the v1 line. The white line shows v2/v1 = 0.2, which is the value that differentiates low and high vibrational ratio sources. Tback is the radiation temperature of the background and τ(u, l) is the line peak opacity. In Eq. 2 g(n) is the degeneracy of rovibra- tional level n, A(u, l) is the Einstein A coefficient of the transition between rovibrational levels u and l, NCO is the CO column, σv is the thermal linewidth, Z(Tex) is the rovibrational partition func- tion of CO, E(n) is the energy above the rovibrational ground state of state n and c, h and k are the speed of light, the Planck constant and the Boltzmann constant as usual. The v1 P(10) and the v2 P(4) lines are used as proxy for the stacked v1 and v2 line from the observations (Section 2). Under the current assumptions the peak line intensity only depends on the excitation temperature, the total column and the background radiation temperature. This last parameter drops out when look- ing at line ratios (assuming that Tback does not vary significantly over the frequency range). 3.1.2. Results The peak line intensity ratio for the v1 and v2 lines are shown in Fig. 3 for a range of temperatures and CO columns. At columns smaller than 1017 cm−2 both lines are optically thin and as such there is no trend with column in the line ratio. At high column densities the line ratio converges to g2(u)A2(u, l)/g1(u)A1(u, l) which is ∼ 1.01 for the lines under consideration, for almost any temperature (T > 200 K). The green line shows where the v1 line becomes optically thick. An increase in CO column to the right of this line no longer elicits a linear response in the line flux. As the v2 flux still in- creases linearly with the column, this increases the line ratio. If the column gets big enough the v2 line also gets optically thick and the line ratio tends to unity. The speed at which this happens with increasing column strongly depends on the population of the upper level of the v2 transition. At temperatures above, 2000 K the column at which the v1 becomes optically thick increases due to the lower fractional population in the lower rotational lev- els of the v1 line. Observed line ratios coming from within 5 AU are generally below 0.2, so from Fig. 3 the conditions to match these obser- vations can be easily read off. If both lines are optically thin, an excitation temperature less than ∼2000 K induces low line ratios. At columns above 1017 cm−2 a line ratio of 0.2 requires 10141016101810201022CO column (cm2)102103104Temperature (K)Optically thickOptically thin0.21031021011v2/v1 Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission lower temperature with increasing column, to values as low as ∼190 K at a CO column of 1022 cm−2. All disks with RCO > 5 AU have line ratios between 0.2 and 0.5. For these high line ratios Fig. 3 shows that, as expected a high line ratio can be due to high temperature, or large columns. For low columns, temperatures between 2000 and 6000 K are needed to produce the right line ratios. Above a column of 1017 cm−2 progressively lower temperatures lead to the observed line flux ratios. 3.1.3. Discussion The vibrational ratio from these lines can be expressed as a vi- brational excitation temperature. However, this only represents the excitation of the gas if both lines are optically thin. CO rovi- brational lines get optically thick at CO columns of 1016 -- 1018 cm−2 (Fig. 3). Assuming a dust mass opacity of 2 × 103 cm2 g−1 (Bruderer et al. 2015, small grains), and a gas-to-dust ra- tio of 100 gives a H column of ∼ 1022 cm−2 before the dust gets optically thick at 4.7 µm. This allows for CO columns up to 1018 cm−2 that can be detected for a canonical CO abundance of 10−4, and thus the generation of optically thick lines above the dust photosphere. Higher CO columns are possible if grains have grown beyond 1 µm or if the dust is depleted with respect to the gas in the emitting layer. Analysis of 13CO lines suggest that columns of 1019 cm−2 are not uncommon for the sources with a high vibrational ratio (van der Plas et al. 2015). For these columns the high v2/v1 ratios can be explained with a tempera- ture between 300 and 500 K. Equation (1) does not include the absorption of the line by dust grains nor the emission of hot dust in the region were τdust < 1. These contributions would lower the line flux, by ab- sorbing line photons and increasing the continuum level. These effects are more pronounced at low line opacities, and so affect the v2 line stronger than the v1 lines. As such, the line ratios are overestimated. In the case of an added dust contribution, the dust opacity sets a maximum to the CO column that can be seen, while the dust emission sets the background temperature. The CO column under consideration is thus only the CO column above the dust photosphere (τdust,4.7µm (cid:46) 1). One critical assumption of this analysis is that both lines are formed in the same region of the disk, either under one set of conditions or under a range of conditions each of which gives rise to a similar line ratio as the region average. The idea being that if the v1 and v2 lines would be coming from different regions of the disk, this would be seen as a significantly different line shape. As the line ratios are determined on the broad component that is seen in both the v2 and v1 lines we can be sure that this assumption holds. 3.2. RADEX models Previously we have assumed a fixed excitation, parametrised by an excitation temperature. Here the excitation processes will be included explicitly by calculating the level populations from the balance between collisions, spontaneous emission and vibra- tional pumping. If no continuum opacity is assumed, the param- eter space is four dimensional: the CO gas column, the kinetic temperature of the gas, the collision partner density (for this pur- pose assumed to be H2 (Yang et al. 2010)1) and the radiation field. For the geometry, a slab that is illuminated from one side is assumed. This configuration is representative for the surface layers of proto-planetary disks, where the infrared continuum photons interacting with the gas are not along the same line of sight as the observations are taken. The pumping radiation inten- sity is parametrized using a 750 K black body diluted by a factor (W) between 0.0001 and 0.3, representative of a region at ∼100 times the radius of the 4.7 µm continuum emitting region and of a region very close to the continuum emitting region. Figure 4 shows the line ratio for the v2 and v1 lines from the RADEX models for different densities. This shows that for columns below 1017 cm−2 line ratios below 0.1 are the norm. Only at high density (> 1014 cm−3) and high temperature (> 1300 K) is the ratio boosted above 0.1, this is similar to the re- sults from the analytic analysis. For W = 0.01, the low density results show the expected subthermal excitation of CO leading to lower line ratios compared to the LTE case. In contrast, in the W = 0.3 case there is a stronger contribution from excitation by infrared photons. This contribution is strongest at low tem- peratures where collisional excitation rates for the vibrational transitions are lowest. 3.3. LTE vs non-LTE The line ratios for CO columns of 1017, 1018 and 1019 cm−2 are plotted in Fig. 5 for both the analytical and RADEX models. For these curves the temperature is assumed to scale as: (cid:32)0.4AU (cid:33)2 T(RCO) = 1500 RCO (3) which is approximately the dust equilibrium temperature around a star of 30 L(cid:12). It is clear that, for these conditions the LTE mod- els can only explain the vibrational ratios at small radii, and those only at columns < 1018 cm−2. The non-LTE RADEX models do somewhat better. With a H2 density of 1012 cm−3, the RADEX models can reproduce the relatively low line ratios at radii smaller than 5 AU at larger columns than the analytical model. The RADEX models can also reproduce the trend in the observed data points in Figure 5, and with a strong enough radiation field, or high enough column, it can also reproduce the absolute line ratios. This indicates that high temperatures or a strongly out of LTE excitation is causing the high vibrational ratio at RCO > 5 AU. Taking into account that the infrared radiation field decreases with radius, Fig. 5 implies that the CO column responsible for the emission needs to increase with RCO. 3.4. Absolute fluxes 3.4.1. Low vibrational ratios in the inner disk To further constrain the conditions of the emitting gas, it is use- ful to compare the absolute fluxes to the observations. First the sources with low vibrational ratios and small CO emitting radii in the lower left corner of Fig. 1 are investigated. Rescaling the Herbig line fluxes from Banzatti et al. (2017) to a common dis- tance of 150 pc leads to a range in fluxes between 4 × 10−15 and 2× 10−12 erg s−1 cm−2 for the v1 line and 3× 10−16 and 4× 10−13 erg s−1 cm−2 for the v2 line. For these sources, the line width implies an emitting radius smaller than 5 AU. Assuming, as a conservative case, that the 1 Results for H as collision partner are similar, but as the collisional rate coefficients are about an order of magnitude larger than the H2 col- lisional rate coefficients, the results for similar densities are shifted to more towards LTE (Song et al. 2015; Walker et al. 2015). Article number, page 5 of 25 A&A proofs: manuscript no. Herbig_rovib Fig. 4. CO vibrational ratio, v2/v1, for different temperatures and columns from the RADEX models using a 750 K radiation field with a dilution factor W of 0.01 (left) and 0.3 (right). The area between the blue and white lines shows where both the vibrational ratio and the v1 flux of the low vibrational ratio sources are reproduced. For the v1 flux an emitting area with a radius of 5 AU is assumed. If a smaller emitting area is assumed the blue lines would shift in the direction of the blue arrows. High vibrational ratio sources can either be explained by gas with a high column (N (cid:38) 1018 cm−2) or a high temperature (T > 2000 K). 3.4.2. High vibrational ratios at larger radii To extract the physical conditions in the emitting regions for the disks with a high vibrational ratio at large radii (low NIR group I disks), the observed fluxes and vibrational ratios were compared with the predicted fluxes from a grid of RADEX models. As the emitting area for these disks is harder to estimate than for sources with a low vibrational ratio at small radii, the emitting area was left as free parameter. As CO is coming from large radii it is ex- pected that the near-infrared radiation field will be weak in the CO emitting area for these sources. Therefore, weaker radiation fields (W = 0.0001 and 0.01) are used in the RADEX modelling. Figure 6 shows the conditions that lead to a vibrational ratio be- tween 0.2 and 0.5 and total integrated v1 fluxes between 3×10−15 and 5× 10−14 erg s−1 cm−2 (normalized to 150 parsec), the range of observed values for low NIR group I sources. Within the RADEX models there are two families of solu- tions. For clarity these solution families have been split in Fig. 6. One solution family is characterised by low temperatures (< 300 K) and very high column densities (> 1018 cm−2), the other so- lution has high temperatures (> 2000 K) and low column densi- ties (< 1014 cm−2). In the low temperature family of solutions, the high line ratio comes primarily from the large columns of gas. The density is virtually unconstrained at small emitting ar- eas and the lowest radiation fields. To allow for a large emitting area, very high densities are needed (> 1013 cm−3), these densi- ties are not reasonable at 10 AU, especially not in the disk sur- face. In the W = 0.0001 case, the radiation field does not domi- nate the excitation. The low temperature branch is also sensitive to the IR continuum radiation field. A stronger IR field moves the solutions to smaller emitting surface areas and higher den- sities as excitation conditions are more easily met and the line surface brightness increases. Strong IR radiation fields are not Fig. 5. CO vibrational ratio versus the inferred radius of emission for observational data (grey points) and analytic (left) and RADEX (right) model results (coloured lines). For the RADEX models, two different assumption for the radiation field are shown weakly irradiated (W = 0.01, dashed) and strongly irradiated (W = 0.3, dotted) cases. For the highest column only the LTE model and the weakly irradiated RADEX model are shown. The density for the RADEX models is 1012 cm−3. flux comes from the full inner 5 AU, it is possible to select con- dition that are able to produce both the correct v1 line flux and the correct line ratio. These conditions are confined between the blue and white lines in Fig. 4. The low vibrational ratios and line fluxes can be repro- duced with CO columns between 1014 -- 1019 cm−2. More con- fined emitting areas would increase the lower limit of the pos- sible columns. Temperatures between 400 and 1300 K are most likely if the CO excitation is dominated by collisions. If IR vi- brational pumping dominates, higher gas temperatures are still consistent with the low vibrational ratios. Article number, page 6 of 25 100200400100020001010cm30.21012cm30.210101013101610191022100200400100020001014cm30.210131016101910221016cm30.21031021011v2/v1Weakly irradiated (W=0.01)CO column (cm2)Gas temperature (K)100200400100020001010cm30.21012cm30.210101013101610191022100200400100020001014cm30.210131016101910221016cm30.21031021011v2/v1Strongly irradiated (W=0.3)CO column (cm2)Gas temperature (K)100101102101100v2/v11019 cm21018 cm21017 cm2Analytic100101RADEXW = 0.010.3W = 0.010.010.30.3CO radius (AU) Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission columns below 1018 cm−2 are needed to explain the low vibra- tional ratios at small RCO. These columns are most likely present in the surface layers of a dust rich inner disk and imply gas-to- dust ratios smaller than 1000, assuming that the dust is optically thick at 4.7 µm. The modest temperatures needed indicate that at these small radii, the temperature of the CO emitting gas cannot be more than a factor ∼ 2 higher than the dust temperature, as gas that is hotter than twice the dust temperature would easily reach 1500 K, especially within the inner AU. This would create higher vibrational ratios than measured. In the next Section, the constraints from the slab models will be used as guidance for the thermo-chemical modelling, and the constrains will be updated with the results from the full disk modelling. To explain the high vibrational ratios coming from large radii a large gap in molecular gas is needed. As these sources also have low near-infrared continuum emission and gaps have been imaged in many of them (e.g. Garufi et al. 2017), a gap devoid of most of the gas and all the dust is assumed. In the case of a large dust gap, the CO column in the cavity needs to be very low, on average lower than 1014 cm−2. If the column were higher, then the v1 flux from within 5 AU would be strong enough to be detected. Alternatively, it can be estimated that the surface area of optically thick CO gas within the cavity needs to be (cid:46) 0.25AU2. Two families of solutions have been found from the RADEX models to fit both the line strengths and the line ratios. The first solution is shown in the left panel in Fig. 7 and needs low tem- peratures and high columns. This solution is preferred as fits of the rotational diagram of rovibrational lines of 12CO and 13CO for disks with a high vibrational ratio (van der Plas et al. 2015) prefer large columns NCO ≈ 1019 cm−2 and moderate tempera- tures (300 -- 500 K). To be able to probe these large columns, local gas-to-dust ratios in the CO emitting regions above 100 are necessary, with many solutions needing gas-to-dust ratios of 10000. The increase in vibrational line ratio with emitting radius seems thus to be an effect of the increase in gas-to-dust ratio of the CO emitting area, with CO coming from gas with a temper- ature that is coupled to the dust for both the low v2/v1 and the high v2/v1 sources. This indicates that the process that clears out the inner disk of gas in the high v2/v1 sources, clears out the dust as well and confines it to larger radii than the gas. This is what would be expected for a dust trap and in line with the low metallicity measured in the accreting material in these sources (Kama et al. 2015). We will discuss these scenarios in Section 5. 4. DALI modelling 4.1. Model setup Armed with an understanding of which conditions reproduce the observations we now run a set of DALI models. Different sets of Herbig disks is modelled with the inner edge, that is, innermost radius at which gas and dust is present (Rin), varied from the classical sublimation radius at 0.4 AU up to 15 AU (see Fig. 8). Table 2 shows the parameters assumed for the model disks. The gas and dust surface densities are given by: Σgas = ∆g−dΣdust (cid:33)−γ (cid:32) R Rc exp − (cid:32) R Rc (cid:33)2−γ, Σdust = Σc 100 (4) where the gas-to-dust ratio is varied between 10 and 10000. The thermo-chemical modelling is done with the code DALI (Brud- erer et al. 2012; Bruderer 2013). The standard setup is used Article number, page 7 of 25 Fig. 6. Parameters that can reproduce the observed CO rovibrational fluxes and line ratios for sources with RCO > 5 AU as function of as- sumed emitting area. Two solution branches are found, a low tempera- ture (left) and a high temperature (right) branch. Different colours show models with different strengths of the infrared radiation field. In the second solution branch, the radiation field does not impact the solutions significantly. expected to be produced by local dust so a local pumping field with W = 0.0001 is preferred over W = 0.01. In the high temperature family the excitation of CO is dom- inated by the collisions with the gas. At these high temperatures both vibrational states are easily populated by collisions and the ratio in which they are populated is similar to the line ratios that is seen. As long as the density is above 1011 cm−3, the result is density independent. Because the lines are optically thin, the line flux is given by the total amount of CO molecules giving rise to a surface area, column degeneracy. The solution is independent of the radiation field assumed. 3.5. Physical conditions in the CO emitting region The slab models provide important constraints on the physical conditions of gas producing the observed CO rovibrational emis- sion. In Fig. 7 these constraints have been put in the context of simple disk geometries. Modest temperatures ((cid:46) 1000 K) and 102103Temperature (K)Solution #1Solution #2101110131015101710191021CO column (cm2)W = 0.01W = 0.0001100101102101110131015Density (cm3)100101102Emitting area (AU2) A&A proofs: manuscript no. Herbig_rovib Fig. 7. Summary of physical condition constraints from the RADEX models on the emitting regions of the CO rovibrational lines. Fig. 14 shows a version of this figure updated with the results of the full disk modelling. The constraints derived here are used to guide the DALI modelling. Regions are not shown to scale. The right panel shows solution #1 from Fig. 6 as observations of 13CO ro-vibrational lines indicate the presence of large columns of CO (van der Plas et al. 2015). Table 2. Fiducial parameters Parameter Stellar Luminosity Stellar Mass Effective Temperature Sublimation radius Critical radius Disk outer radius Gas surface density at Rc Surface density power law slope Disk opening angle Disk flaring angle PAH abun. rel. to ISM Large dust fraction Large dust settling Disk inner radius Gas-to-dust ratio Symbol Rsubl Rc Rout Σc γ hc ψ xPAH, ISM Rin ∆g−d Value 30 L(cid:12) 2.5M(cid:12) 10000 K 0.4 AU 50 AU 500.0 AU 60 g cm−2 1 0.1 0.25 10−20 0.9 0.1 0.4 -- 15 AU 10 -- 10000 have been expanded. For CO five vibrational levels, up to v = 4 each with 41 rotational levels, up to J = 40 are included, with level energies, line positions and Einstein A coefficients taken from the HITRAN database (Rothman et al. 2013). Collision rate coefficients for collisions between CO and H2 (Yang et al. 2010) and H (Song et al. 2015; Walker et al. 2015) are included. The full molecule model is described in Appendix A. The molecule model for H2O has been expanded to include vibrational lines, as these could be important for cooling in the regions that CO is emitting. For H2O the rovibrational datafiles from LAMDA2 are used (Tennyson et al. 2001; Barber et al. 2006; Faure & Josselin 2008). The line profiles are extracted for the CO v2 and v1 tran- sitions using the raytracer as described in Bruderer et al. (2012). For the ray tracing a disk inclination of 45◦ and distance of 150 parsec is used. The extracted line profiles are then convolved to match a re- solving power of R = 100000, and noise is added to achieve a similar signal-to-noise as in the observations (∼ 200). From these line profiles the emitting radius (from the line width) and vibrational line ratio are extracted using the same method as used for observational data by Banzatti & Pontoppidan (2015). For some models the gas temperature and chemistry are not calculated self consistently. These are the LTE models in Fig. 9 and the Tgas = Tdust model in Appendix B. In these models the gas temperature is set equal to the dust temperature as cal- culated by the dust radiative transfer and the CO abundance is parametrised by: xCO = 10−4 × AV 1 + AV , (5) with AV the visual extinction as calculated from the continuum radiative transfer. For large AV the CO abundance converges to the canonical value of 10−4, at AV < 1 the CO abundance is de- creased from the canonical value to mimic the effects of photo- dissociation. The CO abundance globally agrees well with the CO abundance from the thermo-chemical model. These simpli- fied models have been run in LTE conditions and in non-LTE by explicitly calculating the excitation (Tgas = Tdust model). 4.2. Model results 4.2.1. v1 line flux and v2/v1 ratio Results of our fiducial model, with a gas-to-dust ratio of 100, are plotted in Fig. 9 as the black points. The vibrational ratio from 2 Leiden Atomic and Molecular DAtabase http://home.strw. leidenuniv.nl/~moldata/ (Schöier et al. 2005) Fig. 8. Schematic representation of the surface density in the monolithic models. Rin is the same for gas and dust and is varied between 0.4 and 15 AU, while ∆g−d is varied between 10 and 10000. except for a few changes that will be highlighted where rele- vant. Dust temperature is calculated using Monte Carlo radia- tive transfer. Gas temperature, chemical composition and molec- ular excitation are self-consistently calculated. For the thermo- chemical calculation both the CO and H2O molecular models Article number, page 8 of 25 IR emitting layerτdust, 5μm= 1 inner disk edge< 1 AU< 5 AUno/small cavities; v2/v1 < 0.2 and RCO < 5 AUdisk cavityNCO< 1014cm-2cavity walldominates CO emissionTgas< 300 KNCO> 1018cm-2τdust, 5μm= 1 large cavities; v2/v1 > 0.2 and RCO> 5 AU> 5 AURadius (log)Surface density (log)gd10-10000Radial profileGas (variable)Dust (constant)Rin0.4 - 15 AU"Inner edge" Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission in perature. The gas temperature in the emitting regions decreases slowly with increasing inner model radius and is almost constant for models with Rin between 1.4 and 10 AU. While the total area of the inner edge scales as A ∝ R2+ψ , not all of this area con- tributes to the emission. The emission is dominated by two rings, at the top and bottom of the inner edge wall. These rings are sit- uated in the region where the dust temperatures are higher than the midplane dust temperature and the CO excitation is still ther- malized with the gas. In the models the vertical extend of this region increases slightly with radius, leading to faster than linear growth of the emitting area. Coupled with the constant or slowly declining temperature with radius leads to a roughly linear rela- tion between inner model radius and CO v1 flux. The effect of the gas-to-dust ratio on the vibrational ratio is also shown in Fig. 9 (bottom). All models show a similar trend: with increasing CO radius, the v2/v1 drops. Models with an in- creased gas-to-dust ratio produce higher v2/v1 line ratios. This is due to the larger column of gas that can emit, leading to more optically thick lines, driving up the v2 lines compared to the v1 lines. Even so, the models with the lowest gas-to-dust ratio still have a v2/v1 ratio larger than most of the observed disks for emitting radii of less than ∼ 4 AU. For models with the small- est cavities the line becomes undetectable at gas-to-dust ratios lower than 10. With increasing gas-to-dust ratio, there is also an increasing v1 line flux, indicating that the emitting area is getting larger. The LTE models in Fig. 9 show a very different behaviour to the thermo-chemical models. This is fully due to the LTE as- sumption because the parametrisation of the CO abundances and the assumption of coupled gas and dust temperature have only very small effects of the line ratios (see Appendix B). The LTE models consistently have lower vibrational ratios than the fidu- cial models, due to the fact that neither the IR pumping nor ex- citation due to self-absorption are included. Together these pro- cesses explain the different vibrational ratios between the fidu- cial and LTE models. The LTE models with small cavities have vibrational ratios and CO emitting radii that are consistent with with observations. For the disks with larger cavities, the LTE models cannot come close to the observations, indicating that non-LTE processes are definitely important for gas as large radii. The effect of infrared pumping and UV pumping has also been studied in Appendix B but removing IR pumping and including UV pumping only has marginal effects on the excitation and nei- ther can explain the discrepancy between the data and the mod- els. The LTE models consistently show line fluxes that are in good agreement with the data. That the LTE models seem to do so well, certainly for the low v2/v1 sources, is puzzling. The LTE assumption only holds for the CO rovibrational lines if the local gas density is above ∼ 1016 cm−3. These high densities are only expected near the disk midplane and not at the disk surface. Beyond 5 AU there are non-LTE models that overlap with the data and there is a suggestion that higher gas-to-dust ratios are needed to explain the increase in vibrational ratio with in- creasing CO emitting radius. However, fluxes are high for these models, 10−15 W m−2 at 150 parsec, about a factor of 300 higher than most observed high vibrational line ratio sources. The DALI models do not show a hot, tenuous layer of CO such as would be needed for solution #2 of Fig. 6. They show instead that with high gas-to-dust ratios, and thus high CO columns the right vi- brational ratios can be reproduced, consistent with solution #1 of Fig. 6. Article number, page 9 of 25 Fig. 9. v1 line flux (top) and vibrational ratio of CO (bottom) versus the inferred radius of emission for observational data and DALI model results. Lines connect the dots in order of inner model radius. Labels indicate the gas-to-dust ratio for the thermo-chemical models, the LTE model also has a gas-to-dust ratio of 100. The model with the largest cavity has the largest CO radius. The dust surface density is kept constant for models of different gas- to-dust ratios. Due to missing data, not every source with a vi- brational ratio in the lower panel also has a line flux in the upper panel. Clearly none of these models reproduce the trends in the data. these models shows exactly the opposite trend from the data. The model vibrational ratio is roughly flat with a value around 0.4 for CO radii less than 2 AU, while at larger radii the line ratio decreases. The line-to-continuum ratios and line fluxes for the models are generally too high (Fig. 9, top). At small CO emitting radii line fluxes are consistent with the highest observed fluxes, at large CO emitting radii the model line fluxes are a factor ∼ 10 higher than the average flux. The flux is dominated by optically thick lines coming from the inner edge of the model. The gas temperature in the emitting region is higher than ∼ 600 K in all models. This means that the CO rovibrational lines are emitted at wavelengths longer than peak of the relevant Planck function. As a result, the line flux of these optically thick lines scales linearly with the gas tem- 10181017101610151014v1 line flux (W m2)10102103104LTE100101CO radius (AU)102101100v2/v110102103104LTE A&A proofs: manuscript no. Herbig_rovib 4.2.2. Line profiles As shown in Fig. 9, the extracted line ratios and emitting areas of most models are not able to explain the observed behaviour, especially the low line ratios in the inner disk. It is thus necessary to take a closer look at the predicted line profiles, and compare those with the observed line profiles (Fig. 2) for an explanation for this mismatch. The line profiles for subsets of DALI models are shown in Fig. 10 (line profiles for all models are shown in App. C). The models with small holes (< 2 AU) show a clear dif- ference between the full thermo-chemical models and the LTE models. The full DALI models consistently show a two compo- nent line structure. There is a broad, nearly top hat, component of the line which is present in both the v1 and the v2 lines, and a more strongly peaked line profile that is very weak in the v2 line. This second component compares well to the line profile of HD 31648 in Fig. 2. The total line flux and the line ratio are seen to increase with increasing gas-to-dust ratio. Furthermore, the v1 line profile gets narrower with increasing gas-to-dust ratio, consistent with the emitting area getting larger for higher gas-to-dust ratios. None of the observations show the broad plateau-like fea- ture that is in our model line profiles with small Rin (< 2 AU). This indicates that the inner rim of the model disk needs to be adapted to fit the data. Mostly, the v2 flux from the inner disk wall needs to be strongly reduced. The LTE models show that low vibrational ratios are produced if the gas, dust and CO exci- tation are thermalised. To thermalise the CO excitation densities in the emitting area of more than ∼ 1016 cm−3 are necessary, this is an increase in density of about 4 orders of magnitude com- pared to the current density of the inner disk wall. Another op- tion would be to lower the CO abundance from the inner rim regions by at least 4 orders of magnitude, removing most of the contribution of the inner rim to both the v1 and v2 lines. The line profiles from models with an inner radius of 10 AU generally show a narrow double peaked profile in both lines, in- dicating that the directly irradiated inner edge is contributing most of the flux in both transitions. This is consistent with the very steep line profiles without low-level wings seen from disks with a high vibrational ratio (e.g. IRS 48, Fig. 2). The line ratio strongly depends on the gas-to-dust ratio in the disk surface: high gas-to-dust ratios lead to higher vibrational ratios as the v1 line opacity increases. Higher gas-to-dust ratios also lead to larger v1 fluxes. The LTE models with large cavities have no detectable v2 emission. It is, however the only model v1 line for which the flux is within the observed range; the non-LTE models overpre- dict the flux. Comparison of the line profiles in Fig. 2 and Fig. 10 indicates that for observed disks with low vibrational ratios, the line pro- files can be well reproduced by models that have a small cavity radius except that these models have a plateau-like contribution to the line profile in the inner disk. This indicates that emission from the disk surface agrees with the observed line profiles and line ratios. This is consistent with the analytical and RADEX analysis which predict low vibrational ratios for disk surface conditions. Using the spatial information in the model image cube, the emission was decomposed into a disk surface and a disk in- ner rim component. Fig. 11 shows the original (continuum sub- tracted) and decomposed line profile for the model. The line pro- file cleanly separates into a broad, high line ratio component coming from within 0.63 AU and a narrowly peaked, low line ratio component from the rest of the disk. This suggest that the Article number, page 10 of 25 models strongly overestimate the flux coming from the inner rim. The implications of this will be discussed in Sec. 5.1. 4.3. Disk surface emission The removal of the line contribution from the inner rim makes it possible to make a direct comparison between observations and the flux from the disk surface. We restrict ourselves to model disks with a small inner cavity size (< 1.5 AU) as for these radii the vibrational ratio is most strongly over predicted in the models. The inner rim region from which the line emission is removed originally produces ∼ 40% of the v1 flux and ∼90% of the v2 flux. This region also accounts for ∼90% of the 4.7 µm continuum flux in the model. As before, different inner disk radii and gas-to-dust ratios are studied. On top of that, for mod- els with an inner radius of 0.4 AU and gas-to-dust ratios of 100 and 10000, the outer radius and vertical scale height and flaring are also varied. Table 3 gives an overview of the varied parame- ters and model results. Figure 12 compares results of the DALI models without a contribution of the inner rim to the observed data. By isolating the emission from the disk surface, low vibra- tional ratios can be obtained at small CO radii. Increasing the gas-to-dust ratio increases the vibrational ratio and the v1 line flux, while only slightly increasing the CO emitting radius. In- creasing the inner cavity radius to more than 1 AU causes the CO emitting radius to increase beyond 10 AU for gas-to-dust ratios of 100 and 1000. No sources with such a narrow CO line and a low vibrational ratio are seen. Truncating the outer disk, by removing all material beyond a radius of 8, 5 or 3 AU moves the emission inward and gener- ally increases the vibrational ratio, because the emission has less contribution from larger radii and colder gas. The more truncated disks also have lower v1 fluxes, while the NIR continuum emis- sion is not reduced compared to their full disk counterparts. As expected, a more flared disk has emission from further out, and is vibrationally colder, than a geometrically flatter disk. Lowering the scale height moves the emission further out for a non-flared disk, while for flared disks the emitting radius is reduced. Overall, figure 12 shows that emission from the disk surface, especially with gas-to-dust ratios of 100 or 1000, can match the observed CO line fluxes and vibrational ratios at small radii. Dif- ferent inner radii disk cannot explain the full extent of the data. Restricting the emitting region, in this case by truncating the disk, or changing the vertical structure of the inner disk helps in reproducing the spread in vibrational ratio and CO radius. This indicates that rovibrational CO emission is tracing substructures in the inner disk surface. Comparing the model line profiles (Fig. 10) with the ob- served line profiles (Fig. 2) reveals that there is only one disk that is matched well with a full, flared disk (HD 31648, also known as MWC 480). All other line profiles are better matched with a very flat or even truncated model. The ubiquity of emis- sion at large radii in the models, but not in the data, implies that the inner disk structure of the observed disks is different from the smooth, flared geometry assumed in the model. 4.4. Tgas ≈ Tdust The removal of the inner rim for the small Rin models (Sec. 4.3) and the lower temperatures in the rounded models with large Rin (Appendix E) allow us to reproduce line widths, line strengths and vibrational ratios. All these models have in common that the Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. 10. Normalised model line profiles for the v1 (black) and the v2 (blue) lines for a subset of the models at the native resolution of the model, R = 106. The text on the left of each panel de- notes the model set. The top right corner of each panel denotes the inner radius of the model. The vertical bar in the bottom right of each panel shows 0.03 (top two rows) or 0.3 (bottom row) of the continuum flux density. Two models with a "*" match both RCO and v2/v1 for a subset of the data. All lines are modelled assuming a 45 degree inclination. No noise has been added to these lines, noise-like features in the line pro- files are due to the sampling of the DALI grid. Table 3. Model variations for the models with subtracted edge contributions. Inner radius variation Rin (AU) # 1. # 2. # 3. Outer radius variation 0.4 0.6 1.35 Rout (AU) # 1. # 2. # 3. # 4. # 1. # 2. # 3. # 4. 3 5 8 500 h (rad) 0.02 0.1 0.02 0.1 Flaring variation ψ 0.0 0.0 0.25 0.25 g/d = 100 v2/v1 RCO (AU) 0.03 0.02 0.04 2.3 2.8 15.2 g/d = 100 v2/v1 RCO (AU) 0.18 0.12 0.07 0.03 1.4 2.5 3.2 2.3 g/d = 100 v2/v1 RCO (AU) 0.47 0.18 0.08 0.03 0.8 1.4 12.2 2.3 g/d = 1000 v2/v1 RCO (AU) 0.05 0.04 0.03 2.4 3.2 20.7 a Fv1 9.7 10.4 12.1 a Fv1 4.6 4.9 5.5 a Fv1 2.0 2.6 3.0 4.6 a Fv1 0.2 3.5 0.4 4.6 g/d = 10000 v2/v1 RCO (AU) 0.33 0.32 0.18 3.7 5.4 4.5 g/d = 10000 v2/v1 RCO (AU) 0.32 0.25 0.42 0.33 1.5 2.3 1.1 3.7 g/d = 10000 v2/v1 RCO (AU) 0.35 0.67 0.03 0.33 0.7 0.80 2.5 3.7 a Fv1 20.9 21.8 23.8 a Fv1 7.0 9.0 10.4 20.9 a Fv1 1.4 16.6 5.9 20.9 FNIR 0.14 0.16 0.18 FNIR 0.13 0.13 0.14 0.14 FNIR 0.07 0.45 0.03 0.14 Notes. (a) v1 line flux (×10−14 erg cm−2 s−1) Fig. 11. Line profiles for the models with an inner cavity of 0.6 AU and a gas-to-dust ratio of 10000. In the right hand plot contributions from the inner rim and disk surface are separated. gas and dust temperature in the emitting area are similar, with 20% temperature differences in the surface layers of the disks with small holes and difference below 50% for the inner walls of disks with large cavities. Conversely, models that over predicted the flux or vibrational ratio generally had gas temperatures that were at least twice as high as the dust temperature. These results seem contradictory with results from Bruderer et al. (2012) who modelled the pure rotational high J CO lines in HD 100546, a low-NIR group I source in our sample. Bruderer et al. find that they need a gas temperature that is significantly higher than the dust temperature to explain the v = 0 high J CO rotation diagram. However, the emitting area for the high J and rovibrational CO lines is not the same. The high J lines come from the surface of the outer disk, while the rovibrational CO lines come from the cavity wall. This difference in emitting region is due to the difference in critical density of the transi- tions. The critical density of the CO rovibrational lines is around 1015 cm−3 while the v = 0, J = 32 − 31 transition has a critical density around 107 cm−3. The CO rovibrational lines are thus coming from denser (∼ 1010 cm−3), better thermalised gas than the high J CO lines that can be effectively emitted from the more tenuous, thermally decoupled surface layers. Article number, page 11 of 25 0.00.51.00.4Fiducialv2P(4)v1P(10)0.4g/d = 100.4g/d = 10000.4g/d = 100000.4*LTE0.00.51.02.0Fiducial2.0g/d = 102.0g/d = 10002.0g/d = 100002.0LTE400400.00.51.010.0Fiducial4004010.0g/d = 104004010.0g/d = 10004004010.0*g/d = 100004004010.0LTEVelocity (km/s)Normalized line flux40040Velocity (km/s)0246Flux (Jy)v1P(10)v2P(10)40040Velocity (km/s)Inner rimDisk surface A&A proofs: manuscript no. Herbig_rovib Fig. 12. v1 flux (top) and CO vibrational ra- tio (bottom) versus the inferred radius of emis- sion for observational data and model results. The contribution from the inner edge has been subtracted from the models spectra before anal- ysis. Models show variation in inner radius (left panel), variation in outer radius (middle panel) and variation in flaring and disk height (right panel). In the left and middle panels lines con- nect points in increasing order of the parameter varied. In the right panels, lines connect models with the same flaring angle and thus the differ- ence between connected points show the effect of a change in thickness of the disk. Table 3 lists the parameters varied for these models. Models with a gas-to-dust ratio of 100 and 1000 are bet- ter at reproducing the vibrational ratio than the models with gas-to-dust ratios of 10000. Vari- ation in the vibrational ratio can be reproduced by variations in the disk structure, but no single parameter explain all the variation. The thermo-chemical models by Thi et al. (2013) show a slightly stronger gas-dust temperature decoupling in the inner 10 AU at the CO emitting layer with the gas temperature being 2 -- 3 times higher than the dust temperature. The temperature of the CO emitting layer in Thi et al. (2013) is still within the 400 -- 1300 K range. Further testing will have to be done to see if this hotter layer can also reproduce the low vibrational ratios that are observed. In T-Tauri disks, models of the H2O mid-infrared observa- tions have invoked a decoupling of gas and dust temperatures high in the disk atmosphere to explain Spitzer observations (e.g. Meijerink et al. 2009). In these models this decoupling happens at densities below 109 cm−3, which is lower than the density of the gas that produces most of the CO rovibrational lines of > 1010 cm−3. Our models also show a strong decoupling of gas and dust temperatures (Tgas > 3 × Tdust) in this layer, but no CO rovibrational lines are emitted from there. Observations of optically thin CO ro-vibrational lines, i.e. high J 12CO and CO isotopologue lines, can be used to directly probe the gas temperatures predicted here. The high J 12CO will most likely be more sensitive to the hotter, upper or inner layers of the disk atmosphere and so a higher gas temperature would be inferred from these lines compared to the 13CO and possibly C18O ro-vibrational lines. 5. Discussion Our modelling results show that we can reproduce the observed CO emission with low vibrational ratios at small radii versus high vibrational ratios at large radii (Fig. 1) under different and separate conditions. Low vibrational ratios measured at small disk radii require a CO column below 1018 cm−2 and a tem- perature between 400 -- 1300 K. These conditions naturally occur in the denser > 109 cm−3 surface layers of the disk. Emission from a dust-free inner region, and from an inner disk rim di- rectly irradiated from the star, are ruled out based on the high v2/v1 that would be produced under these conditions which are not observed. Line velocity profiles indicate that most group II disks have an emitting area that is radially narrower than what Article number, page 12 of 25 a flared disk model produces. Thus, flared group II disks should be rare, in agreement with the currently accepted paradigm. High vibrational ratios measured at large disk radii require instead an inner disk region strongly devoid in CO (NCO < 1014 cm−2), i.e. the region that would otherwise produce the high- velocity CO line wings that are not observed in these spectra. In the emitting region at larger radii, where CO is still present, the gas must be cold (< 300 K) and CO columns must be high (> 1020 cm −2). To provide these conditions, a high gas-to-dust ratio is necessary (> 10000) coupled with a density structure that allows for efficient cooling. Models with midplane number density and column density that increase with radius are able to match the line flux, vibrational ratio and RCO. The large columns and low gas temperatures are consistent with 13CO observations (van der Plas et al. 2015). These constraints on the disk physical structure, and their relative models, are summarized in Table. 4. Figure 13 shows four representative CO line profiles, sim- ulated images, and cartoons of the disk structures proposed to produce the observed emission as based on the combination of this and previous analyses. In Sec. 5.1 and Sec. 5.2 we will link these structures to the physical and chemical processes proposed to produce them. Three different disk structures for the low vibrational ratios at small radii are shown in Figure 13. Two of these apply to group II disks. They are divided into structures with a compact (upper left) and extended (lower left) rovibrational CO emitting area. CO line profiles and infrared excess show that abundant gas and dust is present within ∼ 5 AU, so any existing disk cavities must be smaller or the dust extends to the sublimation radius. The stellar abundances for these sources are consistent with solar Fe abundances, so transport of dust to the star is relatively unhin- dered (Kama et al. 2015). The NIR and CO flux are not emitted from the same region: the IR flux mostly comes from the in- ner dust edge that is directly irradiated by the star, while the CO must only be emitted from the disk surface. The major difference between the two group II structures is that those with compact CO emission must have a non-flared geometry confining the CO emitting region, possibly with inner disk substructures shadow- ing the rest of the disk, further confining the emission. The one group II disk with extended CO emission (HD 31648) needs to 10181017101610151014v1 line flux (W m2)123123123Different Rin12341234Different Rout12341234Different verticalstructure100101CO radius (AU)102101100v2/v1123123123Observ.gd = 1000100101CO radius (AU)12341234gd = 100gd = 10000100101CO radius (AU)12341234 Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Table 4. Summary of physical constraints from modelling results. CO emission v2/v1 < 0.2 and RCO < 5 v2/v1 > 0.2 and RCO > 5 Conditions T < 1500 K, NCO < 1018 cm−2 T = 400 − 1300 K, 1014 < NCO < 1018 cm−2 Rounded edge, Tgas < 300 K, NCO > 1020 cm−2 g/d> 10000 No CO in dust free gas g/d(cid:46) 1000 No CO at the inner rim Radially constrained emitting area T (cid:46) 300K, NCO > 1018 cm−2 or T (cid:38) 1000K, NCO < 1014 cm−2 Slab LTE, RADEX Model Slab LTE RADEX DALI DALI DALI Comments Fig. 3 Fig. 4 Sec. 3.5 Fig. 12 Sec. 4.3 and Fig. 12 Sec. 4.3 and Figs. 2, C.1 and 10 Slab LTE, RADEX DALI DALI Figs. 3, 4 and 6 Figs. 9 and E.1 Sec. E Fig. 13. Typical line profiles, simulated images and inferred disk proposed disk structures for four types of disks identified in the Herbig sample. Near-infrared continuum and CO emitting areas are shown in red and blue respectively. The simulated images show the velocity integrated CO v1 line flux. These images are discussed in more detail in Sec. 5.4. The disk structures are updated versions of those shown in Fig. 1. have a more flared geometry or have a slow molecular disk wind, analogous to those seen in T-Tauris (Pontoppidan et al. 2011a; Brown et al. 2013). Based on this sample we conclude that both molecular disk winds, as well as flared group II disks should, be very rare. In the case of a flared disk, the size of the emission is measurable is by spectro-astrometry on 8-meter class or IFU spectroscopy on 30-meter class telescopes. The third structure giving rise to low vibrational ratios at small radii is that of the high-NIR group I disks (upper right). These disks have large cavities (typically the largest found in Herbigs, see Table 1), but they have a residual inner dust disk/belt that produces the high near-infrared flux. These disks therefore have a gap between inner and outer disk. CO emission comes from the inner disk, again from the disk surface in or- der to produce the very low vibrational ratios measured in the data. The high near-infrared flux instead, higher than the group II disks, must have come from a larger emitting area than in the group II disks, possibly from both the inner edge and surface of the inner disk. The solar Fe abundance in the surface layers of the star is an independent indicator of accretion from a still gas- and dust-rich inner disk (Kama et al. 2015), possibly implying efficient filtration of small dust from the outer disk to the inner disk. The high vibrational ratios at large radii can all be explained with a single structure (bottom right in Figure 13). These disks have an inner cavity that is strongly reduced in CO surface den- sity. These inner cavities seem to be on average smaller in size than those imaged in high-NIR group I disks (see Table 1). The rovibrational CO emission comes from the cavity wall, i.e. the inner edge of the outer disk, which must be rich in molecular gas but strongly depleted in dust. This structure combined with the low metallicity of the material that is accreted on the stellar surface (Kama et al. 2015), indicates that the dust is efficiently trapped at some radii larger than RCO for these disks. The most Article number, page 13 of 25 solarFe/HsolarFe/Hsub-solarFe/Houter disksolarFe/Hgroup II disks:no/small cavitiesflat + substructures (common)group II disks:no/small cavitiesflaredgeometry (rare)high-NIRgroup I disks:larger cavities, dust trap?low-NIRgroup I disks:smaller cavities + dust trapplanets?planetsdust trapdust trap?<10 AU40-100 AU15-50 AUCOCOCOCOCO?FNIR(intermediate)FNIR(intermediate)FNIR(high)FNIR(very low)dust trap? A&A proofs: manuscript no. Herbig_rovib appropriate explanation for these deep gas cavities and dust traps currently seems to be that they are caused by giant planets and not by photo-evaporation or dead-zones (see also van der Marel et al. 2016). 5.1. Implications for sources with low v2/v1 at small radii 5.1.1. No CO at the inner rim The good match between the line profiles of disks without a con- tribution from the inner disk edge (Sec. 4.3) indicates that CO is not present within or around the dust sublimation radius in any of these disks. The left plot in Figure 14 shows the proposed structure of the inner disk of a group II source as inferred from the data. The dust disk in this case reaches to the dust sublima- tion radius, forming an inner dust rim. Our modelling suggests that, for Herbig disks, the gas gets heated to high enough tem- peratures (> 3000 K) near the sublimation radius to keep the gas atomic (see App. F for a discussion on the chemistry), at least until the gas is hidden under the dust infrared photosphere. This is not seen in the thermo-chemical models, however, the gas and dust temperature in this region are very uncertain. Dust sublima- tion impacts both the dust temperature structure, as well as the gas temperature by changing the composition of the gas. At larger radii the gas can cool enough to become molecu- lar above the dust infrared photosphere. The expectation is that the phase change from atomic to molecular gas also induces a strong extra cooling effect, lowering the gas temperature to the dust temperature. This seems to be the most appropriate scenario to produce the low vibrational ratios measured at such inner disk radii. This scenario assumes that the dust disk extends all the way to the dust sublimation radius. However, even if there is a small inner hole in the dust disk, the radiation field should still be strong enough to heat the gas to temperatures above 3000 K and keep the gas atomic at the inner edge of the dust cavity, at least within the small inner cavities that have been explored above in Fig.12. The CO abundance structure proposed in our modelling, with the low columns and temperatures and the absence of molecular gas within the inner dust rim, also naturally produces very weak or no CO overtone (∆ν = 2) emission. This is consistent with a large survey of 2 µm CO emission towards Herbig AeBe stars, with detection rates as low as 7% (Ilee et al. 2014). The dissociation of H2 and CO as proposed here, should leave a large (NH (cid:38) 1018 cm−2), hot (T > 3000 K) atomic or ionised reservoir around the dust inner radius. Velocity or spa- tially resolved atomic lines, such as can now be measured with near-IR interferometry (e.g. Eisner et al. 2014; Garcia Lopez et al. 2015, and with VLTI-GRAVITY, Gravity Collaboration et al. (2017)), can thus be used to test if CO and H2 are indeed being dissociated around the inner edge of the disk. 5.1.2. Group II disks have more mass within 5 AU than group I high-NIR sources Both the group II disks and the high-NIR group I disks show low vibrational ratios at small radii, with the latter showing larger radii and lower vibrational ratios on average (Table 1). At the same time these group I disks have a higher NIR excess than the group II disks: this is thought to be due to a vertically more extended dust structure in the inner disk (see e.g. discussions in Maaskant et al. 2013; Banzatti et al. 2018). A vertically more extended structure will lead to lower gas densities in the surface Article number, page 14 of 25 layers of the disk. A large population of small grains is needed to populate the tenuous surface layers and convert stellar flux into the observed bright NIR flux. These conditions naturally lead to larger RCO and lower vibrational ratios. In fact, low densities slow down the chemical formation of CO and observable abun- dances of CO are thus only produced at lower UV fluxes, further from the star. Furthermore, a larger population of small grains has a higher NIR opacity per unit mass of gas, so the visible column of CO is smaller than for group II disks. A lower gas density also helps in lowering the excitation in the v = 2 state (Fig. 4), thereby lowering the vibrational ratio and providing a good explanation for the measured difference from the group II disks. Source specific modelling of the rovibrational lines of CO and its isotopes, fitting the full rovibrational sequence using non- LTE models can be used to further constrain the density in the inner regions of the these disks. Furthermore if the grains are indeed small, and the dust mass in the inner disk is low, the con- tinuum might be optically thin at ALMA wavelengths, allowing for inner disk mass and grain size measurements. The radial extent of CO emission in the high-NIR group I ob- jects is harder to estimate than for the group II objects as the ob- served CO lines are intrinsically narrower. Spectro-astrometric measurements of HD 142527 indicate that the CO emission ex- tends up to ∼ 5 AU, twice the measured emitting radius (Pon- toppidan et al. 2011a), and nicely matching an inner dust belt detected by Avenhaus et al. (2017). A narrow emitting region would imply that CO only emits in the inner dust disk and is ab- sent in the disk gap. None of the high NIR group I sources show the double peak structure expected from such a narrow emitting ring. This could be due to emission from the outer disk cavity wall and surface filling up the centre of the observed spectral line, but higher spectral and spatial resolution observations are needed to confirm this scenario. 5.1.3. Inner disk CO emission is confined by substructures One disk in this sample provides exceptional insight into the in- ner dust and gas structures: HD 142666. Its RCO of ∼ 3 AU is relatively large for a group II object, and a bright ring at ∼ 6 AU has recently been found in sub-millimeter dust continuum im- ages taken at very high angular resolution with ALMA (Andrews et al. 2018; Huang et al. 2018). Figure 15 shows the comparison of the inner part of the sub-millimeter radial continuum intensity profile with the CO radial emission profile. The CO radial emis- sion profile was derived by fitting a flat Keplerian disk intensity model to the observed line profile. The observed line profile and the flat disk fit can be seen in the inset in Fig. 15. The CO rovi- brational emission is confined within the inner edge of the sub- millimeter dust ring, indicating that the process that is producing this sub-millimeter ring also confines the CO rovibrational emis- sion. This could happen if the bright sub-millimeter ring traces a vertically extended dust structure that shadows the disk beyond, preventing CO emission from larger radii. Our fit to the CO line profile for HD 142666 shows that the inner 1 AU of the disk is devoid of emission (Figure 15). This would be consistent with IR interferometric observations that re- port an inner disk radius of 1 AU (Schegerer et al. 2013), signifi- cantly larger than the dust sublimation radius expected at 0.3 AU as based on the stellar luminosity. This indicates that a small cav- ity has formed in this disk; similar cavities could also be present in the other group II disks that have RCO (cid:38) 2Rsubl. These small cavities should still allow for efficient transport of both gas and dust (in a ∼ 100 -- 1 ratio) from the inner disk edge to the star, Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. 14. Sketches of the upper right quartile of a disk cross section of the preferred configuration of the CO emitting region in the case of low RCO and low v2/v1 (group II and group I high NIR, left) and large RCO and high v2/v1 (group I low NIR, right). This figure is an update to Fig. 7 including the DALI results. Relevant radial scales for the inner and outer edge are shown on the bottom left and right corner of each sketch. 5.2. Implications for high v2/v1 at large radii 5.2.1. High gas-to-dust ratios by dust trapping The large CO columns needed to produce a high vibrational ra- tio at large radii indicate that around RCO the gas surface density does not deviate strongly from what would be expected in ab- sence of an inner cavity. To get the v2 lines bright enough, CO columns larger than 1020 cm−2 and thus total H columns larger than 1024 cm−2 are needed. This necessitates a drop in gas sur- face density that is less than two orders of magnitude at RCO if CO is at the canonical abundance of 10−4 alternatively, a CO abundance of more than ∼ 10−6 is needed if the gas surface den- sity is continuous within the dust cavity. The large CO column also implies that dust is under abundant by at least two orders of magnitude at RCO. The high gas-to-dust ratios necessary are likely not due to strong settling as ALMA images of millimeter dust show cavities that are consistently larger than RCO indicat- ing that there is a indeed a radial segregation (e.g. HD 100546, HD 97048, IRS 48, HD169142; van der Marel et al. 2016; van der Plas et al. 2017; Fedele et al. 2017; Pinilla et al. 2018). This is consistent with ALMA gas observations of transition disks, showing that also in the sub-millimeter, CO cavities are smaller than dust cavities (van der Marel et al. 2013, 2016). The steep line profiles and the lack of high velocity CO emis- sion indicates that the CO column within RCO is at least six or- ders of magnitude lower than the column at RCO for the observed disks. As CO is hard to photodissociate, a drop in the total gas surface density in the cavity is necessary. A total gas surface den- sity drop of 5 orders of magnitude between the dust ring and the CO poor cavity is needed to produce CO columns below 1014 cm−2 (Bruderer 2013). Observations of atomic oxygen or car- bon could be used to measure gas depletion factors directly, con- straining the depth of the cavity and thus the mass of a possible cavity-forming planet. Fig. 14, right, illustrates the disk struc- ture as reconstructed from modelling the CO rovibrational lines in low-NIR group I disks. The combination of a large cavity in both gas and dust and of a radial segregation between gas and dust at the disk cav- ity wall fits well with what is expected for a giant planet carv- ing an inner disk hole. A sufficiently massive planet can explain the gas depletion in the cavity together with the different cav- ity sizes in dust and gas, as a gas pressure maximum is created that traps dust at a slightly larger radius. Both photo-evaporation and dead zone models, instead, have problems creating transi- tion disks that have an inner region rich in gas, but depleted in micron-sized grains (Pinilla et al. 2012; Gorti et al. 2015; van der Marel et al. 2015, 2016; Pinilla et al. 2016). Article number, page 15 of 25 Fig. 15. Radial intensity cuts for the sub-millimeter dust from Huang et al. (2018) and the radial intensity as inferred from the CO rovibra- tional line profile of HD 142666. Vertical dashed lines show the maxi- mum of the CO and sub-millimeter dust intensity. The CO emission is clearly contained within the bright sub-millimeter ring at 6 AU. The in- set on the top right shows the observed line profile (black) and the fitted profile (blue). The vertical red line shows the inner edge of the dust disk at ∼ 1 AU as inferred from IR interferometry (Schegerer et al. 2013). as the accretion rates are normal and the stellar abundances for these sources close to solar (Kama et al. 2015; Banzatti et al. 2018). The lower sub-mm intensity implies a lower dust surface density at 1 AU. If we assume there is no continuous build up of material around the sub-millimeter ring, then the lower surface density in the inner disk gap should be accompanied with an in- crease in the velocity of the accretion flow. This would be consis- tent with an inner dead-zone edge, possibly due to the thermal ionisation of alkali metals (Umebayashi & Nakano 1988). An- other option would be the presence of a giant planet within the small cavity with a saturated or leaky dust trap. In either case there would be a dust trap or traffic jam. This raises the tanta- lizing possibility that all group II disks with RCO (cid:38) 1 AU could have a bright sub-millimeter ring in the inner regions as found in HD 142666 and HD 163296 (Huang et al. 2018; Isella et al. 2018). CO emission with v2/v1 < 0.2 and RCO < 5 AUCO thermal dissociationIR emitting layerτdust, 5μm= 1 atomic gasdustsublimationfront< 1 AU< 5 AUCO emission with v2/v1 > 0.2 and RCO> 5 AUdisk cavityCO emitting layerdust depletedatomic gasτdust, 5μm= 1 Dust trap> 5 AUdrop of 10-5in Σ𝑔𝑔𝑔𝑔𝑔𝑔051015Radius (AU)0.000.250.500.751.001.25Normalized CO intensitysmall dust RinCO rovibsub-mm dust0.60.81.01.2Norm. sub-mm intensity40040Vel (km s1)HD 142666 A&A proofs: manuscript no. Herbig_rovib 5.2.2. Not all dust traps are equal In Sec. 5.1 we compared the high NIR group I disks to the group II disks on the basis of inner disk structures. However, at long wavelengths and larger radii these high NIR disks appear more similar to the rest of the group I sources showing cavities in sub- millimeter and scattered light imaging (e.g. Garufi et al. 2017). One notable difference is that the inner disk in the high NIR group I sample seems to be misaligned with respect to the outer disk (e.g. Benisty et al. 2017; Banzatti et al. 2018). Due to this misalignment, perhaps the inner disk cannot shadow the cavity wall in the outer disk and thus CO emission in high NIR group I sources could in principle include a component similar to what observed in low NIR group I disks. If so, there should be a nar- row emission component with large v2/v1 at the center of the line. This could be the origin of the narrower v2 emission line observed in HD 135344B (Fig. 2), but even at the current high spectral resolution it is not possible to distinguish a narrower central peak in the v1 line. Part of the problem may also be due to flux filtration by the narrow slit (0.2" for VLT-CRIRES), where the signal from the outer disk will be diluted by any slit that includes a bright inner disk but excludes part of the outer disk (Hein Bertelsen et al. 2014). Another possibility is that the outer disk in the high NIR group I systems may not have as high a gas-to-dust ratio as the rest of the group I systems. A normal gas-to-dust ratio would quench the v2 emission, with v1 emission from the gap outer edge filling in the line center and producing a narrow single peak. A close to ISM gas-to-dust ratio would imply that dust is not ef- ficiently trapped at the gap outer edge. This could fit with a sce- nario in which the dusty inner disk and near solar abundances in the stellar atmosphere are replenished by dust from the outer disk. Mapping of the CO emission, either with multiple slit po- sitions with VLT/CRIRES+ or observations with ELT/METIS integral field unit can determine the brightness and nature of the CO emission from the outer disk in high NIR group I sources (see Sec. 5.4). 5.3. Comparison to T-Tauri disks: distribution of UV flux matters Under several aspects, disks around Herbig AeBe stars are anal- ogous to T-Tauri disks. However, in CO rovibrational emission they exhibit a very different behaviour (Banzatti & Pontoppidan 2015; Banzatti et al. 2018). While the T-Tauri disks have a de- creasing v2/v1 with increasing RCO, the Herbig disks show the opposite trend. This implies a significant difference in the distri- bution of molecular gas between Herbig and T-Tauri inner disks. This is also seen in the line profiles, since many of the T-Tauri disks have a two component CO profile: if both components originate from the Keplerian disk, it means that the CO emitting region is more radially extended than in the Herbig disks. This can be explained if the T-Tauri disks have CO emission from the inner rim (and from within the inner rim), while in Herbig disks CO is dissociated by high temperatures in these regions as explained above. Both Herbigs and highly accreting T-Tauris have strong UV fields. The energy distribution as function of wavelength is very different, however. The UV field of Herbig stars is dominated by the continuum coming from the stellar surface. For T-Tauri stars, on the other hand, most of the UV comes from the accre- tion shocks and is emitted in emission lines, especially Lyman- α (e.g. France et al. 2014). Both CO and H2 cannot be photo- dissociated by Lyman-α photons. As such the photo-dissociation Article number, page 16 of 25 of CO and H2 is much more efficient around Herbig stars. If hy- drogen is mainly in atomic form, formation of other molecules such as CO2 and H2O is significantly slowed down as molecules both need the OH radical for their formation. This radical forms from H2 + O −−−→ OH + H and can be destroyed by OH + H −−−→ H2 + O. Only with abundant H2 can enough OH be pro- duced and can OH survive long enough to form CO2 and H2O. This could explain the lack of H2O and CO2 emission towards Herbig AeBe disks in comparison to T-Tauri disks (e.g. Pontop- pidan et al. 2010; Banzatti et al. 2017). The higher broad-band optical-UV flux in Herbig systems can have a larger impact on the gas temperature compared to the line dominated T-Tauri spectrum as more power can be ab- sorbed by atomic and molecular electronic transitions before the dust absorbs the radiation. Finally, a larger fraction of the stellar flux can generate photo-electrons upon absorption by the dust (Spaans et al. 1994). All these effects will heat the gas more in Herbig than in T-Tauri disks, increasing CO dissociation in the former. 5.4. Predictions for future observations ELT-METIS will be a generation I instrument on the ELT (Brandl et al. 2014). It will be able to do diffraction limited imag- ing and IFU spectroscopy at 3 -- 5 µm. The 39 meter mirror allows for a spatial resolution of ∼ 0.03" at 4.7 µm in a 0.5" by 1" field of view. IFU spectroscopy will be possible at a resolving power of R = 100000. With these capabilities, ELT-METIS will be able to resolve CO rovibrational emission both spatially and in veloc- ity in nearby Herbig disks. Fig. 16 shows continuum subtracted, velocity integrated maps of the v1P(10) line. The four different disk structures proposed in Fig. 13 can be clearly distinguished in these images. Spatially resolving the emission will enable to study asymmetries in the spatial distribution of CO and will help in explaining the single peaked nature of CO lines observed in the high-NIR group I disks. The CO rovibrational ratio is a good tracer of large cavi- ties in Herbig disks, with high ratios (> 0.2) only coming from low-NIR group I disks, intermediate ratios (0.05 -- 0.2) coming from group II disks and very low ratios (< 0.05) coming from the high-NIR group I disks. This could be exploited in more dis- tant and more massive star forming regions, for instance by ob- serving multiple Herbigs within the field of view of the JWST- NIRSPEC multi object spectrograph, providing an efficient clas- sification of large numbers of Herbig sources, either for more detailed follow-up or for population studies. In the shorter term, ground based, high sensitivity observa- tions can be used to constrain densities in the inner disk. In all disk models, CO emission is not in LTE. This results in strongly decoupled vibrational and rotational excitation temperatures. On top of this, most of the v1 lines are also optically thick so there should be a break in the v1 rotational diagram. This is the point where the lines become optically thin. The position and sharp- ness of the break critically depends on the density of the emitting area. A source by source modelling of the CO rovibrational lines over a large number of J levels (J > 30) can derive this density and from that the mass in the inner few AU can be constrained. The same should apply to T-Tauri disks, but as the CO line flux can have contribution from within the sublimation radius it will not be straightforward to measure the mass in the dust rich inner disk. Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. 16. Simulated velocity integrated v1P(10) (top) and v2P(4) (bottom) line maps convolved to METIS resolution (Brandl et al. 2014). The colour scale is log-stretched between 0.1% and 100 % of the maximum of the v1 line flux. The continuum has been subtracted before velocity integration.The contours show 0.1%, 1% and 10% of the peak surface brightness. The disk geometries refer to those presented in Fig. 13. The distance is assumed to be 150 parsec and the inclination is 45 degrees, the far side of the disk is in the north. For the "flat" geometry, a truncated gas disk (5 AU outer radius) is used. The high NIR group I image is composed by combining two models, a truncated disk model, with an inclination of 30 degrees and a disk with a 40 AU hole and a strong dust trap (so strong v2 emission) with an inclination of 45 degrees. No interactions between the inner and outer disk have been taken into account. 6. Conclusions The goal of this work has been to find the physical conditions that can reproduce trends observed in inner disks of Herbig stars, in terms of the CO vibrational ratio v2/v1, the radius of CO emis- sion (from the HWHM of the lines) and the NIR excess (Fig. 1). We have studied the excitation and line profiles of CO rovibra- tional emission from disks around Herbig Ae stars using LTE and non-LTE slab models, as well as using the thermo-chemical model DALI. Our findings are collected in Figs. 13 and 14 and our conclusions can be summarised as follows: -- Emission from the inner disk surface: CO emission with v2/v1 < 0.2 at RCO < 5 AU is reproduced by conditions found in the inner disk surface. CO columns must be (cid:46) 1018 cm−2 and gas-to-dust ratios < 1000. Gas and dust tempera- tures must be coupled and between 400 and 1300 K. Emis- sion from and within the inner disk rim is ruled out on basis of the measured low vibrational ratios. A scenario in which the gas around the dust sublimation radius is hot, > 3000 K, is preferred to explain the absence of CO at the inner rim. At these temperatures, reactions between CO and atomic H should produce a primarily atomic gas that could be observed by IR interferometry. -- Emission from the cavity wall: CO emission with v2/v1 > 0.2 at RCO > 5 AU is reproduced by conditions found in a cavity wall at large disk radii. CO columns must be > 1018 cm−2 and gas-to-dust ratios > 10000. Gas and dust tempera- tures must be coupled and below 300 K, indicating efficient cooling of the gas. Within RCO the gas surface density drops by at least 5 orders of magnitude. A high gas surface density, rather than UV pumping, is the most likely reason for the bright v2 lines providing high v2/v1 ratios. -- Substructures in inner disks: The broad, flat topped or double peaked line profiles that most group II sources exhibit cannot be explained by a smooth, flared disk. The radial extent of the CO emission in these sources is restricted. Flat disk models work better in matching these line profiles, but they generally still have too much flux at large radii. The outermost radius that emits in these sources thus most likely traces some vari- ation in vertical scale-height. This could be the case for HD 142666, where all the CO emission arises from within the first resolved sub-millimeter dust ring at 6 AU. -- Dust trapping: Small cavities in gas and dust are possible in the group II objects. The low vibrational ratios observed in- dicate that dust-free and molecular gas rich cavities are not present. If there are small cavities formed by planets in the sample, then they apparently do not create a very efficient dust trap. This is in contrast with the low NIR group I disks that have high vibrational ratios. The large gas surface den- sity drop and the dust poor gas necessary in these disks fit very well with predictions of giant planets producing a cavity and a strong dust trap. High NIR group I disks, instead, may have dust traps that allow for dust filtration from the outer to the inner disk, sustaining normal elemental accretion onto the star. -- CO as a tracer of disk cavities and inner dust belts/disks: rovibrational CO emission can be used to identify dust cavi- ties also in absence of direct imaging, especially when disks are further away than 200 pc; the difference in CO lines Article number, page 17 of 25 0.30.10.10.3 (")v1group II "flared"group II "flat"Rgap = 40 AUhigh NIR group IRcav = 15 AUlow NIR group I0.30.10.10.3 (")0.30.10.10.3 (")v20.30.10.10.3 (")0.30.10.10.3 (")0.30.10.10.3 (") A&A proofs: manuscript no. Herbig_rovib as observed in high- and low-NIR group I disks moreover shows that, in disks with large cavities, CO emission is a good tracer for residual inner dust belts/disks that may be unseen in direct imaging. -- Molecular gas within the sublimation radius: The lack of broad, high vibrational ratio, CO emission in many of the ob- served sources puts strong constraints on the amounts of dust free molecular gas within the sublimation radius: NCO < 1018 cm−2. This upper limit is consistent with the non-detection of the CO overtone (v = 2-0) emission towards most Herbig AeBe disks. -- Future METIS observations: ELT-METIS will be able to re- solve the emitting area of the CO rovibrational lines. These observations can further constrain the emitting area in group II disks. For low NIR group I disks METIS should find CO rovibrational rings within the scattered light and sub- millimeter dust cavities, while for high NIR group I disks the extent of molecular gas in the inner disk as well as the gas-to-dust ratio, and thus efficiency of dust trapping, in the outer disk can be measured. Acknowledgements. We thank Antonio Garufi for helpful discussions on imag- ing of Herbig disks, Inga Kamp and Daniel Harsono for help with the CO colli- sional rate coefficients and Paul Molliere for providing the results to the chemical equilibrium calculations. Astrochemistry in Leiden is supported by the Nether- lands Research School for Astronomy (NOVA). This work is partly based on observations obtained with iSHELL under program 2016B049 at the Infrared Telescope Facility, which is operated by the University of Hawaii under contract NNH14CK55B with the National Aeronautics and Space Administration. This work is partly based on observations made with CRIRES on ESO telescopes at the Paranal Observatory under programs 179.C-0151, 093.C-0432, 079.C-0349, 081.C-0833, 091.C-0671. This work is partly based on observations obtained with NIRSPEC at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of Cali- fornia, and the National Aeronautics and Space Administration. The observatory was made possible by the generous financial support of the W. M. Keck Founda- tion. This project has made use of the SciPy stack (Jones et al. 2001 -- ), including NumPy (Oliphant 2006 -- ) and Matplotlib (Hunter 2007). Dullemond, C. P. & Dominik, C. 2004, A&A, 417, 159 Eisner, J. A., Hillenbrand, L. A., & Stone, J. M. 2014, MNRAS, 443, 1916 Ercolano, B. & Pascucci, I. 2017, Royal Society Open Science, 4, 170114 Faure, A. & Josselin, E. 2008, A&A, 492, 257 Fedele, D., Carney, M., Hogerheijde, M. R., et al. 2017, A&A, 600, A72 France, K., Schindhelm, E., Bergin, E. A., Roueff, E., & Abgrall, H. 2014, ApJ, 784, 127 Garcia Lopez, R., Tambovtseva, L. V., Schertl, D., et al. 2015, A&A, 576, A84 Garufi, A., Meeus, G., Benisty, M., et al. 2017, A&A, 603, A21 Gordon, S. & McBride, B. J. 1994, Computer Program for Calculation of Com- plex Chemical Equilibrium Compositions and Applications. Part 1: Analysis (Washington, DC: NASA) Gorti, U., Hollenbach, D., & Dullemond, C. P. 2015, ApJ, 804, 29 Gravity Collaboration, Abuter, R., Accardo, M., et al. 2017, A&A, 602, A94 Hein Bertelsen, R. P., Kamp, I., Goto, M., et al. 2014, A&A, 561, A102 Hollenbach, D. & McKee, C. F. 1979, ApJS, 41, 555 Huang, J., Andrews, S. M., Dullemond, C. P., et al. 2018, ApJ, 869, L42 Hunter, J. D. 2007, Computing in Science & Engineering, 9, 90 Ilee, J. D., Fairlamb, J., Oudmaijer, R. D., et al. 2014, MNRAS, 445, 3723 Isella, A., Huang, J., Andrews, S. M., et al. 2018, ApJ, 869, L49 Jones, E., Oliphant, T., Peterson, P., et al. 2001 -- , SciPy: Open source scientific tools for Python, [Online; accessed 5 July 2001] Kaeufl, H.-U., Ballester, P., Biereichel, P., et al. 2004, in Proc. SPIE, Vol. 5492, Ground-based Instrumentation for Astronomy, ed. A. F. M. Moorwood & M. Iye, 1218 -- 1227 Kama, M., Folsom, C. P., & Pinilla, P. 2015, A&A, 582, L10 Maaskant, K. M., Honda, M., Waters, L. B. F. M., et al. 2013, A&A, 555, A64 McBride, B. J. & Gordon, S. 1996, Computer Program for Calculation of Com- plex Chemical Equilibrium Compositions and Applications II. User's Manual and Program Description, Vol. 19 (Washington, DC: NASA), 178 McLean, I. S., Becklin, E. E., Bendiksen, O., et al. 1998, in Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 3354, In- frared Astronomical Instrumentation, ed. A. M. Fowler, 566 -- 578 Meeus, G., Waters, L. B. F. M., Bouwman, J., et al. 2001, A&A, 365, 476 Meijerink, R., Pontoppidan, K. M., Blake, G. A., Poelman, D. R., & Dullemond, C. P. 2009, ApJ, 704, 1471 Menu, J., van Boekel, R., Henning, T., et al. 2015, A&A, 581, A107 Mitchell, G. F. 1984, Astrophysical Journal Supplement Series, 54, 81 Najita, J., Carr, J. S., & Mathieu, R. D. 2003, ApJ, 589, 931 Oliphant, T. 2006 -- , NumPy: A guide to NumPy, USA: Trelgol Publishing, [On- line; accessed 5 July 2019] Owen, J. E. 2016, PASA, 33, e005 Pinilla, P., Benisty, M., & Birnstiel, T. 2012, A&A, 545, A81 Pinilla, P., Klarmann, L., Birnstiel, T., et al. 2016, A&A, 585, A35 Pinilla, P., Tazzari, M., Pascucci, I., et al. 2018, ApJ, 859, 32 Pontoppidan, K. M., Blake, G. A., & Smette, A. 2011a, ApJ, 733, 84 Pontoppidan, K. M., Boogert, A. C. A., Fraser, H. J., et al. 2008, ApJ, 678, 1005 Pontoppidan, K. M., Salyk, C., Blake, G. A., et al. 2010, ApJ, 720, 887 Pontoppidan, K. M., van Dishoeck, E., Blake, G. A., et al. 2011b, The Messenger, 143, 32 References Andrews, S. M., Huang, J., Pérez, L. M., et al. 2018, ApJ, 869, L41 Avenhaus, H., Quanz, S. P., Schmid, H. M., et al. 2017, AJ, 154, 33 Banzatti, A., Garufi, A., Kama, M., et al. 2018, A&A, 609, L2 Banzatti, A. & Pontoppidan, K. M. 2015, ApJ, 809, 167 Banzatti, A., Pontoppidan, K. M., Bruderer, S., Muzerolle, J., & Meyer, M. R. Rayner, J., Tokunaga, A., Jaffe, D., et al. 2016, in Proc. SPIE, Vol. 9908, Ground- based and Airborne Instrumentation for Astronomy VI, 990884 Rothman, L., Gordon, I., Babikov, Y., et al. 2013, Journal of Quantitative Spec- troscopy and Radiative Transfer, 130, 4 Salyk, C., Blake, G. A., Boogert, A. C. A., & Brown, J. M. 2011, ApJ, 743, 112 Salyk, C., Pontoppidan, K. M., Blake, G. A., et al. 2008, ApJ, 676, L49 Schegerer, A. A., Ratzka, T., Schuller, P. A., et al. 2013, A&A, 555, A103 Schöier, F. L., van der Tak, F. F. S., van Dishoeck, E. F., & Black, J. H. 2005, Banzatti, A., Pontoppidan, K. M., Salyk, C., et al. 2017, ApJ, 834, 152 Barber, R. J., Tennyson, J., Harris, G. J., & Tolchenov, R. N. 2006, MNRAS, Song, L., Balakrishnan, N., Walker, K. M., et al. 2015, ApJ, 813, 96 Spaans, M., Tielens, A. G. G. M., van Dishoeck, E. F., & Bakes, E. L. O. 1994, A&A, 432, 369 ApJ, 437, 270 2015, ApJ, 798, L16 368, 1087 K. M. 2011, A&A, 527, A119 A&A, 347, 375 Bast, J. E., Brown, J. M., Herczeg, G. J., van Dishoeck, E. F., & Pontoppidan, Beegle, L. W., Ajello, J. M., James, G. K., Dziczek, D., & Alvarez, M. 1999, Benisty, M., Stolker, T., Pohl, A., et al. 2017, A&A, 597, A42 Blake, G. A. & Boogert, A. C. A. 2004, ApJ, 606, L73 Boehler, Y., Ricci, L., Weaver, E., et al. 2018, ApJ, 853, 162 Brandl, B. R., Feldt, M., Glasse, A., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 914721 Brittain, S. D., Carr, J. S., & Najita, J. R. 2018, PASP, 130, 074505 Brittain, S. D., Simon, T., Najita, J. R., & Rettig, T. W. 2007, ApJ, 659, 685 Brown, J. M., Herczeg, G. J., Pontoppidan, K. M., & van Dishoeck, E. F. 2012, ApJ, 744, 116 Brown, J. M., Pontoppidan, K. M., van Dishoeck, E. F., et al. 2013, ApJ, 770, 94 Bruderer, S. 2013, A&A, 559, A46 Bruderer, S., Harsono, D., & van Dishoeck, E. F. 2015, A&A, 575, A94 Bruderer, S., van Dishoeck, E. F., Doty, S. D., & Herczeg, G. J. 2012, A&A, 541, A91 Chandra, S. & Sharma, A. K. 2001, A&A, 376, 356 Cieza, L., Padgett, D. L., Stapelfeldt, K. R., et al. 2007, ApJ, 667, 308 Currie, T. & Kenyon, S. J. 2009, AJ, 138, 703 di Folco, E., Dutrey, A., Chesneau, O., et al. 2009, A&A, 500, 1065 Du, F. & Bergin, E. A. 2014, ApJ, 792, 2 Article number, page 18 of 25 Stolker, T., Dominik, C., Avenhaus, H., et al. 2016, A&A, 595, A113 Tang, Y.-W., Guilloteau, S., Dutrey, A., et al. 2017, ApJ, 840, 32 Tennyson, J., Zobov, N. F., Williamson, R., Polyansky, O. L., & Bernath, P. F. 2001, Journal of Physical and Chemical Reference Data, 30, 735 Thi, W. F., Kamp, I., Woitke, P., et al. 2013, A&A, 551, A49 Thi, W.-F., van Dalen, B., Bik, A., & Waters, L. B. F. M. 2005, A&A, 430, L61 Umebayashi, T. & Nakano, T. 1988, Progress of Theoretical Physics Supple- van der Marel, N., van Dishoeck, E. F., Bruderer, S., et al. 2016, A&A, 585, A58 van der Marel, N., van Dishoeck, E. F., Bruderer, S., et al. 2013, Science, 340, van der Marel, N., van Dishoeck, E. F., Bruderer, S., Pérez, L., & Isella, A. 2015, ment, 96, 151 1199 A&A, 579, A106 van der Marel, N., Williams, J. P., Ansdell, M., et al. 2018, ApJ, 854, 177 van der Plas, G., van den Ancker, M. E., Waters, L. B. F. M., & Dominik, C. 2015, A&A, 574, A75 van der Plas, G., Wright, C. M., Ménard, F., et al. 2017, A&A, 597, A32 van der Tak, F. F. S., Black, J. H., Schöier, F. L., Jansen, D. J., & van Dishoeck, E. F. 2007, A&A, 468, 627 Walker, K. M., Song, L., Yang, B. H., et al. 2015, ApJ, 811, 27 White, J. A., Boley, A. C., MacGregor, M. A., Hughes, A. M., & Wilner, D. J. 2018, MNRAS, 474, 4500 Woitke, P., Kamp, I., & Thi, W.-F. 2009, A&A, 501, 383 Yang, B., Stancil, P. C., Balakrishnan, N., & Forrey, R. C. 2010, ApJ, 718, 1062 Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Appendix A: CO molecule model Appendix A.1: Rovibrational The CO molecule model used in this work contains 205 energy levels distributed over five vibrational states (J = 0 − 40, v = 0− 5). Energy levels and line strengths have been extracted from the HITRAN database Rothman et al. (2013). Collisional exci- tation by H2 and H was included. For the CO-H2 collisional de- excitation rates the collisional rates from Bruderer et al. (2012), based on Yang et al. (2010) have been used. The collisional rate matrix was expanded using the formalisms in Chandra & Sharma (2001). For the H-CO collisions, the pure rotational rate coefficients were taken from Walker et al. (2015) while the rovibrational rate coefficients were taken from Song et al. (2015). The rates from Walker et al. (2015) were used for all ∆v = 0 transitions. Appendix A.2: Electronic The electronic excitation of CO has been done in post process- ing. From the model output the total rate into and out of each vi- brational state due to collisions and photon absorption and emis- sion was calculated. To these vibrational band-to-band rates the electronic rates are added and a new vibrational equilibrium is calculated. Using the rotational distribution from the model, a new rovibrational level distribution is calculated. This is then used for ray tracing. The electronic transitions and Einstein A coefficients were taken from Beegle et al. (1999). The Einstein A coefficients were used to calculate a transition probability matrix for an absorption and subsequent emission of a UV photon. It is assumed that elec- tronic relaxation into a vibrational level of the ground state with v(cid:48)(cid:48) > 4 will further cascade into v(cid:48)(cid:48) = 4 and will be treated as if the relaxation directly goes into v(cid:48)(cid:48) = 4. By construction the rotational distribution of the molecules was assumed to be unaf- fected by the electronic excitation. Appendix B: Excitation tests To increase our understand of the CO line formation processes in DALI, a few variations on the standard temperature and ex- citation calculations have been done. In one set of models, the thermo-chemistry has been skipped and Tgas = Tdust has been assumed. The CO abundance in these models is parameterized according to Eq. (5). The excitation has been tested without IR pumping and with UV pumping. Results have been plotted in Fig. B.1. Assuming equal gas and dust temperature lowers the vibra- tional ratio somewhat for the models with small inner radii and has slightly increased vibrational ratios for models with 5 to 10 AU gaps. For models with gaps smaller than 5 AU the vibrational ratio is still higher than the data. The models with equal gas and dust temperature also predict very low (> 0.05) vibrational ratios for models with gaps larger than 10 AU. Fig. B.1 shows that the thermo-chemical balance does influence the line fluxes and line flux ratios for CO. A returning topic of discussion for vibrational excitation of molecules is the treatment and relative importance of pumping by radiation of different wavelengths. In DALI the radiation that pumps the lines, radiation that is absorbed by the molecules in- creasing their excitation, is always assumed to come from either radially inward, or vertically upward, whichever has the lowest line optical depth. This could, depending on the situation, both Fig. B.1. Ratio of the flux from the v = 2 and v = 1 levels of CO versus the inferred radius of emission for observational data and DALI model results. Line connect the dots in order of inner model radius. Labels indicate the different assumptions for the excitation calculation. All models have a gas-to-dust ratio of 100. The model with the largest cavity is always at the largest CO radius. Different assumptions on the excitation are tested. over- and under-predict the effective pumping flux. For CO there are two possible ways to excite vibrational levels through the absorption of a photon. The first is the absorption of an infrared photon through the rovibrational lines around 4.7 or 2.3 µm rais- ing the vibrational excitation by one or two quanta directly. In Fig. B.1 a comparison is made between models with and with- out infrared pumping, but otherwise using the same abundance and temperature structure. The general trend is that the infrared pumping lowers the vibrational ratio, especially in the inner re- gions of the disk. Infrared pumping increases the excitation in the v = 1 state more strongly than the v = 2 state. This greatly increases the vertical and radial extent of the emitting region of the v1 lines increasing line flux. Although the v = 2 state can be directly pumped from the ground state, the Einstein A coefficient for these lines is small, contributing little to the overall excitation of the second vibrationally excited state. The inclusions of UV pumping has a small effect of the line fluxes. However it mostly effects the disks with cavities smaller than 10 AU and in these models the vibrational ratio is lowered. The regions were the UV field is strongest is in the cavity wall near the star. These regions show emission with a vibrational ratio close to unity. UV pumping and the vibrational cascade would create a vibrational ratio of ∼ 0.6 thus lowering the con- tribution of the v2 line in the inner region in favour of the v1 line. At radii larger than 10 AU the UV field is so diluted that it can no longer affect the vibrational ratio. Fig. B.1 shows UV pump- ing without taking into account the self-absorption of UV pho- tons by CO molecules. Including self-absorption only changes the vibrational ratios by a few percent. Appendix C: Line profiles Figure C.1 shows the line profiles for the disks with the smaller inner holes. Most line profiles show two components, a broad component that is present in both the v2 and the v1 line and a nar- row component, that is mostly seen in the v1 line. The velocity Article number, page 19 of 25 100101CO radius (AU)102101100v2/v1FiducialTgas=TdustUV pumpingNo IR pumping A&A proofs: manuscript no. Herbig_rovib of the broad component is the component used in the flux ratio extraction and radius determination. This component is consis- tent with emission dominated by the inner wall of the model. The narrow component is more extended and thus has to come from the disk surface. Figure C.2 shows the line profiles for the disks with larger holes. Almost all of the emission in these models is dominated by the inner wall. The lines are very strong with respect to the continuum, in some cases line-to-continuum ratios reach over 100. Part of this is the NIR excess, which is almost non existent, but even taking that into account, some of the line fluxes are a factor of 10 above the brightest lines observed. Figure C.3 shows the line profiles for the disks with emission from the inner rim subtracted. Many of these models show very low v2/v1 emission. Variations in outer radius and vertical struc- ture redistribute the CO emission over the disk surface, creating a large assortment of line profiles. Appendix D: Near-infrared excess The near-infrared continuum is another good probe of the in- ner disk. As all of the models discussed above have the same dust structure, they show identical near-infrared excesses, only depending on the inner edge radius of the disk in the model. Fig. D.1 shows the near-infrared excess as function of RCO for the gas-to-dust = 100 model and the data. The gas-to-dust 1000 and 10000 overlap almost exactly, the LTE and gas-to-dust 10 models deviate at small inner model radii. The models match the near-infrared excess range for most of the observations, except for the four observed disks with FNIR/F(cid:63) around 0.3. These disks have been shown to require very vertically extended dust structures to fit the NIR SED (Maaskant et al. 2013). The good match between the model and the observational near-infrared excesses indicate that the verti- cal extent, hc, used in the model is reasonable for most of the observed sources. Up to RCO = 5 AU there is only very little variation in the NIR-excess in both the models and the group II objects, indication that up to 5 AU the NIR excess is a bad discriminator of gap size. Appendix D.1: CO as tracer of the inner disk radius The half width half maximum of the v1 lines, or of their broad component where present (see Section 2), is used to estimate a characteristic emitting radius for the inner molecular gas in the disk (e.g. Salyk et al. 2008; Brown et al. 2013; Banzatti & Pontoppidan 2015). Using the results from our thermo-chemical models we can infer how well the relation between CO emission radius and inner (molecular) disk radius holds. Figure D.2 shows that RCO is always larger than the model inner radius. For the fiducial thermo-chemical models the deviations are around 50%, if the inner rim of the disk is detected. If the inner rim is too weak to be detected, or in cases where it is artificially removed, the relation between RCO and Rin breaks down as the disk surface far beyond the inner disk radius can dominate the emission. The low vibrational ratios at small RCO can only be matched by disk surface emission. It is thus for these sources that the largest discrepancy is expected between the measured RCO and the innermost radius where dust or molecular gas is actually present. The model explorations in Sec. 4.3 show that a single inner model radius can lead to a spread in measured CO radii. In these cases the line profiles give a clear sign of CO emission within RCO. Article number, page 20 of 25 The high vibrational ratios at large RCO are best matched by emission from a inner cavity wall. As such RCO tightly traces the inner most radius at which CO is present. This is also clearly seen in the observed line profiles for these sources which have very steep sides. (cid:35) , exp Appendix E: Lowering the flux of the outer disk The fiducial, high gas-to-dust ratio models that can match both RCO and the high line ratios have v1 fluxes that are around a fac- tor 50 higher than that of the average source. Comparisons be- tween the RADEX model predictions in Fig. 6 and the emitting area in the DALI models with 10 and 15 AU cavities showed that both the emitting area and the temperature of the gas had to be reduced. Reducing the models scale height from 0.1 to 0.03, in line with van der Marel et al. (2016) lowered the fluxes by about a factor of 3 (e.g. HD 142527, HD 135344B, HD 36112 and HD 31293). The emitting area was further reduced by putting χset = 1. This lowered the flux from the disk surface. To lower the gas temperature at the inner wall of the disk we considered a different structure of the inner wall than the step-function we had. It turned out that a structure with an decreasing density in- wards of the inner radius worked well. The gas density within the inner radius (Rin) was parameterized by: −1 2 ngas(r, z) = ngas(Rin, 0) (r − Rin)2 + z2 1√ 2πH(Rin) (cid:17)ψ, with the parameters (E.1) where ngas(Rin, 0) is the midplane density at the inner disk edge from the fiducial model. H(r) = hcr as defined in Table 2. Rin was taken to be either 10 or 15 AU. Within the cavity radius, it was assumed that there were only small grains and the dust density was decreased to reach gas- to-dust ratios of 20000 and 100000. The CO lines from these models were calculated and are shown in Fig. E.1. The high gas-to-dust ratios are necessary to reach get the vi- brational ratios to the observed values. At these gas-to-dust ratios the dust is nearly transparent to the CO lines and varying the gas- to-dust ratio has little effect on the line ratios. To further increase the line ratios, larger gas columns would be needed. These structures are probably not the only structures that can explain the CO rovibrational observables. Modelling of other ob- servations of CO such as high resolution ALMA images or Her- schel observations of high J CO lines would be able to differen- tiate between different structures. This modelling will have to be done on a source by source basis and is left to future work. H2(Rin) (cid:34) (cid:16) r Rc Appendix F: Thermal dissociation of CO As mentioned before dissociating CO is difficult. The CO photo- dissociation rate is an order of magnitude lower than the disso- ciation rate for CO2 or H2O for a 10000 K black body radiation field. Furthermore, most reactions destroying CO have a large barrier (e.g. CO + H −−−→ C + OH, 77000K Hollenbach & Mc- Kee 1979; Mitchell 1984) or create products that dissociate back into CO (e.g. CO + OH −−−→ CO2 + H). The strong UV fields near a Herbig star, especially when cou- pled with a low dust opacity, can photo-dissociate CO. This is most effective at lower densities. At higher densities (1012− 1014 cm−3), the chemical model in DALI shows that the CO produc- tion rate is faster than the UV radiation field expected at the sub- limation radius. The high radiation field could however increase the gas temperature to far above the dust temperature. Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. C.1. Normalised model line profiles for the v1 (black) and the v2 (blue) lines for a subset of the models at the native resolution of the model, R = 106. The text on the left of each panel denotes the model set. The top right corner of each panel denotes the inner radius of the model. The vertical bar in the bottom right of each panel shows 0.03 of the continuum flux density. All lines are modelled assuming a 45 degree inclination. No noise has been added to these lines, features in the line profiles are due to the sampling of the DALI grid. Article number, page 21 of 25 0.00.51.00.4Fiducialv2P(4)v1P(10)0.91.352.03.00.00.51.00.4g/d = 100.91.352.03.00.00.51.00.4g/d = 10000.91.352.03.00.00.51.00.4g/d = 100000.91.352.03.00.00.51.00.4LTE0.91.352.03.00.00.51.00.4Tdust0.91.352.03.00.00.51.00.4UV pump0.91.352.03.0500500.00.51.00.4No IR pump500500.9500501.35500502.0500503.0Velocity (km/s)Normalized line flux A&A proofs: manuscript no. Herbig_rovib Fig. C.2. Same as Fig. C.1, but for models with inner radii > 4.5 AU. The right bar in the bottom right of each panel shows 0.3 times the continuum. Article number, page 22 of 25 0.00.51.04.5Fiducialv2P(4)v1P(10)6.7510.015.00.00.51.04.5g/d = 106.7510.015.00.00.51.04.5g/d = 10006.7510.015.00.00.51.04.5g/d = 100006.7510.015.00.00.51.04.5LTE6.7510.015.00.00.51.04.5Tdust6.7510.015.00.00.51.04.5UV pump6.7510.015.0500500.00.51.04.5No IR pump500506.755005010.05005015.0Velocity (km/s)Normalized line flux Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. C.3. Same as Fig. C.1, but for models were the inner rim contribution has been subtracted. Article number, page 23 of 25 0.00.51.00.4g/d = 102,Rinv2P(4)v1P(10)0.61.350.00.51.00.4g/d = 103,Rin0.61.350.00.51.00.4g/d = 104,Rin0.61.350.00.51.03g/d = 102,Rout580.00.51.03g/d = 104,Rout580.00.51.00.0, 0.02g/d = 102,,hc0.0, 0.10.25, 0.02400400.00.51.00.0, 0.02g/d = 104,,hc400400.0, 0.1400400.25, 0.02Velocity (km/s)Normalized line flux A&A proofs: manuscript no. Herbig_rovib Fig. D.1. CO emitting radius versus the near-infrared excess flux as frac- tion of the total stellar flux. Black circles show the inner model radius against the near-infrared excess for the fiducial, g/d=100, model (see Sec. 4 for details). The data are shown following the same color coding as in Fig. 1. Fig. D.2. Model radius versus the deviation of RCO from the inner model radius. Filled circles correspond to the normal DALI models, the stars correspond to the DALI models with the contribution of the inner rim re- moved. Only the models sets with varying inner rim location are shown. At temperatures above 3000 K Gibbs free energy minimis- ers find CO abundances lower than the canonical 10−4 (Fig. F.1). Around this temperature, the kinetical network also shows a de- crease in the CO abundance as the endothermic CO +H −−−→ C + OH and OH + H −−−→ O + H2 reactions coupled with collisional dissociation of H2 push the gas towards a fully atomic state. Figure F.1 clearly shows that the molecular to atomic transi- tion happens slower and at higher temperatures than in the equi- librium models. The kinetic model does not have all the reaction pathways included that could be important for the production and destruction of CO and H2 at high temperatures. Between the uncertainty in the gas temperature and the uncertainty in the chemistry, it is thus not unlikely that the CO abundance in the inner disk rim is overestimated by DALI. Article number, page 24 of 25 Fig. E.1. v1 line flux (top) and vibrational ratio of CO (bottom) versus the inferred radius of emission for observational data and DALI model results. Cyan and red markers show models with a gas-to-dust ratio of 20000 and 100000 respectively. Circles show models with Rin = 10 AU, squares show models with with Rin = 15 AU 100101CO radius (AU)0.00.10.20.30.40.5NIR excess (FNIR/F)v2/v10.16v2/v10.05v2/v1>0.2Fiducial model100101Model Rin (AU)100101RCO/RinRin=RCOFiducialg/d = 10g/d = 1000g/d = 10000LTERim subtracted101910181017101610151014v1 flux (W m2)100101CO radius (AU)102101100v2/v1Observ. Arthur D. Bosman et al.: Probing the inner disk of Herbig Ae sources with CO rovibrational emission Fig. F.1. CO abundance as function of temperature under inner disk conditions. Black points show the results from the chemical model in DALI under conditions relevant for the inner disk (n = 1012 -- 1014 cm−3, 104 -- 1010G0 radiation field). Blue line show the results from equilibrium chemistry for the same range of densities (Chemical Equilibrium with Applications, Gordon & McBride 1994; McBride & Gordon 1996). Article number, page 25 of 25 100020003000400050006000Temperature (K)109108107106105104CO abundanceEquilibrium chem.Dali chem
1810.12971
1
1810
2018-10-30T19:31:47
Fully scalable forward model grid of exoplanet transmission spectra
[ "astro-ph.EP" ]
Simulated exoplanet transmission spectra are critical for planning and interpretation of observations and to explore the sensitivity of spectral features to atmospheric thermochemical processes. We present a publicly available generic model grid of planetary transmission spectra, scalable to a wide range of H$_2$/He dominated atmospheres. The grid is computed using the 1D/2D atmosphere model ATMO for two different chemical scenarios, first considering local condensation only, secondly considering global condensation and removal of species from the atmospheric column (rainout). The entire grid consists of 56,320 model simulations across 22 equilibrium temperatures (400 - 2600 K), four planetary gravities (5 - 50 ms$^{-2}$), five atmospheric metallicities (1x - 200x), four C/O ratios (0.35 - 1.0), four scattering haze parameters, four uniform cloud parameters, and two chemical scenarios. We derive scaling equations which can be used with this grid, for a wide range of planet-star combinations. We validate this grid by comparing it with other model transmission spectra available in the literature. We highlight some of the important findings, such as the rise of SO$_2$ features at 100x solar metallicity, differences in spectral features at high C/O ratios between two condensation approaches, the importance of VO features without TiO to constrain the limb temperature and features of TiO/VO both, to constrain the condensation processes. Finally, this generic grid can be used to plan future observations using the HST, VLT, JWST and various other telescopes. The fine variation of parameters in the grid also allows it to be incorporated in a retrieval framework, with various machine learning techniques.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 12 (2018) Preprint 1 November 2018 Compiled using MNRAS LATEX style file v3.0 Fully scalable forward model grid of exoplanet transmission spectra Jayesh. M. Goyal 1(cid:63), Hannah. R. Wakeford2, Nathan. J. Mayne1, Nikole. K. Lewis3, Benjamin. Drummond1, David. K. Sing1,4 1Astrophysics Group, Physics Building, Stocker Road, University of Exeter, Devon EX4 4QL, UK 2Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 3Department of Astronomy and Carl Sagan Institute, Cornell University, 122 Sciences Drive, Ithaca, NY, 14853, USA 4Department of Earth and Planetary Sciences, Johns Hopkins University, Baltimore, MD, USA Accepted October 29th 2018. Received October 15th 2018; in original form July 31st 2018 ABSTRACT Simulated exoplanet transmission spectra are critical for planning and interpreta- tion of observations and to explore the sensitivity of spectral features to atmospheric thermochemical processes. We present a publicly available generic model grid of plan- etary transmission spectra, scalable to a wide range of H2/He dominated atmospheres. The grid is computed using the 1D/2D atmosphere model ATMO for two different chem- ical scenarios, first considering local condensation only, secondly considering global condensation and removal of species from the atmospheric column (rainout). The en- tire grid consists of 56,320 model simulations across 22 equilibrium temperatures (400 - 2600 K), four planetary gravities (5 - 50 ms−2), five atmospheric metallicities (1x - 200x), four C/O ratios (0.35 - 1.0), four scattering haze parameters, four uniform cloud parameters, and two chemical scenarios. We derive scaling equations which can be used with this grid, for a wide range of planet-star combinations. We validate this grid by comparing it with other model transmission spectra available in the litera- ture. We highlight some of the important findings, such as the rise of SO2 features at 100x solar metallicity, differences in spectral features at high C/O ratios between two condensation approaches, the importance of VO features without TiO to constrain the limb temperature and features of TiO/VO both, to constrain the condensation processes. Finally, this generic grid can be used to plan future observations using the HST, VLT, JWST and various other telescopes. The fine variation of parameters in the grid also allows it to be incorporated in a retrieval framework, with various machine learning techniques. Key words: planets and satellites: atmospheres -- planets and satellites: composition -- planets and satellites: gaseous planets -- techniques: spectroscopic 1 INTRODUCTION Planning and interpretation of exoplanet atmospheric char- acterization observations necessarily rely on theoretical spectra generated from atmospheric models. There currently exist several exoplanet atmospheric forward model grids and simulators providing publicly available theoretical transmi- sison spectra (e.g. Fortney et al. 2010; Molli`ere et al. 2016; Kempton et al. 2017; Goyal et al. 2018) all of which make specific choices about the atmospheric physics, chemistry, radiative transfer and spectroscopic line lists incorporated (cid:63) E-mail: [email protected] © 2018 The Authors into their models. However, inter-comparisons between these modeling frameworks still prove difficult. Many of the exo- planet spectral databases produced to date cover very spe- cific non-overlapping parts of exoplanet atmospheric phase space. Often choices made by specific teams concerning the underlying spectroscopic line lists and physical processes like condensation and rainout are not clearly outlined, which can lead to disagreements and confusion when the models are ap- plied by scientists outside the team. Furthermore, the sen- sitivity of these theoretical spectra to these specific physics choices are often not explored. Publicly accessible databases that provide representative theoretical spectra for exoplanet atmospheres spanning a broad range of atmospheric proper- 2 J. M. Goyal et al. ties and physical assumptions are necessary to determine the validity of our understanding of these distant worlds to our theoretical constructs. These grids also provide a straight forward way to test spectral sensitivity both within a given modeling framework and across modeling frameworks. Transmission spectra observations of exoplanet atmo- spheres have been increasing in quality and resolution since the commissioning of the Wide Field Camera 3 (WFC3) in- strument on board the Hubble Space Telescope (HST) (e.g. Deming et al. 2013; Kreidberg et al. 2014; Wakeford et al. 2016; Sing et al. 2016; Evans et al. 2016, 2017; Wakeford et al. 2018) and FORS2 on the Very Large Telescope (VLT) (Nikolov et al. 2018). This increase in data fidelity has also motivated development of grids of models covering fine vari- ations of model parameters, especially those that alter the chemistry most significantly. Robust line-lists are at the core of any spectral gener- ation tool. The HITRAN (High Resolution TRANsmission) database (Rothman et al. 2013) has been the source of spec- tral line-lists for most of the previous models. This database is established at a reference temperature of 296 K (Rothman et al. 2010), with HITEMP (Rothman et al. 2010) its high temperature version available only for certain molecules. Some hot Jupiter exoplanet atmospheres can reach tempera- tures as high as 3000 K, where HITRAN line-lists can under- estimate absorption of radiation by several orders of mag- nitude. However, more accurate spectral line-lists for hot atmospheres have been developed by the EXOMOL project (Tennyson et al. 2016), which has also motivated develop- ment of updated more accurate grids of models. In this paper we present a publicly available1 new generic grid of forward model simulations that can be scaled to a wide range of star-planet pairs for atmospheric trans- mission spectra. We provide a user-friendly generic grid of simulated transmission spectra to interpret observations, where the word "generic" implies a grid of models that can be scaled to a wide range of H2/He dominated exoplanet atmo- spheres. We highlight the sensitivity of our model transmis- sion spectra to choices in atmospheric physics, such as con- densation schemes. We also provide comparisons between our scalable grid and planet specific grid as well as spec- tra generated by other atmospheric modeling frameworks. These comparisons critically highlight how choices made in generating atmospheric models influence our interpretation of the underlying physics captured by observations. In Sec- tion 2 we first detail the model, its setup for the grid and treatment of condensation. In Section 3 we describe the pa- rameter space of the grid. In Section 4 we present the scien- tific results obtained from the grid by detailing the effects of sensitivity tests on atmospheric chemical composition and the resultant transmission spectra. In Section 5 we detail how the models can be scaled to any planet-star combina- tion and compare them with other published model grids in the literature for validation. In Section 6 we discuss the application of the grid with specific reference to the trans- mission spectral index established in Sing et al. (2016) and finally we conclude in Section 7. 1 https://drive.google.com/open?id=1ZFbkPdqg37_ Om7ECSspSpEp5QrUMfA9J 2 ATMO FORWARD MODELS We use ATMO, a 1D-2D radiative-convective equilibrium model for planetary atmospheres (Tremblin et al. 2015, 2016; Amundsen et al. 2014; Drummond et al. 2016; Tremblin et al. 2017; Goyal et al. 2018) to compute a grid of generic forward models, which can be scaled to represent a wide range of H2/He dominated atmospheres. For this work we use isothermal P-T profiles under the assumption of chemical equilibrium. We include H2-H2 and H2-He collision induced absorption (CIA) opacities. We also include opacities due to H2O, CO2, CO, CH4, NH3, Na, K, Li, Rb, Cs, TiO, VO, FeH, CrH, PH3, HCN, C2H2, H2S and SO2. The source of these opacities and their pressure broadening parameters can be found in Amundsen et al. (2014) and Goyal et al. (2018). We note that in this work we adopt Na and K pressure broadened line profiles as derived in Burrows et al. (2000), instead of Allard et al. (2003), adopted for planet specific grid presented in Goyal et al. (2018). This was motivated by the results of (Nikolov et al. 2018) where the profiles of Burrows et al. (2000) had a slightly better fit to observa- tions than that of Allard et al. (2003), although the final results were statistically inconclusive while comparing these different pressure broadened profiles. This grid of model simulations is baselined for a Jupiter radius planet (1 RJ at 1 millibar pressure) around a Solar radius star (1 Rsun). Our isothermal P-T profiles extend from 10−6 bar at the top of the atmosphere to 10 bar at the bot- tom, with the radius of the simulated planet that is 1 RJ , defined at the 1 millibar pressure level, which approximates the region of the atmosphere probed with transmission spec- tra (Lecavelier Des Etangs et al. 2008). The upper and the lower boundary conditions of the atmosphere, for the equil- brium chemical abundances calculation are set by the input pressure grid. We use 50 model levels for each isothermal P-T profile, which are evenly spaced in log(P) space. Over a large pressure range the assumption of an isothermal atmo- sphere is an extreme assumption, especially for isothermal P-T profiles with rainout (see Section 2.1). Therefore, we set the bottom of the atmosphere pressure to 10 bar, approxi- mately similar to previous works (e.g Fortney et al. 2010). We compute transmission spectra in 5000 correlated-k bins, evenly spaced in wavenumber, which corresponds to R∼5000 at 0.2 µm while decreasing to R∼100 at 10 µm (see Goyal et al. 2018, for details). 2.1 The Treatment of Condensation and Rainout Within our model we adopt three different approaches to compute chemical equilibrium abundances; gas-phase only, local condensation and rainout condensation. We do not pro- duce a grid for the gas phase only approach, as it can lead to unphysical spectral features for planetary atmospheres, where condensates can form. In the local condensation (no- rainout) approach, each model level is entirely independent of all other model levels along the profile, and the chemi- cal composition is only dependent on the local temperature and pressure, and the element abundances. In this approach, each model level starts with the same elemental abundance, and condensates that form, only deplete the material in that specific layer of the atmosphere. In contrast, in the condensation with rainout approach, MNRAS 000, 1 -- 12 (2018) Generic Grid of Transmission Spectra 3 Figure 1. Figure showing transmission spectra with the effects of local condensation (left) and rainout condensation (right). We show the effects of varied temperature (top row) with a fixed metallicity and C/O ratio set to solar, varied metallicity (middle row) with a fixed Teq = 1500 K and C/O at solar, varied C/O ratio (bottom row) with a fixed Teq = 1500 K and solar metallicity. In each case we consider a clear atmosphere with haze and cloud parameter set to 1.0 and 0.0, respectively, for a planet with radius 1 RJ around a star with radius 1 Rsun and gp =10 ms−2. An offset in transit depth has been added to each spectrum for clarity. once the condensates are formed, the elements that com- prise that condensate are depleted stoichiometrically from the concerned layer and all the layers above it. This is driven by the assumption that once condensates are formed they will tend to sink in the atmosphere (a limiting case), and the elements that compose those condensate species will there- fore be depleted for lower pressures. For example, forma- tion of Mg2SiO4 condensate will remove 2:1:4 of the Mg, Si and O atoms in the layers above, this will continue until one or more of the elements are depleted in their entirety. This setup is designed to mimic cloud particles forming and "settling" or "sinking" in an atmosphere (i.e. raining out) (Burrows & Sharp 1999; Mbarek & Kempton 2016). There- fore, in the rainout approach each layer is dependant on the deeper (high pressure) layers of the atmosphere for its chem- MNRAS 000, 1 -- 12 (2018) ical abundances. We note that, since O is one of the most abundant element that gets locked up in the condensates, rainout processes primarily effect the availability of O in the atmosphere for the upper layers. For a detailed description of the implementation of the rainout calculation in our model, see Section 3.3.2 of Drummond (2017)2. P-T profiles in radiative-convective equilibrium are sub- stantially hotter in deeper levels (>10 or 100 bar) compared to upper region (<10 bar) of the atmosphere (except when there is an inversion). Since the rainout process is depen- dant on the adopted P-T profile, assumption of isothermal P-T profile in deeper levels of the atmosphere can be an ex- 2 https://ore.exeter.ac.uk/repository/handle/10871/27993 1100.55Wavelength(µm)100001050011000115001200012500Transitdepth(ppm)1100.55Wavelength(µm)1000011000120001300014000Transitdepth(ppm)1100.55Wavelength(µm)100001050011000115001200012500Transitdepth(ppm)1100.55Wavelength(µm)1000011000120001300014000Transitdepth(ppm)1100.55Wavelength(µm)10000110001200013000Transitdepth(ppm)1100.55Wavelength(µm)10000110001200013000Transitdepth(ppm)Local CondensationRainout CondensationTeq = 1500 K
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2[M/H] = 0.0
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2Teq = 1500 K [M/H] = 0.0
haze = 1 cloud = 0 gp = 10 ms-22500K2000K1500K1000K600K2500K2000K1500K1000K600K200x100x50x10x1x200x100x50x10x1x1.00.70.560.35C/O RatioAtmospheric MetallicityEquilibrium Temperature1.00.70.560.35H2OH2OCOTiOTiO/VONaKCH4H2OH2OCO2/CO/H2OCH4CH4PH3CH4VOH2OH2OCOTiOH2OH2OVONaKCH4CO2/CO/H2OCH4CH4PH3CH4NaKNaKSO2CO2/CO/H2OCO2/CO/H2OSO2TiO/VOTiO/VONaKCOCO2CO2H2OH2OCO2/CO/H2OH2OH2OH2OH2OH2OH2OH2OCH4VOCO2/CO/H2OH2OH2OH2OCH4HCNCH4/HCNCH4HCNCOCH4H2OCH4/HCNCH4/HCN/C2H2CH4/HCN 4 J. M. Goyal et al. treme assumption, as it would form more condensates com- pared to a hotter radiative-convective equilibrium P-T pro- file and thereby effect the chemical abundances in the upper region (low pressure region) of the atmosphere probed by transmission spectra. Therefore, we adopt 10 bars as the lower boundary of our model domain in this work, which is generally the deepest part of the atmosphere, transmission spectra can probe (Fortney et al. 2010). These two differ- ent condensation approaches can have large effects on the chemical equilibrium abundances of various molecules across different parameter spaces and therefore the transmission spectra. Therefore, to make the computed grids applicable to differing scenarios we produce a generic grid for both lo- cal condensation and rainout condensation conditions. We note that the assumption of isothermal P-T profiles is not as accurate as P-T profiles in radiative-convective equilibrium (e.g. Molli`ere et al. 2016, Goyal et al. in prep.). However, isothermal models have been shown to be sufficient to in- terpret current transmission spectra observations (Fortney 2005; Heng & Kitzmann 2017; Goyal et al. 2018), and are much more computationally efficient. 3 GRID PARAMETER SPACE Our aim is to produce a forward model grid which can be scaled to any planet-star pair for both, local condensation and rainout condensation cases. The grid requires a size- able number of parameters which must be sampled with sufficient fineness to allow accurate interpolation. For each of the two treatments of condensation we compute 28,160 forward models over a range of 22 planetary equilibrium temperatures (400, 600, 700, 800, 900, 1000, 1100, 1200, 1300, 1400, 1500, 1600, 1700, 1800, 1900, 2000, 2100, 2200, 2300, 2400, 2500, 2600; all listed in K), four planetary sur- face gravities (5, 10, 20, 50; all listed in ms−2), five atmo- spheric metallicities (1, 10, 50, 100, 200; all in × solar), four C/O ratios (0.35, 0.56, 0.7, 1.0), four scattering haze pa- rameters (1, 10, 100, 1100× standard Rayleigh-scattering) and four uniform cloud parameters (0, 0.06, 0.2, 1). Haze is treated as a small scattering aerosol particles and im- plemented as a parameterized enhanced multi-gas Rayleigh scattering, while clouds are treated as large particles with grey opacity. Haze parameterization is implemented via the equation σ(λ) = αhazeσ0(λ) where σ(λ) is the total scattering crossection with haze, αhaze is the haze enhancement factor and σ0(λ) is the scattering crossection due to all other gases. The cloud parameterization is implemented vi the equation κ(λ)c = κ(λ) + αcloudκH2 , where κ(λ)c is the total scattering opacity in cm2/g, κ(λ) is the scattering opacity due to nom- inal Rayleigh scattering in same units, αcloud is the variable cloudiness factor governing the strength of grey scattering and κH2 is the scattering opacity due to H2 at 350 nm which is ∼ 2.5×10−3cm2/g (see Goyal et al. 2018, for more details). We do not extend the grid to sub-solar metallicities and limit the high metallicity end to 200× solar. Above 200× solar the atmosphere becomes abundant in species other that H2 and He, such as CO2, H2O, CO etc. This would require the inclusion of pressure broadening effects due to these species, and no existing studies have solved this problem due to lack of lab-based observational data (Fortney et al. 2016). The choice of C/O ratios in the grid is guided by important tran- sition regimes as found by previous studies (Madhusudhan 2012; Molli`ere et al. 2015; Goyal et al. 2018). Each of the cloud and haze parameters apply a scaling from 0.0 or 1×, re- spectively, up to extreme values of almost total obscuration (1.0) for uniform clouds and 1100× the scattering parameter for haze. 4 SENSITIVITY ANALYSIS OF THE GRID This model grid explores a wide range of giant planet param- eter space across temperatures and chemical abundances, for two different condensation regimes. Therefore, we detail the sensitivity of spectra to different parameter and physics choices in this section. We show a subset of our simulations for both local condensation (left) and rainout condensation (right) in Fig. 1, demonstrating the changes with tempera- ture (top), metallicity (middle) and C/O ratio (bottom). For the same 3 grid parameters, we also show the changes in the mean chemical abundances of various spectrally important species in Fig. 2. 4.1 Impact of changing temperature Changes in temperature have a more profound impact on the transmission spectrum compared to any other parame- ter, as the chemical composition is strongly dependent on the temperature, under the assumption of chemical equilib- rium (Burrows & Sharp 1999). To demonstrate the impact of temperature on the transmission spectrum we fix the metal- licity and C/O ratio to solar value and use a clear atmo- spheric (haze and cloud parameters of 1.0 and 0.0, respec- tively) spectrum for a planetary gravity of 10 ms−2, as shown in Fig. 1(top row). In both condensation cases, at low tem- peratures <1000 K the spectra are dominated by CH4 with minor contributions from H2O. Above ∼1000 K the spectra are dominated by H2O absorption with contributions in the IR from CO and CO2. This can also be noticed in Fig. 2 bot- tom row, where the CH4 abundance drops dramatically after ∼1000 K. The main difference between the two condensation cases can again be seen in the optical. In the local conden- sation case TiO/VO absorption is expected at temperatures above ∼1400 K, while in rainout condensation, absorption by TiO/VO is suppressed until ∼1700 K. This implies that the atomic Na and K line absorption is only apparent in the local condensation grid between ∼ 800 -- 1400 K. In the rainout condensation grid, the presence of Na and K is also impacted by condensation, therefore Na and K features are only expected between ∼ 1100 -- 1700 K 4.2 Impact of changing metallicity The effects of metallicity are shown for a clear atmosphere at an equilibrium temperature of 1500 K, solar C/O ratio(0.56) and gravity of 10 ms−2. Generally, it can be seen that all the spectral features tend to be reduced in amplitude with in- crease in metallicity as seen in Fig. 1 (middle row), although the abundance of all the spectrally important species in- crease, as seen in Fig. 2. This is caused by the increasing metallicity leading to a reduction in the scale height of the atmosphere, as the mean molecular weight increases. With MNRAS 000, 1 -- 12 (2018) Generic Grid of Transmission Spectra 5 Figure 2. Figure showing change in mean chemical abundances between 0.1 and 100 millibar for various molecules for the same model simulations as in Fig. 1. Changes in mean chemical abundances due to local condensation (left) and rainout condensation (right) for varied temperature (top row) with a fixed solar metallicity and C/O ratio, varied metallicity (middle row) with a fixed Teq = 1500 K and solar C/O ratio, and for varied C/O ratios with a fixed Teq = 1500 K and solar metallicity (bottom row)]. increase in metallicity, there is also increase in CO2 abun- dance (Moses et al. 2013) as seen in Fig. 2, which can also be noticed by rise in strong CO2 feature at 15 µm. In the local condensation case, the strength of TiO/VO features in the optical decreases with increasing metallicity at 1500 K, which is a combination of decrease in TiO/VO abundance as well as muting of the features due to in- creased molecular weight of the atmosphere. In contrast, rainout condensation entirely removes TiO and VO features at 1500 K due to condensation and removal of the species MNRAS 000, 1 -- 12 (2018) Local CondensationRainout CondensationTeq = 1500 K [M/H] = 0.0
haze = 1 cloud = 0 gp = 10 ms-2C/O Ratio1x10x100xMetallicity(xsolar)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO1x10x100xMetallicity(xsolar)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVOTeq = 1500 K
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2Atmospheric Metallicity050010001500200025003000Temperature(K)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO050010001500200025003000Temperature(K)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO[M/H] = 0.0
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2Equilibrium Temperature1100.55Wavelength(µm)950010000105001100011500Transitdepth(ppm)gp 5 ms-2 10 ms-2 20 ms-2 50 ms-20.30.40.50.60.70.80.91.01.1C/Oratio102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2C2H2HCNSO2TiOVO0.30.40.50.60.70.80.91.01.1C/Oratio102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2C2H2HCNSO2TiOVO 6 J. M. Goyal et al. deeper in the atmosphere, leading to their lower abundances. The lower abundances of TiO and VO, leaves the atomic lines of Na and K as the dominant transmission features in the optical. There is, however, little difference (between rainout and no-rainout) in the infrared (IR) with only mod- erate changes in the absolute abundance of the oxygen based species due to condensate formation. The IR spectra are mainly dominated by H2O features. However, CO and CO2 features can be seen between 4 and 5 µm. Additionally, there is a SO2 feature between 7 to 8 µm which first appears in the 100× solar metallicity models, with a stronger feature at 200× solar metallicity. The presence of SO2 can possibly be used to constrain the metallicity of exoplanet atmospheres with mid-IR observations and strong near-IR constraints on the H2O abundance. SO2 is one of the most prominent sul- phur gases in the atmosphere of Venus, having substantial effect on its radiative balance (Vandaele et al. 2017). There- fore, investigating the possibility of SO2 detection, can in- crease our understanding of the sulphur cyle in giant plane- tary atmospheres. 4.3 The impact of changing C/O The effects of varying the carbon-to-oxygen (C/O) ratio are shown in Fig. 1 and 2 (bottom row) for a clear atmosphere at an equilibrium temperature of 1500 K, solar metallicity and gravity of 10 ms−2. The C/O ratio primarily dictates the dominance of various carbon and oxygen bearing molec- ular species in the atmosphere and thereby the spectra. As seen in Fig. 2 bottom row, for C/O ratios less than equal to solar values, atmosphere is dominated by CO, H2O and CO2. However, for C/O ratios greater than solar, the abun- dances of molecules with carbon but without oxygen such as CH4, HCN, and C2H2 start increasing rapidly. This effect is visible in the transmission spectra as shown in Fig. 1 (bot- tom row), where the infrared spectra is primarily dominated by H2O features upto solar C/O ratio (0.56), 0.7 being the transitional value and 1 being the C/O ratio at which the infrared spectra is primarily dominated by CH4 and HCN features. Choice of condensation also has an effect on spec- tral features, especially for C/O ratios greater than solar. It can be seen in Fig. 1 that at C/O ratio of 1, spectral features are different between rainout and local condensa- tion cases. This is because the rainout case has a slighly larger abundance of CH4, HCN and C2H2, as compared to local condensation case seen in Fig. 2. It can also be seen from this figure that the abundances of spectrally important molecules such as H2O, C2H2 etc. are more strongly depen- dent on the C/O ratio in the rainout case, compared to local condensation case. 4.4 The presence of VO without TiO Recent observations have shown the possibility of VO with the absence of TiO in the atmosphere of extremely irra- diated hot Jupiter WASP-121b (Equilibrium temperature (Teq) = 2358 K) (Evans et al. 2017). Our models suggest that this may only be possible across a narrow range of tempera- tures, namely ∼ 1200 -- 1400 K, under the assumption of local condensation as seen in Fig. 2 (top row). In this rather nar- row temperature regime, the abundance of VO is higher than that of TiO and sufficient enough to impart its features in the transmission spectrum. This is because the primary Ti condensate Ti3O5 is more abundant than the V condensate V2O3, thereby locking more Ti in condensates than V at these temperatures. For temperatures higher than 1400 K TiO starts dominating the optical spectrum. However, for the rainout condensation case, this narrow range where VO is present without TiO, is ∼ 1700 -- 1800 K. Therefore, pres- ence of VO features in the spectrum without TiO can help constrain the limb temperature of the planet's atmosphere, if the rainout and local condensation processes in the plan- etary limb are constrained. Additionally, if the planetary limb temperature is constrained using the Rayleigh scatter- ing slope, TiO or VO features can reveal which process is dominant in these atmospheres (rainout or local condensa- tion). 5 WORKING WITH THE GRID The generic exoplanet ATMO model grid has been produced such that it can be scaled to a wide range of planet/star combinations and can be applied as an interpretive tool as well as a preparation tool for exoplanet atmospheric stud- ies. Temperature and gravity are two of the most important parameters shaping the transmission spectrum of giant exo- planets as they effectively control the scale height of the at- mosphere. The temperature also governs the chemical state of the atmosphere, where different temperatures can lead to different chemical properties (see Fig. 2). As such, the parameter space of the temperature (400-2600 K) is broken down into fine bins (of ∼100 K). The gravity, however, can be represented over fewer values and scaled to a more pre- cise value. The amplitude of features in the transmission spectra are strongly tied to the scale height of the planet's atmosphere, which is inversely proportional to the planet's gravity as shown in Fig. 3. In short, all else remaining equal, as the gravity increases, the amplitude of spectral features decrease. To demonstrate the scalability of the gravity, for each gravity parameter in the grid (5, 10, 20, 50 ms−2) we scale the model to a variety of different planetary gravities and compare it to a model specifically generated for that grav- ity. In Fig. 4, we demonstrate the accuracy associated with scaling the four gravities supplied in the grid to a finer pa- rameter space between 3 -- 100 ms−2. The residuals are well below half the scale height of the atmosphere, when scaled from the 5, 10 or 20 ms−2 models as seen in Fig. 4. How- ever, residuals tend to increase beyond half scale height for 50 ms−2 models. One of the reasons for this is the difference between the gravity values (δg) used for scaling, for lower values of g smaller δg is used, thus resulting in smaller resid- uals and for higher values, larger δg results in comparatively larger residuals. It can also be noticed that the residuals are maximum between 0.5 and 1 µm, specifically around Na and K absorption bands at 0.58 and 0.76 µm respectively. This is because when transmission spectra is computed for a specific value of gravity it corresponds to specific value of pressure and therefore specific pressure broadening of opacities. How- ever, scaling to a different gravity changes the pressure level probed by transmission spectra, since it varies as square root of g (see Equation 2) and therefore ideally requires pressure MNRAS 000, 1 -- 12 (2018) Generic Grid of Transmission Spectra 7 Figure 4. Figure showing the residual comparison of the generic model scaled to different planetary gravities to models computed at the specific gravities, for each of the four planetary gravity parameters in the grid (5 ms−2 top, 10 ms−2 middle top, 20 ms−2 middle bottom, 50 ms−2 bottom). Figure 3. Figure showing how changing the gravity changes the size of the absorption features in the transmission spectrum, due to a change in the scale height, for each of the gravities represented in the grid (5 ms−2, 10 ms−2, 20 ms−2, 50 ms−2 ). Each model is for a clear solar metallicity ans solar C/O ratio atmosphere with Teq = 1500 K. broadened opacities at this new pressure levels. But while scaling generic models between different gravities, we don't alter the opacities, leading to largest residuals for Na and K bands which are very strongly affected by pressure broaden- ing. The other parameters covered by the grid represent dif- ferent scaling parameters applied to either the chemistry (e.g. metallicity, C/O ratio) via a scaling of the abundances, or in the opacity (e.g. haze and cloud, via scaling factors with and without wavelength dependence; see Goyal et al. 2018 for details). 5.1 Scaling to Specific Planetary Parameters In order for these models to be applied to the desired ex- oplanet they will have to be scaled based on the planetary radius, stellar radius, surface gravity, and planetary equilib- rium temperature. The wavelength-dependent observed (apparent) radius of the planet Rp(λ) can be seen as a combination of the wavelength-independent bulk planet radius Rp,bulk and a wavelength-dependent contribution from the atmosphere z(λ), Rp(λ) = z(λ) + Rp,bulk , (1) where the term z(λ) is dependent on the physical and chemical properties of the atmosphere. mate analytic solution for z(λ), Lecavelier Des Etangs et al. (2008) provide an approxi- (cid:32) (cid:115)2πRp,bulk (cid:33) z(λ) = H ln ξabsPz=0σabs(λ) τeq kbT µg = H ln α, (2) where T is temperature, µ is the mean molecular weight, g is gravity, kb is the Boltzmann constant and H = (kbT)/(µg) is the atmospheric scale height. σabs(λ) and ξabs are the absorption cross section and mole fraction of the dominant absorbing species, respectively. Pz=0 is the pres- sure at the effective altitude(z) of 0 (i.e at the base of the MNRAS 000, 1 -- 12 (2018) Local CondensationRainout Condensation0.30.40.50.60.70.80.91.01.1C/Oratio102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO0.30.40.50.60.70.80.91.01.1C/Oratio102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVOTeq = 1500 K [M/H] = 0.0
haze = 1 cloud = 0 gp = 10 ms-2C/O Ratio1x10x100xMetallicity(xsolar)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO1x10x100xMetallicity(xsolar)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVOTeq = 1500 K
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2Atmospheric Metallicity050010001500200025003000Temperature(K)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO050010001500200025003000Temperature(K)102010181016101410121010108106104102MeanMoleFractionCOH2OCH4CO2NaKSO2TiOVO[M/H] = 0.0
C/O = 0.56 haze = 1 cloud = 0 gp = 10 ms-2Equilibrium Temperature1100.55Wavelength(µm)950010000105001100011500Transitdepth(ppm)gp 5 ms-2 10 ms-2 20 ms-2 50 ms-21100.55Wavelength(µm)1.00.50.00.51.0AtmosphericScaleHeightDi↵erencebetween5ms2modelscaledtoothergravities1100.55Wavelength(µm)1.00.50.00.51.0AtmosphericScaleHeightDi↵erencebetween10ms2modelscaledtoothergravities1100.55Wavelength(µm)1.00.50.00.51.0AtmosphericScaleHeightDi↵erencebetween20ms2modelscaledtoothergravities1100.55Wavelength(µm)1.00.50.00.51.0AtmosphericScaleHeightDi↵erencebetween50ms2modelscaledtoothergravities3ms-2 4ms-2 5ms-2 6ms-2 7ms-2 8ms-2 9ms-2 10ms-28ms-2 9ms-2 10ms-2 12ms-2 14ms-2 16ms-2 18ms-2 20ms-216ms-2 18ms-2 20ms-2 25ms-2 30ms-2 40ms-2 50ms-2 60ms-230ms-2 40ms-2 50ms-2 60ms-2 80ms-2 100ms-2 8 J. M. Goyal et al. atmosphere). Optical depth, τeq, is set to 0.56 (Lecavelier Des Etangs et al. 2008). We note that we do not use Eq. 2 to calculate the transmission spectra in our model grid, but use the numerical approach detailed in Amundsen (2015); Goyal et al. (2018). In our model grid we present a large number of trans- mission spectra for specific sets of model parameters, where, (λ) = zgrid(λ) + Rgrid Rgrid , as shown in Eq. 1. However, we p derive a scaling relation by solving Eq. 2 simultaneously for grid and planet parameters, to fine-tune a particular model from the grid to a specific set of planetary parameters, p,bulk zgrid(λ) Hgrid − zplanet(λ) H planet = ln αgrid − ln αplanet (3) where terms denoted "grid" are values from the model grid and terms denoted "planet" are the new parameters for a specific case. Rearranging Eq. 3 and canceling constants we obtain the scaling relation, zplanet(λ) = zgrid(λ)T planet ggrid T gridgplanet −0.5 ln Rgrid p,bulkT planet gplanet Rplanet p,bulk T gridggrid (4) . Importantly, we note that we have made the assumption that σabs, ξabs and µ are constants, while scaling from the nearest grid to planetary parameter, which is a reasonable assumption given the fine variation of parameters in the grid, as demonstrated in Section 5.2 The wavelength dependent planetary radius (Rp(λ)) scaled to the parameters of a specific planet can then be found using Eq. 1 and Eq. 4. The transmission spectrum can then be obtained simply by including the rele- vant stellar radius R∗. We provide a python code on GitHub and the grid google drive to scale these models3,4. (cid:16) Rp(λ) (cid:17) R∗ 5.2 Comparison to Planetary Specific Grid We test this new scalable ATMO grid by comparing it to the previously published planetary specific transmission spectra (Goyal et al. 2018)5. WASP-39b has Teq = 1116 K, gp = 4.07 ms−2, Rp = 1.27 RJ, and R∗ = 0.90 Rsun adopted from TEPCAT database (Southworth 2011). As shown in Wakeford et al. (2018), the best fit model for WASP-39b data from the Goyal et al. (2018) grid has a metallicity of 100× solar, uniform cloud = 0.2, and scattering haze = 150×. To demonstrate the applicability of our generic model spectra grid for different planetary systems, we scale to the planetary parameters of WASP-39b, and compare to this best fit model in Fig. 5. From the generic grid we scale the Teq = 1100 K, gp = 5.0 ms−2, 200× solar, uniform cloud = 0.2, and scattering haze = 150× model to the best fit model parameters for WASP-39b. For this temperature range the spectrum is relatively independent of the treatment of the 3 https://github.com/hrwakeford/Generic_Grid 4 https://drive.google.com/open?id=1ZFbkPdqg37_ Om7ECSspSpEp5QrUMfA9J 5 https://bd-server.astro.ex.ac.uk/exoplanets/ Figure 5. Figure showing the use of the generic grid (red) to fit atmospheric data (black points, Wakeford et al. 2018), in com- parison to the planetary forward model grid (blue) presented in Goyal et al. (2018). In both cases we use the grid for local con- densation only. condensation, so for clarity we only show the local conden- sation spectrum in Fig. 5. This test demonstrates the flexibility of this grid of models to be adapted to specific planetary parameters. There are minor differences, ∼100ppm, in the IR near 4.5 microns with the scaled CO/CO2 feature, however, these do not significantly affect the fit to the data. In the rest of the spectra, the differences between the models average 50 ppm, which is well below the data uncertanties. The minor differ- ences in the Rayleigh scattering slope can be attributed to differences in adopted pressure broadening profiles for Na in this work and that in (Goyal et al. 2018), explained in detail in Section 2. This test also demonstrates that differences shown in Fig. 4 when scaling the 5 ms−2 model to new planetary pa- rameters are negligible between specifically generated mod- els and our re-scaled generic models, where differences are well within the constraints of the current observational mea- surements. Compared to the planetary specific grid presented in (Goyal et al. 2018) this grid can be applied to a wide range of H2/He dominated exoplanet atmospheres and is not lim- ited to the 117 well studied planets. This grid is applicable to targets detected using TESS, NGTS, HATS and any number of other H2/He dominated planets discovered in the future over a wide parameter space. Due to the scalable nature of the grid it can also be implemented within a retrieval frame- work, for which the planetary specific grid is not suitable. 5.3 Comparison to other Forward Models Figure 6 shows the "scaled" simulated spectrum from the generic exoplanet ATMO forward model grid, with similar generic forward model simulations from Fortney et al. (2008, 2010) and Exo-transmit from Kempton et al. (2017). Each simulation is computed for isothermal P-T profiles assum- ing thermochemical equilibrium. The simulations in Fig. 6 are computed for solar metallicity atmospheres without cloud opacities and Teq = 600 K (top), Teq = 1500 K (bot- tom), with gp = 10 ms−2, Rp = 1 RJ, and R∗ = 1 Rsun. Both Exo-transmit (Kempton et al. 2017) and the models based on Fortney et al. (2008, 2010), included thermochemical equilibrium scheme which account for condensation (Lod- MNRAS 000, 1 -- 12 (2018) ATMO WASP-39b Specific Model ATMO Scaled Generic Model Model Params. R* = 0.90 Rsun, Rp = 1.27 RJ Teq = 1116K, gp = 4.07 ms-2, [M/H] = 2.3, cloud = 0.2, haze = x150. Generic Grid of Transmission Spectra 9 et al. (2017) models primarily use HITRAN line-lists. This can lead to differences in spectral features. 6 TRANSMISSION SPECTRAL INDEX To evaluate the impact that changing individual scaling parameters has on the atmospheric transmission spectrum we use the grid of generic models to compute the differ- ent transmission spectral indices as detailed in (Sing et al. 2016). We compute spectral indices for both local conden- sation and rainout condensation grids, and measure the dif- ference in the radius ratio with an increase in atmospheric metallicity, uniform cloud scattering, and wavelength depen- dent haze scattering. We compute the H2O amplitude versus ∆ZU B−L M/Heq index for both the grids and show the model trends in Fig. 7. For each model, the H2O amplitude is computed based on the radius ratio of each model between 1.34 -- 1.49 µm rel- ative to the standard solar model over the same wavelength range. The relative change in the radius ratio for each model is then converted to a percentage amplitude of the H2O fea- ture compared to solar. The ∆zU B−L M/Heq index represents the measured radius ratio of the model in the blue-optical UB-band (0.3 -- 0.57 µm) compared to the mid-infrared LM- band (3 -- 5 µm) in terms of atmospheric scale height. A neg- ative number here indicates that the blue-optical is lower in altitude than the mid-IR, with the opposite true for positive numbers. We show the model trends for the uniform cloud, scattering haze, and super solar metallicity parameters cov- ered by both grids (Fig. 7). The transmission spectral indices show that the dif- ferences between local condensation and rainout condensa- tion are predominantly in the ∆zU B−L M/Heq index, where ∆zU B−L M/Heq = -0.8 and -1.75 for the clear solar model for local condensation and rainout condensation respectively. In all model cases the H2O amplitude is decreased, with the ex- ception of the 10× solar model which would require the op- tical and mid-IR data to distinguish from a clear solar case. There is a distinct separation between each of the model tracks. Increasing the metallicity decreases the amplitude of the water feature while increasing the radius ratio in the mid-IR compared to the optical, mainly due to the presence of CO2 at higher metallicities (see Fig. 1). Increasing the uniform cloud parameter decreases the H2O amplitude with little change to the relative radius ratio between the optical and mid-IR. Increasing the wavelength dependent scatter- ing haze decreases the amplitude of the H2O feature while increasing the relative radius ratio between the optical and mid-IR. We note that these computed transmission spectral tracks show the trend of the individual parameters com- pared to a clear solar abundance atmosphere and for exam- ple, do not depict the combined impact of scattering haze and high metallicity. However, these are a useful indication of dominant factors in the transmission spectra of exoplan- ets and have successfully been used to make predictions for atmospheric measurements (Sing et al. 2016; Wakeford et al. 2018). Each of the measurements shown on the diagram are from Sing et al. (2016) with the addition of WASP-39b from Wakeford et al. (2018). A majority of the currently mea- sured transmission spectra from the optical to the mid-IR Figure 6. Figure showing comparison of ATMO rainout condensa- tion model (red) to the Fortney et al. (2010) equilibrium model with TiO/VO (blue), and Exo-transmit with condensation and rainout (yellow). Each are isothermal models for gp=25 ms−2, Rp/R∗=0.1, and Teq= 600 K (top) and 1500 K (bottom). We show the spectra aligned in the infrared 1.4 µm. The main differences in the infrared are due to the opacity database used for H2O and CH4. Bottom: In the optical, Exo-transmit and the Fortney et al. (2010) show evidence for VO absorption while this is removed in the rainout process in ATMO. ders 1999, 2002, 2006; Lodders & Fegley 2002, 2006; Visscher et al. 2006; Freedman et al. 2008). We see that all three models agree well at low tem- perature (Teq = 600 K) (with a slight difference in the Ex- oTransmit model between 1 -- 2 µm). The difference in the relative radius ratio in the optical slopes at Teq = 1500 K between models from Fortney et al. (2010) and ATMO, can be attributed to raining out of TiO/VO in ATMO at these temperatures. Here we discuss some of the reasons for differences be- tween these models as seen in Fig. 6. Both Exo-transmit (Kempton et al. 2017) and the models based on Fortney et al. (2008, 2010) use elemental abundances from Lodders (2003). However, ATMO uses elemental abundances from As- plund et al. (2009). The major difference between these two sources is in Helium abundance which is greater in Lodders (2003) compared to Asplund et al. (2009). This can lead to some differences in equilibrium chemical abundances. The differences in adopted polynomial coefficients for chemical equilibrium calculations can also lead to differences. In Fort- ney et al. (2010) the base radius is either at 10 or 100 bar, while in this work we adopt 10 bar. The grid generated us- ing ATMO uses high temperature line-lists primarily from Exomol with H2/He pressure broadening applied wherever possible. In comparison, Fortney et al. (2010) and Kempton MNRAS 000, 1 -- 12 (2018) ATMO Local Condensation ATMO Rainout Condensation Fortney et al. (2010) with TiO/VO Fortney et al. (2010) without TiO/VOATMO Local Condensation ATMO Rainout Condensation Fortney et al. (2010) with TiO/VO Fortney et al. (2010) without TiO/VOAligned in the opticalAligned in the infraredATMO Rainout Condensation Fortney et al. (2010) eq. chem. with TiO/VO Exo-transmit with condensation and rainoutTeq = 1500 KTeq = 600 KATMO RainoutFortney et al. (2010) eq. chem. Exo-Transmit with Rainout 10 J. M. Goyal et al. Figure 7. Figure showing ATMO grid based transmission spectral index tracks first (e.g. Sing et al. (2016)). We show the H2O amplitude versus the ∆zU B−L M index for [M/H] (red), C/O (green), uniform cloud (grey), and scattering haze (purple) parameters. We show the tracks for both the grid with rainout condensation (solid lines) and local condensation (dashed lines). Planetary data points are taken from Sing et al. (2016) and Wakeford et al. (2018). follow the track of increased scattering haze with a majority of measured H2O amplitudes between 20 -- 50% of the clear solar model. These indices demonstrate the importance of the optical and mid-IR data to distinguish between different model parameters, in this case the cloud, haze, and enhanced metallicity trends. 7 CONCLUSIONS We present a publicly available6 generic forward model grid of exoplanet transmission spectra, computed using the ATMO code, based on both local condensation and rainout con- densation conditions. This grid can be scaled to a wide 6 https://drive.google.com/open?id=1ZFbkPdqg37_ Om7ECSspSpEp5QrUMfA9J range of H2/He dominated planetary atmospheres. The en- tire grid consists of 56,320 model simulations across 22 isothermal temperatures, four planetary gravities, five atmo- spheric metallicities, four C/O ratios, four uniform cloud pa- rameters, four scattering haze parameters, and two chemical condensation scenarios. We demonstrate by sensitivity anal- ysis, the reasoning behind selection of different grid parame- ters and their values. We derive scaling equations which can be used with this grid, for a wide range of planet-star com- bination. This grid of forward models is validated against published models of Fortney et al. (2010), Kempton et al. (2017) and Goyal et al. (2018). Degeneracies are known to exist in the interpretation of various characteristics of planetary atmospheres from obser- vations. The grid of atmospheric models presented here al- lows us to decouple and better understand the thermochem- ical processes shaping observable spectra. We demonstrate how changes in temperature and chemical abundances cause MNRAS 000, 1 -- 12 (2018) WASP-39bHD 209458bWASP-31bWASP-17bWASP-19bHAT-P-1bWASP-12bHAT-P-12bHD 189733bCloudSuper SolarScatteringHazeRainout CondensationLocal CondensationClear Solarx200x100x50x101.00.20.06x1100x150x10420246ZUBLM/Heq20020406080100120H2Oamplitude(%)Rainout CondensationLocal CondensationCloud[M/H]ScatteringHaze1.0C/Ox1100x150x100.20.06Clear SolarWASP-39bHD 209458bWASP-31bWASP-17bWASP-19bHAT-P-1bWASP-12bHAT-P-12bHD 189733bx200x100x50x100.350.560.71.0 different trends in the spectra, and identify the following ma- jor factors that can effect interpretation of observations. • Adopting different condensation processes (rainout and local) can lead to different interpretation of observations. • SO2 features at 6 -- 8µm, along with H2O, can be used to constrain the metallicity of the exoplanet atmosphere, since the SO2 spectral feature only appears for metallcities greater than 100x solar. • The presence of VO without TiO can help constrain the temperature of the atmospheric limb, and that both TiO/VO features can reveal dominant physical process (rainout or local condensation) in the planet's atmosphere. • At high C/O ratios (∼1), spectral features in the in- frared are different between the rainout and local conden- sation case, because rainout case has higher abundance of carbon bearing species without oxygen such as CH4, C2H2 and HCN. • The difference in solar elemental abundances between Asplund et al. (2009) and Lodders (2003) used for model initialisation, can lead to differences in equilibrium chemical abundances and therefore the spectral features. This grid can be used to interpret the observations of H2/He dominated hot jupiter exoplanet atmospheres, as well as to plan future observations using the HST, VLT, JWST and various other telescopes. It can be used directly with the HST and JWST simulator PandExo (Batalha et al. 2017) for planning observations. The scaling flexibility provided by this grid for a wide range of planet-star combinations will be extremely valuable to efficiently choose and plan observa- tions, for soon to be discovered TESS targets.The fine vari- ation of parameters in the grid also allows it to be incorpo- rated in a retrieval framework with various machine learning techniques, as demonstrated in Marquez-Neila et al. (2018). ACKNOWLEDGEMENTS We would like to thank the anonymous reviewer for their constructive comments that improved the paper substan- tially. We would like to thank Mark Marley and Channon Visscher for valuable discussions on rainout condensation and comparative studies to their chemical models. J.M.G and N.M are part funded by a Leverhulme Trust Research Project Grant, and in part by a University of Exeter Col- lege of Engineering, Mathematics and Physical Sciences PhD studentship. H.R.W acknowledges funding by Association of Universities for Research in Astronomy (AURA) through a Giacconi Fellowship appointed at Space Telescope Sci- ence Institute (STScI). D.K.S and B.D acknowledge support from the European Research Council under the European Unions Seventh Framework Programme (FP7/2007-2013)/ ERC grant agreement number 336792. We acknowledge sup- port from the Space Telescope Science Institute (STScI) Re- search Visitors Program. This work used the DiRAC Complexity system, oper- ated by the University of Leicester IT Services, which forms part of the STFC DiRAC HPC Facility. This work also used the University of Exeter Supercomputer, a DiRAC Facility jointly funded by STFC, the Large Facilities Capital Fund of BIS and the University of Exeter. MNRAS 000, 1 -- 12 (2018) Generic Grid of Transmission Spectra 11 Author Contributions J.M.G computed the model grids. J.M.G, B.D, and H.R.W benchmarked the models with the help of N.K.L, D.K.S and N.M. J.M.G and H.R.W wrote the manuscript. H.R.W made the figures. All authors commented on and approved the final manuscript. REFERENCES Allard N. F., Allard F., Hauschildt P. H., Kielkopf J. F., Machin L., 2003, A&A, 411, L473 Amundsen D. S., 2015, PhD thesis Amundsen D. S., Baraffe I., Tremblin P., Manners J., Hayek W., Mayne N. J., Acreman D. M., 2014, A&A, 564, A59 Asplund M., Grevesse N., Sauval A. J., Scott P., 2009, ARA&A, 47, 481 Batalha N. E., et al., 2017, preprint, (arXiv:1702.01820) Burrows A., Sharp C. M., 1999, ApJ, 512, 843 Burrows A., Marley M. S., Sharp C. M., 2000, ApJ, 531, 438 Deming D., et al., 2013, ApJ, 774, 95 Drummond B., 2017, PhD thesis, University of Exeter Drummond B., Tremblin P., Baraffe I., Amundsen D. S., Mayne N. J., Venot O., Goyal J., 2016, A&A, 594, A69 Evans T. M., et al., 2016, ApJ, 822, L4 Evans T. M., et al., 2017, Nature, 548, 58 Fortney J. J., 2005, MNRAS, 364, 649 Fortney J. J., Lodders K., Marley M. S., Freedman R. S., 2008, ApJ, 678, 1419 Fortney J. J., Shabram M., Showman A. P., Lian Y., Freedman R. S., Marley M. S., Lewis N. K., 2010, ApJ, 709, 1396 Fortney J. J., et al., 2016, preprint, (arXiv:1602.06305) Freedman R. S., Marley M. S., Lodders K., 2008, ApJS, 174, 504 Goyal J. M., et al., 2018, MNRAS, 474, 5158 Heng K., Kitzmann D., 2017, preprint, (arXiv:1702.02051) Kempton E. M.-R., Lupu R., Owusu-Asare A., Slough P., Cale B., 2017, PASP, 129, 044402 Kreidberg L., et al., 2014, ApJ, 793, L27 Lecavelier Des Etangs A., Pont F., Vidal-Madjar A., Sing D., 2008, A&A, 481, L83 Lodders K., 1999, ApJ, 519, 793 Lodders K., 2002, ApJ, 577, 974 Lodders K., 2003, ApJ, 591, 1220 Lodders K., 2006, ApJ, 647, L37 Lodders K., Fegley B., 2002, Icarus, 155, 393 Lodders K., Fegley Jr. B., 2006, Chemistry of Low Mass Substellar Objects. p. 1, doi:10.1007/3-540-30313-8 1 Madhusudhan N., 2012, ApJ, 758, 36 Marquez-Neila P., Fisher C., Sznitman R., Heng K., 2018, preprint, p. arXiv:1806.03944 (arXiv:1806.03944) Mbarek R., Kempton E. M.-R., 2016, ApJ, 827, 121 Molli`ere P., van Boekel R., Dullemond C., Henning T., Mordasini C., 2015, ApJ, 813, 47 Molli`ere P., van Boekel R., Bouwman J., Henning T., Lagage P.- O., Min M., 2016, preprint, (arXiv:1611.08608) Moses J. I., et al., 2013, ApJ, 777, 34 Nikolov N., et al., 2018, Nature, 557, 526 Rothman L. S., et al., 2010, J. Quant. Spectrosc. Radiative Trans- fer, 111, 2139 Rothman L. S., et al., 2013, J. Quant. Spectrosc. Radiative Trans- fer, 130, 4 Sing D. K., et al., 2016, Nature, 529, 59 Southworth J., 2011, MNRAS, 417, 2166 Tennyson J., et al., 2016, Journal of Molecular Spectroscopy, 327, 73 Tremblin P., Amundsen D. S., Mourier P., Baraffe I., Chabrier G., Drummond B., Homeier D., Venot O., 2015, ApJ, 804, L17 12 J. M. Goyal et al. Tremblin P., Amundsen D. S., Chabrier G., Baraffe I., Drummond B., Hinkley S., Mourier P., Venot O., 2016, ApJ, 817, L19 Tremblin P., et al., 2017, ApJ, 841, 30 Vandaele A. C., et al., 2017, Icarus, 295, 1 Visscher C., Lodders K., Fegley Jr. B., 2006, ApJ, 648, 1181 Wakeford H. R., Sing D. K., Evans T., Deming D., Mandell A., 2016, ApJ, 819, 10 Wakeford H. R., et al., 2018, AJ, 155 This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1 -- 12 (2018)
1901.07078
1
1901
2019-01-21T21:01:15
Transitional disk archeology from exoplanet population synthesis
[ "astro-ph.EP", "astro-ph.SR" ]
Increasingly better observations of resolved protoplanetary disks show a wide range of conditions in which planets can be formed. Many transitional disks show gaps in their radial density structure, which are usually interpreted as signatures of planets. It has also been suggested that observed inhomogeneities in transitional disks are indicative of dust traps which may help the process of planet formation. However, it is yet to be seen if the configuration of fully evolved exoplanetary systems can yield information about the later stages of their primordial disks. We use synthetic exoplanet population data from Monte Carlo simulations of systems forming under different density perturbation conditions, which are based on current observations of transitional disks. The simulations use a core instability, oligarchic growth, dust trap analytical model that has been benchmarked against exoplanetary populations.
astro-ph.EP
astro-ph
Origins: from the Protosun to the First Steps of Life Proceedings IAU Symposium No. 345, 2019 Bruce G. Elmegreen, L. Viktor T´oth, Manuel Gudel, eds. c(cid:13) 2019 International Astronomical Union DOI: 00.0000/X000000000000000X Transitional disk archeology from exoplanet population synthesis Germ´an Chaparro Molano1, Frank Bautista2, Yamila Miguel3 1Vicerrector´ıa de Investigaci´on, Universidad ECCI Bogot´a, Colombia email: [email protected] 2Departamento de F´ısica, Universidad Nacional de Colombia Bogot´a, Colombia email: [email protected] 3Sterrewacht Leiden, Leiden University Leiden, The Netherlands Abstract. Increasingly better observations of resolved protoplanetary disks show a wide range of conditions in which planets can be formed. Many transitional disks show gaps in their radial density structure, which are usually interpreted as signatures of planets. It has also been sug- gested that observed inhomogeneities in transitional disks are indicative of dust traps which may help the process of planet formation. However, it is yet to be seen if the configuration of fully evolved exoplanetary systems can yield information about the later stages of their primordial disks. We use synthetic exoplanet population data from Monte Carlo simulations of systems forming under different density perturbation conditions, which are based on current observa- tions of transitional disks. The simulations use a core instability, oligarchic growth, dust trap analytical model that has been benchmarked against exoplanetary populations. Keywords. planetary systems: formation, protoplanetary disks 1. Introduction Planet-forming disks can have either smooth or density perturbed profiles (van der Marel et al. 2015). Gas-depleted, density perturbed disks are often called transitional disks. Gaps, cavities, or radial perturbations in such disks sometimes show dust traps in which planet formation may be taking place. Such dust traps may be caused by pressure bumps, which are often theorized as a consequence of the presence of an already-formed planet in the disk (Pinilla et al. 2011). In this work, we are interested in the impact of disk density perturbations in planetary populations while making no assumptions about the exact mechanism that causes such inhomogeneities in the disk. In order to explore the link between transitional disks and exoplanetary systems, we intend to compare synthetic and observed populations of exoplanetary systems. However, instead of looking at individual cases (Raymond et al. 2018) we take a Bayesian inference approach to reconstruct probability distributions of general properties of 3000+ simulated synthetic planetary systems. We thus study the effect of radial density perturbations in the disk structure on the formation of exoplanetary systems. For many transitional disks, a radial density perturbation can be described as a succession of over-dense and under-dense regions which appear as the radial distance r changes (Pinilla et al. 2011), (cid:18) (cid:18) (cid:19)(cid:19) r f H(r) Σp(r) = Σ(r) 1 + A cos 2π 1 . (1.1) 9 1 0 2 n a J 1 2 . ] P E h p - o r t s a [ 1 v 8 7 0 7 0 . 1 0 9 1 : v i X r a 2 Germ´an Chaparro Molano Here Σ(r) is the surface density distribution (Miguel et al. 2011), A is the amplitude of the perturbation, f is the length scale of the perturbation and H(r) is the scale height of the disk. Here we consider A = 0 for smooth disks, and A = 0.3 for transitional disks. 2. Synthetic planetary systems Population synthesis models are often used for modeling individual exoplanetary sys- tems (Raymond et al. 2018) in order to estimate properties of individual planets in the system. We extend this method for 3000 synthetic planetary systems formed in smooth and transitional disks (following the perturbation recipe described above). Each system has initial conditions drawn from prior probability distributions on stellar mass, disk mass and radial extent, stability, metallicity, and gas dissipation timescale from Miguel et al. (2011). This planet population synthesis framework is also described as part of the review in Benz et al. (2014). The result of each simulation is a system with orbital data like planetary semi-major axis and mass (solid+gas). We calculate consolidated quanti- ties that summarize general properties of each simulated system. Thus, we use quantities like the number of terrestrial and giant planets, total terrestrial planetary mass, average terrestrial planet mass, total planetary mass/disk mass ratio, and what we refer here to as center of mass. This center of mass is not the barycenter of the system but rather the first moment of the mass distribution of planets with respect to their semi-major axes. With the consolidated results per system, we can reconstruct the posterior probability distribution from the simulated quantities mentioned above. This reconstructed poste- rior can be used to make educated predictions for existing exoplanetary systems based on known properties like stellar mass, metallicity, planetary masses and semi-major axes distributions, etc. 3. Results As an initial benchmark, we looked at the resulting distributions of the location of the center of mass and the total planetary mass/disk mass ratio for systems formed in smooth disks. Figure 1(a) shows that systems with giant planets are more spread inwards (closer to the star) than systems that only formed terrestrial planets, due to giant planet migration. On the other hand, more of the original disk mass goes to form planets in systems with giant planets than in terrestrial planet-only systems (Figure 1(b)). Both of these results are expected from exoplanet and planet-forming disk observations. The resulting synthetic planetary systems were divided among systems with giant planets (20-25% of all systems), and systems with terrestrial planets only. We benchmarked our results of the distribution of planetary systems with respect to metallicity. Figure 2), shows that systems with giant planets tend to be formed in metal-rich systems, both in our simulations and in observations from exoplanet.eu. Comparatively, most exoplanetary systems discovered to date tend to be distributed uniformly around the solar value in the HARPS 2011 catalog (Mayor et al. 2011), which we used as a prior for our simulations. For the same initial disk mass, transitional disks form more systems with giant planets than smooth disks (Figure 3). This is due to transitional disks having over-dense regions that favor planet growth. Figure 4 shows the distribution of stellar mass vs. center of mass for synthetic and observed planetary systems. The 1σ and 2σ contours show that there is a significant overlap in parameter space between synthetic and observed planetary systems. For exoplanetary systems with a low center of mass the overlap breaks down, which is likely a result of observational bias. Transitional disk archeology from exoplanet population synthesis 3 (a) (b) Figure 1. Marginalized distributions of center of mass and total planet mass/disk mass ratio for systems with giant planets (green) and with only terrestrial planets (blue). The distributions are approximated by a Kernel Density Estimation (solid lines). Figure 2. Marginalized distribution of exoplanetary systems metallicities for synthetic systems with giant planets (green), observed systems with giant planets (red) and from the prior used in the simulation (blue) from Mayor et al. (2011). The distributions are approximated by a Kernel Density Estimation (solid lines). 4. Conclusions Our comparison between smooth and transitional disks shows that transitional disks favor the formation of giant planets at lower disk masses than smooth disks. Benchmark- ing of the results of our simulations show a very good parameter space overlap between synthetic and observed exoplanetary systems. The method of approximating a posterior probability distribution for exoplanetary parameters from our simulations can be used to make educated predictions that direct future surveys of observed exoplanetary systems. References Benz, W., Ida, S., Alibert, Y., Lin D.N.C., & Mordasini, C. 2014, in H. Beuther, R.S. Klessen, C.P. Dullemond, & T. Henning (eds.) Protostars and Planets VI, Univ. of Arizona Press, 944 Hughes, A.M., Duchene, G., & Matthews B.C. 2018, Annu. Rev. Astron. Astrophys., 56, 541 Mayor G.M., Marmier M., Lovis, C., et al. 2011, arXiv1109.2497 Miguel, Y., Guilera, M., & Buruni, A. 2011, MNRAS, 417, 314 Pinilla, P., Birnstiel, T., Ricci, L., Dullemond, C.P., Uribe, A.L., Testi, L., & Natta, A. 2012, A&A, 538, A114 4 Germ´an Chaparro Molano Figure 3. Marginalized posterior distribution of parent disk masses for synthetic systems with giant planets for smooth and transitional disks. The dashed black lines show lower and upper estimations for the minimum mass solar nebula. The distributions are approximated by a Kernel Density Estimation (solid lines). Figure 4. Center of mass vs. stellar mass for observed systems (red) and synthetic systems (blue) formed in smooth disks (left) and in transitional disks (right). The 1σ (solid line) amd 2σ (dashed line) contours were obtained using a Kernel Density Estimation for each set of data points. Pinilla, P., Tazzari, M., Pascucci, I., et al. 2018, ApJ, 859, 32 Pinilla, P., van der Marel, N., P´erez, L.M., van Dishoeck, E.F., Andrews, S., Birnstiel, T., Herczeg, G., Pontoppidan, K.M., & van Kempen, T. 2015, A&A, 584, A16 Raymond, S.N., Boulet, T., Izidoro, A., Esteves, L., & Bitsch, B. 2018, MNRAS, 479, L81 van der Marel, N., van Dishoeck, E.F., Bruderer, S., Andrews, S.M., Pontoppidan, K.M., Her- czeg, G.J., van Kempen, T. & Miotello, A. 2016, A&A, 585, A58 van der Marel N., van Dishoeck E.F., Bruderer, S., P´erez, L., & Isella, A. 2015, A&A, 579, A106 Discussion Alfaro: What kind of priors are you using in the Bayesian inference methods? Transitional disk archeology from exoplanet population synthesis 5 Chaparro: The functional form of priors for each input parameter are taken from the literature, which are summarized in the 2011 work by Miguel et al.
1803.06614
1
1803
2018-03-18T07:01:23
On existence of out-of-plane equilibrium points in restricted three-body problem with oblateness
[ "astro-ph.EP" ]
We analyze in this paper the existence of the "out-of-plane" equilibrium points in the restricted three-body problem with oblateness. From the series expansion of the potential function of an oblate asteroid, we show analytically all equilibrium points locate on the orbital plane of primaries and how artificial equilibrium points may arise due to an inappropriate application of the potential function. Using the closed form of the potential of a triaxial ellipsoid, we analytically demonstrate that the gravitational acceleration in $z$-direction is always pointing toward the equatorial plane, thus it could not be balanced out at any value of $z\neq 0$ and the out-of-plane equilibrium points cannot exist. The out-of-plane equilibrium points appear only when additional acceleration other than the gravitation from primaries is taken into account. We suggest that special attention must be paid to the application of the spherical harmonics expansion of potential to find the equilibrium points, especially when these points may be very close to the celestial body.
astro-ph.EP
astro-ph
On existence of out-of-plane equilibrium points in restricted three-body problem with oblateness Xuefeng Wang, Nan Wu, Liyong Zhou and Bo Xu School of Astronomy and Space Science, Nanjing University, Nanjing 210046, China Key Laboratory of Modern Astronomy and Astrophysics in Ministry of Education, Nanjing University [email protected] ABSTRACT We analyze in this paper the existence of the "out-of-plane" equilibrium points in the restricted three-body problem with oblateness. From the series expansion of the potential function of an oblate asteroid, we show analytically all equilibrium points lo- cate on the orbital plane of primaries and how artificial equilibrium points may arise due to an inappropriate application of the potential function. Using the closed form of the potential of a triaxial ellipsoid, we analytically demonstrate that the gravitational acceleration in z-direction is always pointing toward the equatorial plane, thus it could not be balanced out at any value of z 6= 0 and the out-of-plane equilibrium points cannot exist. The out-of-plane equilibrium points appear only when additional acceler- ation other than the gravitation from primaries is taken into account. We suggest that special attention must be paid to the application of the spherical harmonics expansion of potential to find the equilibrium points, especially when these points may be very close to the celestial body. Subject headings: Celestial Mechanics – minor planets, asteroids: general – methods: analytical 1. Introduction To find the equilibrium points and their stabilities is a key step to understand the structure of phase space of a dynamical system as well as the orbital behaviours. In the restricted three-body problem concerning the motion of a small (celestial or artificial) object in the gravitational field of two massive primaries, the equilibrium points suggest the existence of temporary (unstable) or permanent (stable) periodic trajectories around the primaries, which not only provide explanations of the orbital configuration of celestial bodies but also benefit the trajectory design for space missions. In the classical restricted three-body problem, it is well-known that there are five equilib- rium points, i.e. three collinear points (L1, L2, L3) and two triangular points (L4, L5), all on the – 2 – orbital plane of the primaries (see e.g. Murray & Dermott 1999). In the perturbed restricted three-body problem, however, these so called Lagrangian points (L1∼5) still exist but with dif- ferent locations and stabilities. Detailed studies of their dynamical features can be found e.g. in Abouelmagd & Guirao (2016); Elshaboury et al (2016). With the effect of radiation pressure, apart from these five coplanar ones, other equilibrium points than L1 ∼ L5 are found and ana- lyzed in literatures, e.g. Simmons, McDonald & Brown (1985); Ragos & Zagouras (1988); Todoran (1994); Roman (2001); Ansari (2017). When the oblateness of primaries (one or both) is taken into account, the existence of out-of-plane equilibrium points, locating out of the orbital plane of primaries, is also declared (Douskos & Markellos 2006), and the dynamics of orbits around these new equilibrium points have been discussed in some other literatures. However, the existence of such out-of-plane equilibrium points in the latter case is doubtful. For example, Zamaro & Biggs (2015) found in their calculation that such additional equilibrium points in the Mars-Phobos system locate beneath the surface of the Martian moon Phobos, making them physically meaningless. Intuitively, at any point out of the orbital plane, each (oblate) primary's gravitational force on the third body has a component along the z direction, and both of them point to the orbital plane that in general coincides with the primary's equatorial plane. Not like in the photogravitatioanl system where the radiation pressure offers a force opposite to the gravitation, there is no other force in this "oblate-primary" system that can balance out the combined gravitation in z direction. Thus no force balance can be attained only except for on the orbital plane (z = 0). We will check again in this paper where all the equilibrium points locate in the restricted three-body problem with oblateness, and particularly, we will find out whether we should expect to see equilibrium points out of the orbital plane. In Section 2, we will briefly review the potential function of an oblate homogenous body and the calculation of the equilibrium points, as well as show how artificial out-of-plane equilibrium points may arise due to an inappropriate employment of the potential function. In Section 3, a more general case where one or both of the primaries are triaxial ellipsoids is investigated, and we demonstrate that equilibrium points must locate on the orbital plane because the force balance can only be attained on the same plane. And we discuss the condition for the existence of out-of-plane equilibrium points and draw conclusion in Section 4. 2. Motion under the truncated potential 2.1. Background of the potential function The expansion of a non-spherical body's gravitational potential can be found in many textbooks of celestial mechanics, e.g. Szebehely (1967); Murray & Dermott (1999). Here we briefly review the perturbation generated by the J2 term for convenience. Assume a homogenous body M of – 3 – arbitrary shape and a coordinate frame centered in the center of mass, as sketched in Fig. 1. The gravitational potential of M at any point P out of the body can be described as an integral Expand 1/∆ using Legendre polynomials, we get V =Z Gdm ∆ =Z Gdm (r2 + R2 − 2rR cos α)1/2 . (1) (2) V = = G r Z dm r Z "1 + G ∞ r(cid:19)n Xn=0(cid:18) R R r cos α + Pn(cos α) r(cid:19)2 2(cid:18) R 1 (cid:0)3 cos2 α − 1(cid:1) + ···# dm where Pn is the Legendre polynomials. The n = 1 term R cos α/r and n = 2 term (R2/r2)(cid:0)3 cos2 α − 1(cid:1) /2 respectively are just the J1 and J2 zonal terms in the spherical harmonics expansion of the gravi- tation potential. We now simplify the model by supposing that M is a triaxial ellipsoid with three Fig. 1.- Gravitation potential created by a mass element dm on an exterior point P . axes Rx, Ry, Rz along the OX, OY, OZ axes and the moments of inertia about these three axes are Ix, Iy, Iz. Because the origin of the Cartesian coordinate system O − XY Z coincides with the center of mass, we have Z dm = M, Z R cos αdm = 0, Z R2dm = 1 2 (Ix + Iy + Iz) . The potential V now reads V = GM r + G(Ix + Iy + Iz) 2r3 − 3G 2r3 Z R2 sin2 αdm + ··· . From Fig. 1, we know that R R2 sin2 αdm = IOP is the moment of inertia of M about axis OP while the direction cosine of OP is (x/r, y/r, z/r), thus IOP = (cid:0)Ixx2 + Iyy2 + Izz2(cid:1) /r2. So now, – 4 – V becomes V = GM r + G(Ix + Iy + Iz) 2r3 3G(cid:0)Ixx2 + Iyy2 + Izz2(cid:1) 2r5 − + ··· . When M is a rotational ellipsoid with equatorial and polar radii Re and Rp, we have Ie ≡ Ix = Iy = 1 5(cid:0)R2 e + R2 p(cid:1) M, Ip ≡ Iz = 2 5 R2 eM. If M is an oblate ellipsoid, then Re > Rp and Ie < Ip. Consider the equation: Ixx2 + Iyy2 + Izz2 = Ie(r2 − z2) + Ipz2 = (Ip − Ie)z2 + Ier2, then V in Eq. (3), truncated up to J2 term, can be rewritten as V ≈ GM r + GM(cid:0)R2 e − R2 p(cid:1) 10r3 (cid:18)1 − 3z2 r2 (cid:19) . (3) (4) This potential function is exactly the one widely employed in literatures such as Sharma & Subba Rao (1976); Oberti & Vienne (2003); Perdiou, Markellos & Douskos (2005); Douskos & Markellos (2006); Singh & Mohammed (2013); Singh & Umar (2013); Abouelmagd & Mostafa (2015); Song, He & He (2016); Zeng, Baoyin & Li (2016); Zeng, Liu & Li (2017), etc. One must note that the expansion in Eq. (2) only converges when R < r, or in other words, the convergence of spherical harmonics series can be guaranteed only outside the Brillouin sphere, which is the smallest sphere enclosing all mass elements (see e.g. Romain & Jean-Pierre 2001). 2.2. The equilibrium points in the synodic frame Here we will find the equilibrium points in the model of a massless body moving around a non-spherical body with J2 potential term as given in Eq. (4). Then we will extend the analysis to the restricted three-body problem by introducing another point mass in next subsection. In this model, an equilibrium point in a rotational frame corresponds to a circular orbit in the inertial coordinate system. In fact, it is well known (see e.g. Murray & Dermott 1999) that any orbit with nonzero eccentricity and/or nonzero inclination will not be closed thus not correspond to an equilibrium point, because the longitudes of pericentre and node Ω will precess due to the J2 perturbation. Therefore, we focus only on the circular orbit (e = 0) on (or parallel to) the equatorial plane (i = 0◦) hereafter. In the rotational frame (O − xyz) with angular velocity n (see Fig. 2), the equation of motion reads x − 2n y = ∂U ∂x , y + 2n x = ∂U ∂y , z = ∂U ∂z , (5) – 5 – where U = 1 2 n2(x2 + y2) + V and V is the potential function truncated up to the J2 term, as Eq. (4). The stationary solutions to Eq. (5) are just the equilibrium points in the rotational frame and circular periodic orbits in the inertial frame. We find all of these solutions in two cases as follows. Case I: z ≡ 0 It's easy to obtain the following equations: n2x = n2y = GM r3 "1 + r3 "1 + GM 3(R2 3(R2 e − R2 p) 10r2 e − R2 p) 10r2 # x, # y. (6a) (6b) Except for the trivial solution (x, y, z) = (0, 0, 0), countless solutions can be obtained by solving the equation: n2 = GM r3 "1 + 3(R2 e − R2 p) 10r2 # . (7) For any given r (surely r > Re), there exists an n = n(r) that satisfies the above equation. The equilibrium points (corresponds to periodic circular orbits in inertial frame) are countless on the O − xy plane. Fig. 2.- The motion of a massless body under the gravitational attraction from an oblate ellipsoid. O − XY Z is an inertial frame centered at the mass center of M while O − xyz is a rotational frame with angular velocity n. Case II: z 6= 0 The equilibrium equation from Eq. (5) now reads: n2x = GM r3 "1 + 3(R2 e − R2 p) 10r2 (cid:18)1 − 5z2 r2 (cid:19)# x, (8a) – 6 – n2y = z = − 3(R2 e − R2 p) 10r2 GM r3 "1 + e − R2 p) 10r2 3(R2 (cid:18)1 − r2 (cid:19) z. 5z2 (cid:18)3 − 5z2 r2 (cid:19)# y, Obviously, one solution to Eq. (8) is x0 = y0 = 0, z0 = ±r3 5(cid:0)R2 p(cid:1). e − R2 (8b) (8c) (9) At these two points (x, y, z) = (0, 0, z0) on the Oz axis, the gravitation from the main part of the oblate body M is seemingly canceled out by the acceleration generated by the J2 term. However, the distance from this point to the center of M is r0 = z0 < Re, i.e. this force balance point in fact is inside the Brillouin sphere, where the expansion adopted above is invalid. Therefore, these equilibrium points are not true. For x 6= 0 or y 6= 0, since n2 ≥ 0, the terms inside the square brackets in Eq. (8) must satisfy 1 + 3(R2 e − R2 p) 10r2 (cid:18)1 − 5z2 r2 (cid:19) ≥ 0. And from the 3rd equation in Eq. (8), 1 = − 3(R2 e − R2 p) 10r2 (cid:18)3 − 5z2 r2 (cid:19) . Substitute Eq. (11) into Eq. (10), we get 3(R2 e − R2 p) 10r2 −2 · ≥ 0. (10) (11) (12) This inequality is impossible for any rotational ellipsoid (Re > Rp). This contradiction indicates that the "out-of-plane" equilibrium points with z 6= 0 does not exist in this model where the J2 term is taken into account. 2.3. The equilibrium in restricted three-body problem Now we introduce another body (point mass) with mass m on the equatorial plane of M to make a circular restricted three-body problem. Without loss of generality, assume m locates on the Ox axis at a distance D to the first primary M . As usual, the distance between primaries D is set as the length unit, the total mass (M + m) as the mass unit, and the time unit is chosen so as to make the gravitational constant G = 1. Then the oblateness of M can be described by the normalized equatorial and polar radii re = Re/D and rp = Rp/D respectively. After m is introduced, the barycenter of the system is shifted and in the new synodic frame the stationary – 7 – coordinates of M and m are (−µ, 0, 0) and (1 − µ, 0, 0), where µ = m/(m + M ). And the motion of the massless third body can be described by x − 2n y = ∂W ∂x , y + 2n x = ∂W ∂y , z = ∂W ∂z , (13) where W = n2(x2 + y2) 2 + µ r2 + r1 = p(x + µ)2 + y2 + z2, e − r2 r2 5 e − R2 R2 5D2 = A = p p 3z2 r2 A 2r2 1 (cid:18)1 − r1 (cid:20)1 + 1 − µ r2 =p(x + µ − 1)2 + y2 + z2, 1 (cid:19)(cid:21) , . Eq. (13) is the same as Eq. (5) only except that the effective potential function W here has the additional potential µ/r2 from the second primary m. Simple calculation shows that the mean motion rate now is n =p1 + 3A/2. The equilibrium points should be found by solving the equation ∂W/∂x = ∂W/∂y = ∂W/∂z = 0. Since we are mainly interested in the "out-of-plane" equilibrium points in this paper, we focus on the 3rd equation ∂W ∂z = − µ r3 2 z − (1 − µ) r3 1 h1 + 3A 2r2 1 (cid:18)3 − 5z2 r2 1 (cid:19)iz = 0. (14) ∂W/∂z is the acceleration along z direction. It's easy to see from Eq. (14) that the acceleration generated by m and the main part of M , i.e. −(cid:18) 1 − µ r3 1 + µ r3 2(cid:19) z, (15) always points toward the Oxy-plane for z 6= 0, and only the component generated by the J2 term may point outward from the Oxy-plane. When they are balanced, the "out-of-plane" equilibrium points appear. However, as we are going to show below, such points are inside the Brillouin sphere, where the expansion of the potential function is invalid. For z 6= 0, since −µ/r3 2 is definite negative, Eq. (14) can be possibly satisfied if and only if: f (r1, z) = 1 + 3A 2r2 1 (cid:18)3 − 5z2 r2 1 (cid:19) < 0. Because z2 ≤ r2 1, the minimum of f (r1, z) may be attained when r2 1 = z2: min f (r1, z) = 1 + 3A 2r2 1 (cid:18)3 − 5z2 z2 (cid:19) = 1 − 3A r2 1 . Thus, for any (r1, z) of the "out-of-plane" equilibrium point, min f (r1, z) ≤ f (r1, z), i.e. 3A 1 ≤ f (r1, z) < 0. r2 1 − (16) (17) – 8 – Consequently, from this inequality an estimation can be obtained for the out-of-plane equilibrium point: r1 < √3A =s 3(R2 e − R2 p) 5D2 < Re D . (18) Bearing in mind that D is the unit length, Eq. (18) means that r1 < Re, i.e. the equilibrium point locates inside the Brillouin sphere around M . In addition, we show in Fig. 3 numerical calculations of the equilibrium points in two cases with arbitrarily chosen µ = 0.3. The oblate primaries in both examples have the same equatorial radius Re = 0.4, thus they have the same Brillouin sphere with a radius of 0.4. For the polar radius, it's Rp = 0.1 for case (a) and Rp = 0.3 for case (b). Therefore, the oblateness parameter A = 0.03 for (a) and A = 0.014 for (b) respectively. The equilibrium points, denoted by crosses in Fig. 3, are clearly inside the Brillouin sphere in case (a) and even inside the ellipsoid surface in case (b) when A is smaller. Apparently, these numerical results support our conclusion drawn above. It is worth to note that we chose relatively large ellipsoid size and parameter A so that the fact can be seen clearly in the figures. 0.6 0.4 0.2 0.0 z -0.2 -0.4 -0.6 a b z x x -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 Fig. 3.- The artificial out-of-plane equilibrium points are inside the Brillouin sphere or even the ellipsoid. We illustrate the projections on the Oxz-plane of the primary ellipsoids (green), Brillouin spheres (dashed black), the contour of surface ∂W/∂x = 0 (red) and the contour of surface ∂W/∂z = 0 (blue). The equilibrium points at the intersections of red and blue curves are denoted by crosses. As a matter of fact, these out-of-plane equilibrium points arise only due to the existence of acceleration pointing outward from the orbital (equatorial) plane, and we know now that for an oblate primary body such acceleration component is an artificial effect of the improper application – 9 – of the potential series of spherical harmonics. We can also show that this conclusion holds if higher order terms J2n are included. We are going to show in next Section that there is no such outward acceleration at all using the closed form of gravitational potential of a triaxial ellipsoid model. However, if the asymmetry of the gravity (e.g. J3 or tesseral harmonic terms) is considered, there will be the acceleration component pointing outward from the Oxy-plane, and the classical equilibrium points L1−5 perhaps will not stay on the same plane. But in this case, we cannot assume that the two primaries will move on a circular orbit on the primaries' equatorial plane, and this will make a totally different model, out of the scope of this paper. 3. Potential of a triaxial ellipsoid A triaxial ellipsoid is a good model for describing an asteroid if it is not too much distorted, and it severs as one of the basic models in many studies. The gravitational potential of a triaxial ellipsoid will be given by an integral in this Section, so that no artificial effects may be introduced due to the truncation of series expansion. We will show that the acceleration arising from a triaxial ellipsoid is always pointing to the equatorial plane, so that there will be no chance for the force balance out of the equatorial plane. Therefore, all equilibrium points must be on the equatorial plane and the out-of-plane equilibrium points do not exist. In a frame of reference rotating with the ellipsoid, the ellipsoid has a surface defined by X 2 A2 + Y 2 B 2 + Z 2 C 2 = 1, (19) where A ≤ B ≤ C are the long, intermediate, and small semi-axes of the ellipsoid. Apparently, these axes rest exactly on the coordinate axes. After normalizing the coordinates by A, the semi-axes read a ≡ 1, b ≡ C A . B A , c ≡ And the gravitational potential outside the ellipsoid is (see e.g. Scheeres 1994; Romannov & Doedel 2012) V (x, y, z) = 3GM 4A3 Z ∞ u (cid:18)1 − x2 a2 + s − y2 b2 + s − z2 c2 + s(cid:19) × where u is a positive root of the following equation ds p(a2 + s)(b2 + s)(c2 + s) , (20) x2 a2 + u + y2 b2 + u + z2 c2 + u = 1. Therefore, the acceleration in the Oz direction can be calculated (Romannov & Doedel 2012) ∂V ∂z = − 3GM z 2A3 Z ∞ u ds (c2 + s)p(a2 + s)(b2 + s)(c2 + s) . (21) (22) Because the integral in Eq. (22) is definite positive, we have – 10 – ∂V ∂z < 0 ∂V ∂z = 0 ∂V ∂z > 0 for all z > 0, for z = 0, for all z < 0. The acceleration always points toward the Oxy-plane, and the acceleration balance in z-direction can be attained only when z = 0. The out-of-plane equilibrium points do not exist. 4. Conclusion We have shown analytically that in the restricted three-body problem with one oblate primary the "out-of-plane" equilibrium points do not exist. The same arguments can be carried out and the same conclusion can be made if both primaries are oblate. The existence of such "out-of- plane" equilibrium points is due to an inappropriate application of the potential series of spherical harmonics, which gives rise to a fake vertical acceleration pointing outward the orbital plane. Surely, when the outward vertical acceleration does exist in a system, the out-of-plane equilib- rium points may appear. For example, when one of the primaries are very luminous, the radiation pressure may overcome the gravitation from it, then such kind of new equilibrium points may ap- pear. This is possible for dust particles or spacecrafts with very large area-mass-ratio (A/m), like solar sails. Another example is the artificial equilibrium points created by low thrust. No matter in which case, we suggest that special attention should be paid to the application of the spherical harmonic expansion in the close vicinity of a celestial body. Last but not least, the argument presented in this paper hardly affects the calculation of location and stability of the classical equilibrium points, L1 ∼ L5, provided that they are not too close to the oblate body. We thank the National Natural Science Foundation of China (NSFC, Grants No.11473016 & No.11333002) for financial support. REFERENCES Abouelmagd, E.I, & Guirao, J.L.G. 2016, Applied Mathematics and Nonlinear Sciences, 1, 123 Ansari, A.A. 2017, Applied Mathematics and Nonlinear Sciences, 2, 529 Abouelmagd, E.I., & Mostafa, A. 2015, Ap&SS, 357, 58 Douskos, C.N., & Markellos, V.V. 2006, A&A, 446, 357 – 11 – Elshaboury, S.M., Abouelmagd, E., Kalantonis, V.S., & Perdios, E.A. 2016, Ap&SS, 361, 315 Murray, C.D., & Dermott, S.F. 1999, Solar System Dynamics, Cambridge Univ. Press, Cambridge Oberti, P., & Vienne, A. 2003, A&A, 397, 353 Perdiou, A.E., Markellos, V.V., & Douskos, C.N. 2005, Earth, Moon and Planets, 97, 127 Ragos, O., & Zagouras, C. 1988, Cel. Mechanics, 44, 135 Romain, G., & Jean-Pierre, B. 2001, Celest. Mech. Dyn. Astron., 79, 235 Roman, R. 2001, Ap&SS, 275, 425 Romannov, V.A., & Doedel, E.J. 2012, International Journal of Bifurcation and Chaos, 22, 1230035 Scheeres, D. 1994, Icarus, 110, 225 Sharma, R.K., & Subba Rao P.V. 1976, Celes. Mech. 13, 137 Simmons, J.F.L., McDonald, A.J.C., & Brown, J.C. 1985, Celest. Mech. Dyn. Astron., 35, 145 Singh, J., & Mohammed, H.L. 2013, Ap&SS, 345, 265 Singh, J., & Umar, A. 2013, Ap&SS, 344, 13 Song, M., He, X., & He, D. 2016, Ap&SS, 361, 327 Szebehely, V. 1967, Theory of Orbits, Academic Press. Todoran, I. 1994, Ap&SS, 215, 237 Zamaro, M., & Biggs, J.D. 2015, Celest. Mech. Dyn. Astron., 122, 263 Zeng, X., Baoyin, H., & Li, J. 2016, Ap&SS, 361, 15 Zeng, X., Liu, X., & Li, J. 2017, R.A.A. (Research in Astronomy and Astrophysics) 17, 2 (10pp) This preprint was prepared with the AAS LATEX macros v5.2.
1703.00011
1
1703
2017-02-28T19:00:00
1 to 2.4 micron Near-IR spectrum of the Giant Planet $\beta$ Pictoris b obtained with the Gemini Planet Imager
[ "astro-ph.EP" ]
Using the Gemini Planet Imager (GPI) located at Gemini South, we measured the near-infrared (1.0-2.4 micron) spectrum of the planetary companion to the nearby, young star $\beta$ Pictoris. We compare the spectrum obtained with currently published model grids and with known substellar objects and present the best matching models as well as the best matching observed objects. Comparing the empirical measurement of the bolometric luminosity to evolutionary models, we find a mass of $12.9\pm0.2$ $\mathcal{M}_\mathrm{Jup}$, an effective temperature of $1724\pm15$ K, a radius of $1.46\pm0.01$ $\mathcal{R}_\mathrm{Jup}$, and a surface gravity of $\log g = 4.18\pm0.01$ [dex] (cgs). The stated uncertainties are statistical errors only, and do not incorporate any uncertainty on the evolutionary models. Using atmospheric models, we find an effective temperature of $1700-1800$ K and a surface gravity of $\log g = 3.5$-$4.0$ [dex] depending upon model. These values agree well with other publications and with "hot-start" predictions from planetary evolution models. Further, we find that the spectrum of $\beta$ Pic b best matches a low-surface gravity L2$\pm$1 brown dwarf. Finally comparing the spectrum to field brown dwarfs we find the the spectrum best matches 2MASS J04062677-381210 and 2MASS J03552337+1133437.
astro-ph.EP
astro-ph
Draft version March 2, 2017 Preprint typeset using LATEX style AASTeX6 v. 1.0 1 TO 2.4 MICRON NEAR-IR SPECTRUM OF THE GIANT PLANET β PICTORIS b OBTAINED WITH THE GEMINI PLANET IMAGER 7 1 0 2 b e F 8 2 . ] P E h p - o r t s a [ 1 v 1 1 0 0 0 . 3 0 7 1 : v i X r a Jeffrey Chilcote1, Laurent Pueyo2, Robert J. De Rosa3, Jeffrey Vargas3, Bruce Macintosh4, Vanessa P. Bailey4, Travis Barman5, Brian Bauman6, Sebastian Bruzzone7, Joanna Bulger8, Adam S. Burrows9, Andrew Cardwell10,11, Christine H. Chen2, Tara Cotten12, Daren Dillon13, Rene Doyon14, Zachary H. Draper15,16, Gaspard Duchene3,17, Jennifer Dunn16, Darren Erikson16, Michael P. Fitzgerald18, Katherine B. Follette4,35, Donald Gavel13, Stephen J. Goodsell19,20, James R. Graham3, Alexandra Z. Greenbaum21, Markus Hartung10, Pascale Hibon22, Li-Wei Hung18, Patrick Ingraham23, Paul Kalas3,24, Quinn Konopacky25, James E. Larkin18, J´erome Maire25, Franck Marchis24, Mark S. Marley26, Christian Marois15,16, Stanimir Metchev7, Maxwell A. Millar-Blanchaer27,36, Katie M. Morzinski28, Eric L. Nielsen4,24, Andrew Norton13, Rebecca Oppenheimer29, David Palmer6, Jennifer Patience30, Marshall Perrin2, Lisa Poyneer6, Abhijith Rajan30, Julien Rameau14, Fredrik T. Rantakyro10, Naru Sadakuni31, Leslie Saddlemyer16, Dmitry Savransky32, Adam C. Schneider30, Andrew Serio10, Anand Sivaramakrishnan2, Inseok Song12, Remi Soummer2, Sandrine Thomas23, J. Kent Wallace27, Jason J. Wang3, Kimberly Ward-Duong30, Sloane Wiktorowicz33, and Schuyler Wolff34 1Dunlap Institute for Astronomy & Astrophysics, University of Toronto, Toronto, ON M5S 3H4, Canada 2Space Telescope Science Institute, Baltimore, MD 21218, USA 3Astronomy Department, University of California, Berkeley; Berkeley CA 94720, USA 4Kavli Institute for Particle Astrophysics and Cosmology, Department of Physics, Stanford University, Stanford, CA, 94305, USA 5Lunar and Planetary Laboratory, University of Arizona, Tucson AZ 85721, USA 6Lawrence Livermore National Laboratory, 7000 East Ave., Livermore, CA 94550 7Department of Physics and Astronomy, Centre for Planetary Science and Exploration, the University of Western Ontario, London, ON N6A 3K7, Canada 8Subaru Telescope, NAOJ, 650 North A'ohoku Place, Hilo, HI 96720, USA 9Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA 10Gemini Observatory, Casilla 603, La Serena, Chile 11Large Binocular Telescope Observatory, 933 N Cherry Ave., Tucson AZ 85721, USA 12Department of Physics and Astronomy, University of Georgia, Athens, GA 30602, USA 13University of California Observatories/Lick Observatory, University of California, Santa Cruz; Santa Cruz, CA 95064, USA 14Institut de Recherche sur les Exoplan`etes, D´epartment de Physique, Universit´e de Montr´eal, Montr´eal QC H3C 3J7, Canada 15University of Victoria, 3800 Finnerty Rd, Victoria, BC V8P 5C2, Canada 16National Research Council of Canada Herzberg, 5071 West Saanich Rd, Victoria, BC V9E 2E7, Canada 17Univ. Grenoble Alpes/CNRS, IPAG, F-38000 Grenoble, France 18Department of Physics & Astronomy, University of California, Los Angeles, CA 90095, USA 19Gemini Observatory, 670 N. A'ohoku Place, Hilo, HI 96720, USA 20Department of Physics, Durham University, Stockton Road, Durham DH1, UK 21Department of Astronomy, University of Michigan, Ann Arbor, MI 48109, USA 22European Southern Observatory, Alonso de Cordova 3107, Vitacura, Santiago, Chile 23Large Synoptic Survey Telescope, 950N Cherry Av, Tucson, AZ 85719, USA 24SETI Institute, Carl Sagan Center, 189 Bernardo Avenue, Mountain View, CA 94043, USA 25Center for Astrophysics and Space Science, University of California San Diego, La Jolla, CA 92093, USA 26Space Science Division, NASA Ames Research Center, Mail Stop 245-3, Moffett Field CA 94035, USA 27Jet Propulsion Laboratory, California Institute of Technology Pasadena CA 91125, USA 28Steward Observatory, University of Arizona, Tucson AZ 85721, USA 29American Museum of Natural History, Department of Astrophysics, New York, NY 10024, USA 30School of Earth and Space Exploration, Arizona State University, PO Box 871404, Tempe, AZ 85287, USA 31Stratospheric Observatory for Infrared Astronomy, Universities Space Research Association, NASA/Armstrong Flight Research Center, 2825 East Avenue P, Palmdale, CA 93550, USA 32Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853, USA 33The Aerospace Corporation, 2310 E. El Segundo Blvd., El Segundo, CA 90245 34Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218, USA 2 35NASA Sagan Fellow 36NASA Hubble Fellow Chilcote et al. ABSTRACT Using the Gemini Planet Imager (GPI) located at Gemini South, we measured the near-infrared (1.0– 2.4 µm) spectrum of the planetary companion to the nearby, young star β Pictoris. We compare the spectrum obtained with currently published model grids and with known substellar objects and present the best matching models as well as the best matching observed objects. Comparing the empirical measurement of the bolometric luminosity to evolutionary models, we find a mass of 12.9± 0.2MJup, an effective temperature of 1724 ± 15 K, a radius of 1.46 ± 0.01RJup, and a surface gravity of log g = 4.18± 0.01 [dex] (cgs). The stated uncertainties are statistical errors only, and do not incorporate any uncertainty on the evolutionary models. Using atmospheric models, we find an effective temperature of 1700 − 1800 K and a surface gravity of log g = 3.5–4.0 [dex] depending upon model. These values agree well with other publications and with "hot-start" predictions from planetary evolution models. Further, we find that the spectrum of β Pic b best matches a low-surface gravity L2±1 brown dwarf. Finally comparing the spectrum to field brown dwarfs we find the the spectrum best matches 2MASS J04062677–381210 and 2MASS J03552337+1133437. Keywords: (stars:beta Pictoris) planetary systems - instrumentation: adaptive optics - techniques: spectroscopic - infrared: general 1. INTRODUCTION Since the discovery of 51 Pegasi b in 1995 (Mayor & Queloz 1995), the search for and discovery of ex- trasolar planets has broadly changed our understand- ing of planetary systems. Direct imaging allows for the discovery of planets on solar system-scale orbits, pro- vides new insight into the formation and characteristics of extrasolar systems, and enable direct spectroscopic observations of their atmospheres. Despite decades of efforts to image young Jupiter-mass exoplanets still lu- minous as a result of their formation process, only a handful of extrasolar planets have ever been directly imaged. Examples of such planets include 2M1207b (Chauvin et al. 2005), Fomalhaut b (Kalas et al. 2008), the HR8799 system (Marois et al. 2008, 2010), β Pic b (Lagrange et al. 2010), IRXS J1609 b (Lafreni`ere et al. 2010), HD 95086 b (Rameau et al. 2013), 51 Eri b (Mac- intosh et al. 2015), and HD 131399 Ab (Wagner et al. 2016). β Pictoris (β Pic, HD 39060) is a 24±3 Myr (Bell et al. 2015), A6V star located at a distance of 19.44 ± 0.05 pc (Gray et al. 2006; van Leeuwen 2007). β Pic repre- sents the earliest example of high contrast imaging to directly detect a circumstellar disk (Smith & Terrile 1984). The disk is seen edge-on and shows an asym- metric structure that has been attributed to planetary perturbations (Burrows et al. 1995; Kalas & Jewitt 1995; Golimowski et al. 2006; Mouillet et al. 1997; Heap et al. 2000; Augereau et al. 2001). The planet β Pic b was first detected by VLT/NaCo (Lagrange et al. 2010). Since then, its orbit has been constrained via careful astro- metric monitoring (Chauvin et al. 2012; Nielsen et al. 2014; Millar-Blanchaer et al. 2015; Wang et al. 2016). The atmospheric properties of the planet have been es- timated from photometric and spectroscopic measure- ments using a number of adaptive optics (AO) fed in- struments such as Gemini/NICI (Boccaletti et al. 2013), Magellan AO (Males et al. 2014; Morzinski et al. 2015), Gemini/GPI (Chilcote et al. 2015), and VLT/SPHERE (Baudino et al. 2015). Because of the presence of a dynamically perturbed debris disk (Mouillet et al. 1997; Lagrange et al. 2012a; Millar-Blanchaer et al. 2015), along with well docu- mented constraints on its age (Bell et al. 2015), the β Pic planetary system is an ideal laboratory to understand the formation and evolution of sub-stellar objects near the planet/brown dwarf limit. The luminosity and col- ors of β Pic b are indeed similar to early-type brown dwarfs (Males et al. 2014; Morzinski et al. 2015; Bon- nefoy et al. 2014; Currie et al. 2013). However, con- straints from radial velocity observations place its dy- namical mass well below the value expected for an iso- lated field object of the same luminosity (Lagrange et al. 2012b). The recently identified population of young isolated brown dwarfs with low-surface gravity (Kirk- patrick et al. 2008; Allers & Liu 2013; Cruz et al. 2009; Delorme et al. 2012; Faherty et al. 2013; Gagn´e et al. 2015; Schneider et al. 2016) is a more appropriate sam- ple to compare to β Pic b. Recent work has provided a preliminary look at the near-infrared low resolution spectrum of β Pic b (Bonnefoy et al. 2014; Chilcote et al. 2015; Baudino et al. 2015) and has highlighted the similarities between β Pic b and low-gravity brown dwarfs in young associations and moving groups. Such results naturally lead to questions regarding the forma- tion mechanisms underlying these two type of objects that have apparently very different dynamical origins (isolated vs. orbiting another star) and yet look similar from a spectro-photometric standpoint (Baudino et al. 2015). In this paper, we provide the empirical basis for such future investigations by presenting the most comprehen- sive spectrum of the β Pic b planet to-date. Our data, obtained with the Gemini Planet Imager between 2014 and 2016, covers the Y, J, H, and K bands. In Section 2, we discuss the observations, data reduction, and spectral extraction. In Section 3.2, we compare the spectrum of β Pic b to those of a wide array of brown dwarfs, and present the best fitting objects along with comparisons to low-surface gravity brown dwarf spectral standards. An analysis of the spectrum, along with existing pho- tometry, and comparison to existing models is presented in Section 3. Finally, conclusions are discussed in Sec- tion 4. 2. OBSERVATIONS AND DATA REDUCTION The Gemini Planet Imager (GPI) was designed and built to directly image and spectroscopically character- ize young, Jupiter-sized, self-luminous extrasolar planets (Macintosh et al. 2006; Graham et al. 2007). Installed at Gemini South in the Fall of 2013, GPI underwent commissioning from the Fall of 2013 to the Fall of 2014 before becoming part of the standard instrument suite at Gemini South. β Pic was observed by the GPI Verification and Com- missioning team on 2013 November 16 and 18, 2013 De- cember 10 and 11, and 2014 March 23. A log of the ob- servations is given in Table 1. Observations performed during the instrument commissioning period (November 2013 - November 2014) were not all taken in a stable sci- ence environment, and various operational modes were used during a specific data set to evaluate performance of the instrument. For instance, during the 2013 Novem- ber 18 observations, 32 individual 59.6 s images were ob- tained in coronagraphic mode, with the cryocoolers op- erating at a reduced power level to reduce the effects of vibration introduced into the telescope (Chilcote et al. 2012; Larkin et al. 2014). Observations taken for testing purposes including changes in the AO performance pa- rameters and the vibrations levels of the IFS cryocoolers affect the shape and stability of the GPI point spread function (PSF) on a shorter time scale than would be expected from typical stable operations of Gemini. Y-band data were obtained as part a Gemini Large and Long Program focused on the study of debris disks with GPI (GS-2015B-LP-6). These observations oc- curred when the planet had already moved significantly inwards, resulting in a higher level of noise in the es- timated spectrum. The average seeing measured using a Differential Image Motion Monitor (DIMM), total ex- 3 posure times, and instrument configurations of the ob- servations presented in this paper are listed in Table 1, along with the specifics of the data-sets that were ob- tained during verification and commissioning. Each of these data sets was individually and indepen- dently reduced using the GPI pipeline, with standard recipes provided by the GPI Data Reduction Pipeline (Perrin et al. 2014). A short arc lamp exposure was taken with each science observation set to account for offsets of the lenslet spectra due to flexure within the IFS. The GPI data reduction pipeline was used to reduce all images by applying dark corrections, fitting and re- moving vibration-induced microphonics noise (Chilcote et al. 2012; Ingraham et al. 2014), removing bad pix- els, fitting satellite spot locations (Wang et al. 2014), and extracting each microspectra to create a 37-channel spectral cube. For K-band data sky frames were sub- tracted, if available, to remove the thermal background. High-contrast imaging surveys for faint substellar companions typically use Angular Differential Imaging (ADI, Marois et al. 2006) and/or Spectral Differential Imaging (SDI, Sparks & Ford 2002). These PSF sub- traction processes often lead to self-subtraction of any resolved faint companions, creating systematic biases in the extracted photometry that need to be corrected for. Previous studies have used forward modeling approaches where a negative version of the PSF is injected into the reduced images prior to PSF subtraction to estimate the flux and position of faint companions detected in PSF-subtracted images (e.g., Hinkley et al. 2013; Op- penheimer et al. 2013; Crepp et al. 2015). For this study, we use a different forward modeling approach that an- alytically models the effect of stellar PSF subtraction on the PSF of the planet to find the best planet spec- trum that matches the signal of the planet after stel- lar PSF subtraction. We use the generalized method KLIP-FM, described in Pueyo (2016), which combines the Karhunen-Lo`eve Image Projection algorithm (KLIP, Soummer et al. 2012) and forward modeling. Pueyo (2016) demonstrated the effectiveness of KLIP-FM at reducing the systematic biases inherent in ADI/SDI PSF subtraction by injecting and recovering point sources with known spectra into into GPI J-band β Pic data. KLIP-FM has also been used to measure the astrometry of β Pic b with GPI at milliarcsecond precision (Wang et al. 2016). The final PSF-subtracted images of β Pic b in each of the five GPI filters are shown in Figure 1. As described in Wang et al. (2016), we use the four satellite spots in each spectral channel to estimate the PSF of β Pic b. When using such a PSF fitting method, biases on the spectrum can arise due to a mismatch be- tween the full width at half maximum (FWHM) of the planet and model PSFs. Such a mismatch can occur as a result of the preliminary high-pass filter step before the 4 Chilcote et al. Figure 1. PSF-subtracted images of β Pic in each of the five GPI filters, with the location of β Pic b highlighted. The images −5 have been rotated such that North is up and East is to the left, with a linear color scale in units of contrast between ±2.5× 10 (±11.5 mags). The significant orbital motion of the planet between 2013 and 2015 is apparent (Wang et al. 2016). Each data set was processed using the same KLIP parameters (seven annuli, four segments per annulus, one pixel minimum movement criteria), with a reference PSF constructed from the first ten KL modes. The final images were created by averaging these PSF-subtracted data cubes along the wavelength axis. Figure 2. (top panel): The GPI spectrum of β Pic b extending from the Y band to K band (black points). Photometric measurements of β Pic b, as compiled by Morzinski et al. (2015), are also plotted (color and symbols given in legend, Lagrange et al. 2009; Quanz et al. 2010; Bonnefoy et al. 2011; Currie et al. 2011; Bonnefoy et al. 2013; Currie et al. 2013; Absil et al. 2013; Males et al. 2014; Morzinski et al. 2015). (bottom panel): Normalized filter transmission curves for the various photometric measurements of β Pic b. KLIP PSF subtraction, carried out in order to mitigate the impact of the residual atmospheric halo. We cali- brated this effect in an ad-hoc fashion by exploring an increasing sequence of high-pass filtering cutoff frequen- cies. Typically the signal-to-noise ratio (SNR) of the planet increases with a more aggressive filter that elim- inates the residual AO halo, but such filtering schemes create spurious slopes in the spectrum since they affect the morphology of the planet PSF and of the satellite spots differently. Fortunately, in all data sets considered here the planet SNR is high enough so that there exists a large range of high-pass filtering parameters for which the planet spectrum is stable (between 8 and 15 px, as defined in Wang et al. 2016). The resulting spectrum that minimizes the residuals at the location of β Pic b was then estimated by forward modelling, normalized by the average satellite spot intensity in each wavelength channel. An 8000 K, log g = 4.0 [dex] (Gray et al. 2006) BT-NextGen1 model (Allard et al. 2012) convolved to the resolution of GPI, was used to approximate the A6V stellar spectrum of β Pic A. This allows the instru- mental and telluric features under identical conditions to be estimated for the planet spectrum and then re- moved. To exclude low SNR data at the edges of the filter band-passes, we trim the β Pic b spectrum to ex- clude wavelength channels where the filter transmission is below 80 %, excluding 27 of the 185 channels of the 1 https://phoenix.ens-lyon.fr/Grids/BT-NextGen/ 2500−250−2500250∆DE(mas)Y/2015-12-052500−250J/2013-12-102500−250∆RA(mas)H/2013-11-182500−250K1/2013-11-162500−250K2/2013-11-160.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)PhotometrycompiledbyMorzinski+15:Lagrange+09Quanz+10Bonnefoy+11Currie+11Bonnefoy+13Currie+13/14Absil+13Males+14Morzinski+1512345Wavelength(µm)0.00.51.0λRλ full spectrum. The average SNR per resolution element of the final trimmed spectrum, plotted in Figure 2, was 3 in Y, 17 in J, 15 in H, 14 in K1, and 19 in K2. 3. RESULTS & DISCUSSION 3.1. Bolometric Luminosity The bolometric luminosity of β Pic b was most re- cently estimated as log Lbol/L(cid:12) = −3.78 ± 0.03 [dex] by Morzinski et al. (2015)2, calculated using GPI J- and H-band spectroscopy and a re-calibration of optical through thermal-infrared photometry to remove any sys- tematic bias introduced in previous studies. The bolo- metric luminosity of β Pic b was reassessed using the new spectroscopic measurements presented in Section 2, which significantly improves the sampling of its spectral energy distribution (SED) in the Y (0.98–1.13 µm) and K (1.91–2.38 µm) bands. The procedure is similar to that employed by Morzin- ski et al. (2015). The GPI spectrum was combined with band-averaged photometry at 3.31 µm, 3.34 µm, L(cid:48) (3.80 µm), 4.10 µm, and M(cid:48) (4.72 µm), the values for which are given by Morzinski et al. (2015). The YS (0.985 µm) and K (2.27 µm) photometry measure- ments used by Morzinski et al. (2015) were rejected from this analysis due to the significant overlap with the new GPI spectrum presented in Figure 2. The measured SED (0.98–4.72 µm) was extended to shorter and longer wavelengths using two blackbody functions. The op- tical blackbody (0.01–0.97 µm) was normalized to the integrated flux of the GPI Y -band spectrum, while the infrared blackbody (4.73–1000 µm) was normalized to the M(cid:48) photometric point. The bolometric luminosity was then measured by integrating the synthetic spec- trum formed by the combination of the measured SED and the two blackbody functions. The final luminosity and its uncertainty were esti- mated using a Monte Carlo approach by repeating the integration 105 times. Random draws were made for each trial from each of the band-average photometric points. The individual GPI spectra were varied by draw- ing from a normal distribution created from a quadratic sum of the band-averaged uncertainty and the satellite spot ratio uncertainty. Conservatively, each point within an individual GPI spectrum was adjusted by the same amount to account for correlation between the spectral channels. The temperature of the two blackbody func- tions was drawn from a uniform distribution between 1500 and 1900 K for each trial, and were normalized as described previously. Despite the large range of tem- 5 Figure 3. Histograms of the values of the four parameters from the Monte Carlo analysis comparing the empirical lu- minosity and age of β Pic b to the Baraffe et al. (2003) hot-start evolutionary models. This analysis was performed using the luminosity and age presented in this paper (black histogram), and using the values prented in Morzinski et al. (2015) (log Lbol/L(cid:12) = −3.78± 0.03 [dex] and t = 23± 3 Myr; gray histogram). The asymmetric distribution for mass is caused by the significant increase in the predicted luminos- ity due to the onset of deuterium burning. peratures, there was no correlation between the choice of temperature for the blackbody extensions and the re- sulting luminosity. Using the median and 1 σ range of the 105 trials, the bolometric luminosity of β Pic b was found to be log Lbol/L(cid:12) = −3.76 ± 0.02 [dex], consis- tent with the value reported in Morzinski et al. (2015). While the choice of a blackbody function is a simplistic one, it only has a small contribution to the total flux of β Pic b, with the short- and long-wavelength blackbody extensions contributing 3 ± 1 % and 14 ± 1 %, respec- tively. The bolometric luminosity of β Pic b and the age es- timate for the system of 24 ± 3 Myr (Bell et al. 2015) were compared to the Baraffe et al. (2003) hot-start evo- lutionary models3 to derive a model-dependent estimate of the mass (M ), temperature (Teff ), radius (R), and surface gravity (g) of β Pic b. A Monte Carlo procedure was used to propagate the uncertainty of the luminosity (Lbol) and age (t) to the four derived parameters. At each step, a random luminosity and age were drawn from two normal distributions, one in log Lbol and the other in t. The model grid was linearly interpolated first in log t to the randomly selected age, and then in log M to an arbitrarily high resolution. Interpolation was performed in log Lbol, log Teff , R, and log g due to their behavior as a function of log t and log M . The randomly selected luminosity was then used to select a model within the interpolated grid. This process was repeated 105 times yielding M = 12.9 ± 0.2MJup, Teff = 1724 ± 15 K, R = 1.46 ± 0.01RJup, and log g = 4.18 ± 0.01 [dex] (Figure 3). These are consistent consistent with the re- sults of a previous analysis by Morzinski et al. (2015), 2 We adopt the same convention as Morzinski et al. (2015) where script letters are used to denote nominal Solar and Jovian values as defined by IAU resolutions. 3 https://phoenix.ens-lyon.fr/Grids/AMES-Cond/ISOCHRONES/ 121314Mass(MJup)Number16501750Teff(K)1.421.461.50Radius(RJup)4.154.20logg(dex) 6 Chilcote et al. who reported a mass, effective temperature, and ra- dius for β Pic b of 12.7 ± 0.3MJup, 1708 ± 23 K, and 1.45 ± 0.02RJup, respectively. 3.2. Comparison with field objects The spectrum of β Pic b was compared with a li- brary of 1600 M-, L-, and T-dwarf spectra compiled from the SpeX Prism library4 (Burgasser 2014), the IRTF Spectral Library5 (Cushing et al. 2005), the Montreal Spectral Library6 (e.g., Gagn´e et al. 2015; Robert et al. 2016), and the sample of young ultracool dwarfs pre- sented in Allers & Liu (2013). The spectral types for the objects within the library were obtained from a number of literature sources, and are given for the individual ob- jects described later in this section. The near-infrared spectral type was used for objects with both an opti- cal and near-infrared spectral type. The literature was also searched to obtain the surface gravity classifications for each object, using either of the schemes outlined by Kirkpatrick (2005); Kirkpatrick et al. (2006); Cruz et al. (2009) (α, β, γ, δ, in descending order of surface grav- ity), or Allers & Liu (2013) (fld-g, int-g, vl-g, simi- larly). Briefly, both classification schemes categorize ul- tracool dwarfs into three groupings: field surface gravity consistent with that seen for old field dwarfs (α, fld- g), intermediate surface gravity (β, int-g), and very low surface gravity consistent with that seen for young brown dwarfs (γ, vl-g). Kirkpatrick (2005) define a fourth classification, δ, for objects which exhibit even stronger low-gravity features in their spectra than those classified as γ. The spectrum of each object was degraded to the spec- tral resolution of GPI (between λ/δλ = 35 at Y and λ/δλ = 79 at K2) by convolution with a Gaussian func- tion of the appropriate width. The uncertainties were similarly degraded, normalized by the effective number of spectral channels within the convolution window. The spectrum of β Pic b was compared to this library using three different procedures. First, the five GPI bands were fit independently to explore the sensitivity of each bandpass to the spectral type and surface gravity of β Pic b. Second, the five bands were fit simultaneously but were each normalized independently and without constraint to account for both the dispersion in near- infrared colors of young low-gravity brown dwarfs (e.g., Leggett et al. 2003), and for the uncertainty in the pho- tometric calibration of the GPI data. Third, the five bands were fit simultaneously as before, but the normal- 4 http://pono.ucsd.edu/~adam/browndwarfs/spexprism 5 http://irtfweb.ifa.hawaii.edu/~spex/IRTF Spectral Library 6 https://jgagneastro.wordpress.com/the-montreal-spectral-library/ Figure 4. χ2 ν as a function of spectral type for the M, L and T dwarfs within the spectral library fit to the spectum of β Pic b in each of the GPI bandpasses. The K1 and K2 spectra were combined to create a single K-band spectrum. Different markers were used to indicate the different gravity classes using the scheme described in Allers & Liu (2013), with the legend given at the top of the figure. The optical and near infrared gravity classification have been grouped together for clarity. Objects without any gravity classifica- tion are plotted as gray circles. Spectral standards for field- gravity (Burgasser et al. 2006; Kirkpatrick et al. 2010) and low-gravity (Allers & Liu 2013) objects are highlighted with large red and yellow crosses, respectively. ization of each band was restricted by the uncertainty of the photometric calibration of the GPI data (Maire et al. 2014). We find that all three methods yield simi- lar results in terms of the spectral type of β Pic b, and provide strong evidence for a low surface gravity. 3.2.1. Fits to the individual bands The spectrum of β Pic b in each of the four near- infrared bands (YJHK) was fit to the corresponding spectrum of each comparison object within the library. 10−1100101χ2νYField-L2.8±3.0VLG-L1.6±2.0noneα/FLD-Gβ/INT-Gγ/VL-Gδ10−1100101χ2νJField-L5.9±1.4VLG-L2.5±1.210−1100101χ2νHField-L6.6±1.2VLG-L1.9±1.9M5L0L5T0SpectralType10−1100101χ2νK1&K2Field-L2.1±2.7VLG-L2.0±1.0 The K-band spectrum of β Pic b was created by com- bining the GPI K1 and K2 spectra, discarding the over- lapping spectral channels within the K1 spectrum due to systematics in the K1 spectrum. The spectrum of the comparison object was multiplied by a scaling factor to account for the different distance and radius between that object and β Pic b. The optimal scaling factor was found analytically for each object and band (e.g., Burgasser et al. 2016). The uncertainty on the spec- trum of β Pic b and the comparison object were added in quadrature. The number of degrees of freedom was typically 29 for Y band, 32 for J band, 34 for H band, and 56 for K band. The minimum χ2 ν for each object in each band is plotted as a function of spectral type in Figure 4. The sensitivity of the J-, H-, and K-band spectra to surface gravity is apparent in Figure 4, with the low- surface gravity objects typically providing a better fit to β Pic b than field-gravity objects of the same spec- tral type. Given the low resolution of the GPI data, this sensitivity is primarily due to differences in the shape of the continuum between field and low-gravity objects (Allers & Liu 2013), rather than differences in the strengths of gravity-sensitive absorption lines. The difference between the spectra of field and low-gravity objects is most pronounced in the H and K bands, where the minimum χ2 ν of the low-surface gravity objects is sig- nificantly lower than that of field-gravity objects of the same spectral type (Fig. 4). We estimated the spectral type of β Pic b in each band by a comparison to the spectra of field surface grav- ity standards (Burgasser et al. 2006; Kirkpatrick et al. 2010), and low-surface gravity standards (Allers & Liu 2013). The weighted average of the numerical spectral types of the standards, weighted according to the ra- tio of their χ2 to the minimum χ2 of all the standards (e.g., Burgasser et al. 2010), was adopted as the spec- tral type. A systematic uncertainty of one half subtype was assumed for the standards. This process was re- peated for both surface gravity subsets, and for each of the five bands. The adopted spectral type and corre- sponding uncertainty for β Pic b are given for each band in Figure 4, ranging from L2 to L6.5 for the field surface gravity standards and from L1.5 to L2.5 for the low- surface gravity standards, both rounded to the nearest half subtype. 3.2.2. Unrestricted fit to the full spectrum The full GPI spectrum of β Pic b was then fit to each object within the library. Each band of the spectrum was scaled independently to account for the dispersion in near-infrared colors seen for brown dwarfs of a given spectral type (Leggett et al. 2003), and for the uncer- tainty in the absolute flux calibration of the GPI data 7 Figure 5. χ2 ν as a function of spectral type for each object within the spectral library for the unrestricted fit described in § 3.2.2 (top panel), and for the restricted fit described in § 3.2.3 (bottom panel), to the GPI spectrum of β Pic b. The symbols are as in Figure 4. In both cases, the low-gravity ob- jects typically have lower χ2 ν values than field-gravity objects of the same spectral type. The spectral type of β Pic b was estimated for both gravity subsets, with the estimates being consistent between the two different fitting procedures. (Maire et al. 2014). This was achieved by summing the χ2 of each object in each band, equivalent to fitting the five bands simultaneously with five independent scale factors. The resulting minimum χ2 ν values for each ob- ject are plotted in Figure 5 (top panel). The number of degrees of freedom was typically 152, but was lower for objects that had limited spectral coverage. As with the previous fit, the spectral type of β Pic b was estimated as L4± 2.5 using the field-gravity standards, and L2± 1 using the low-gravity standards, consistent with previ- ous estimates based on fits to the broadband photometry of β Pic b (Males et al. 2014). The significantly lower χ2 ν values for the low-gravity objects within the library provides strong evidence for the low surface gravity of β Pic b, consistent with previous photometric and spec- troscopic analyses (Chilcote et al. 2015; Morzinski et al. 2015). Of all the objects within the library, the best fit object from the unrestricted fit was found to be 2MASS J03552337+1133437 (2M 0355+11, χ2 ν = 0.45), a nearby (8–9 pc, Faherty et al. 2013; Liu et al. 2013) 100101χ2νUnrestrictedfit:Field-L4.4±2.0VLG-L2.1±0.7M5L0L5T0SpectralType100101χ2νRestrictedfit:Field-L3.4±1.1VLG-L2.2±1.2 8 Chilcote et al. and extremely red (Cruz et al. 2009) brown dwarf with a near-infrared (optical) spectral type of L3 vl-g (L5γ). The spectrum of 2M 0355+11 is plotted with β Pic b in Figure 6 (top panel). Based on a kinematic analysis and the spectral signatures of youth, 2M 0355+11 is a confirmed member of the 149+51−19 Myr (Bell et al. 2015) AB Doradus moving group (Faherty et al. 2016). Due to the unusual near-infrared spectrum of 2M 0355+11, Gagn´e et al. (2015) assign it a special spectral classifi- cation of L3–L6γ, and classify objects with similar spec- tra as J0355-type. These objects are visually similar to L4γ objects but with a shallower CO band at 2.3 µm (Gagn´e et al. 2015). The spectrum of 2M 0355+11 ex- hibits strong indicators of low surface gravity, and the unusual near- and mid-infrared colors were explained by flux redistribution to longer wavelengths due to en- hanced dust or thick clouds in the photosphere (Faherty et al. 2013, 2016). A good fit was also found to the spectrum of 2MASS J22351658–3844154 (χ2 ν = 0.52), an L1.5γ can- didate member of the 45 ± 4 Myr (Bell et al. 2015) Tucana-Horologium moving group (Gagn´e et al. 2015). Of the low-gravity near-infrared standards defined by Allers & Liu (2013), the best fit was found to be the L2 vl-g standard 2MASSI J0536199–192039 (χ2 ν = 0.55), a candidate member of both the 42+6−4 Myr (Bell et al. 2015) Columba moving group (Gagn´e et al. 2014, 2015), and the 24 ± 3 Myr (Bell et al. 2015) β Pictoris mov- ing group (Faherty et al. 2016). This object, and the other early- to mid-L near-infrared spectral standards from Allers & Liu (2013) are plotted in Figure 7. 3.2.3. Restricted fit to the full spectrum Finally, we fit the morphology within each band and the relative flux levels of the different bands by restrict- ing the range over which the scale factor for each band can vary based on the expected photometric accuracy of GPI. In order to restrict this range, the χ2 equation was modified with a cost term based on a comparison of the scale factor for a band, and the uncertainty on the satellite spot ratio in that band (Maire et al. 2014). The χ2 for the kth comparison object was calculated as: 4(cid:88) nj(cid:88) j=0 i=0 χ2 k =  Fj(λi) − αkβj,kCj,k(λi) (cid:113) 4(cid:88) σ2 Fj (λi) + σ2 Cj,k (λi) 2 (cid:20) βj,k − 1 σmj + nj j=0 (cid:21)2 (1) where Fj(λi) and σFj (λi) are the flux and uncertainty of β Pic b in the ith wavelength channel of the jth band and Cj,k(λi) and σCj,k (λi) is the flux and uncertainty of the kth comparison object in the same channel and The minimum χ2 band. The spectrum of the comparison object is multi- plied both by a scale factor αk which is the same for each band, and by an additional scale factor for the jth band βj,k. The first term of Equation (1) gives the standard χ2 equation, summed over all nj wavelength channels in the jth band, and over all five bands. This is modified by a cost term which compares the band-dependent scaling factor βj,k to the fractional uncertainty of the satellite spot flux ratio σmj for the jth band (Maire et al. 2014). ν for each object is plotted for this restricted fit in Figure 5 (bottom panel). The number of degrees of freedom was typically 151, but was lower for objects that had limited spectral coverage. As with the two previous fits, the spectral type of β Pic b was estimated using the field-gravity and low-gravity stan- dards as L3.5 ± 1.0 and L1.5 ± 1.5, respectively. While these estimates of the spectral type are consistent with the results of the unrestricted fit, the minimum χ2 ν val- ues are higher for each object due to the additional cost term included in the restricted fit, an effect that is most pronounced for the mid to late-type M-dwarfs within the library. For the restricted fit case, the best fitting result from the spectral library is 2MASS J04062677–381210 (2M 0406–38, χ2 ν = 1.04), a brown dwarf with an L0γ/L1 vl-g (optical/near-infrared, Faherty et al. 2013; Allers & Liu 2013) spectral type (Figure 6, bottom panel). The kinematics of 2M 0406–38 are ambiguous in terms of nearby moving group membership, being a probable member of several nearby moving groups as well as having consistent space motion as old field ob- jects (Faherty et al. 2016). Other objects with a good fit include 2MASS J01415823–4633574 (χ2 ν = 1.46); an L0γ/L2 (optical/near-infrared, Cruz et al. 2009; Schnei- der et al. 2014) high-probability candidate member of the Tucana-Horologium moving group (Gagn´e et al. 2014); and the Allers & Liu (2013) near-infrared low- gravity standards 2MASSW J2208136+292121 (L3 vl- g, χ2 ν = 1.10), a candidate member of the β Pictoris moving group (Liu et al. 2016), and 2MASSI J0518461– 275645 (L1 vl-g, χ2 ν = 1.30), a probable member of the Columba moving group (Liu et al. 2016). All of the Allers & Liu (2013) low-gravity standards are shown in Figure 7 (right panel). 3.2.4. Spectral type and gravity classification of β Pic b Based on a comparison of the full GPI spectrum of β Pic b to the Allers & Liu (2013) low-gravity stan- dards, the spectral type of β Pic b was estimated using the unrestricted and restricted procedures as L2.1 ± 0.7 and L2.2 ± 1.2, respectively. While these two estimates are consistent with one another for β Pic b, we would expect the unrestricted fit to more reliably estimate the spectral type of young low-gravity objects due to the 9 Figure 6. The best fit object to the spectrum of β Pic b within the spectral library for the unrestricted (top panel, Allers & Liu 2013) and restricted (bottom panel, Kirkpatrick et al. 2010) fits. The optical and near-infrared spectral type and gravity classifications are given for both objects. observed range of their near-infrared colors. Rounding to the nearest half subtype, we adopt a spectral type of L2± 1 for β Pic b, consistent with previous photometric and spectroscopic estimates of L2–5 (Currie et al. 2013), L2γ ± 2 (Bonnefoy et al. 2013), L2.5 ± 1.5 (Males et al. 2014), and L1+1−1.5 (Bonnefoy et al. 2014). The signif- icantly improved fits to the low-gravity objects within the spectral library (Figures 4 and 5) demonstrates that β Pic b has a near-infrared spectrum consistent with that of a low-surface gravity object, and as such we as- sign it a surface gravity classification of γ. We do not as- sign a classification in the Allers & Liu (2013) scheme as the bandwidths of the indices used to define this scheme are smaller than the spectral resolution of the GPI spec- trum. This spectral type estimate was converted into a bolo- metric luminosity using the J- and KS-band empiri- cal spectral type to bolometric correction relations for young, low-gravity objects derived by Filippazzo et al. (2015). We estimate an absolute J-band magnitude in the MKO system (Tokunaga et al. 2002) of β Pic b from the flux-calibrated GPI spectrum of MJ = 12.56±0.08, a bolometric correction of BCJ = 1.48±0.28, a bolometric magnitude of Mbol = 14.04 ± 0.29, and a bolometric lu- minosity of log Lbol/L(cid:12) = −3.72 ± 0.12 [dex]. Similarly, for KS-band: MK = 10.86 ± 0.15, BCK = 3.26 ± 0.13, Mbol = 14.11 ± 0.20, log Lbol/L(cid:12) = −3.75 ± 0.08 [dex]. Both of these luminosity estimates are consistent with the empirical bolometric luminosity of β Pic b presented in Section 3.1. We also convert the absolute H-band magnitude of β Pic b (MH = 11.80 ± 0.09) into an ef- fective temperature of 1681 ± 64 K using the relations derived by Filippazzo et al. (2015). The spectral type was also converted into an effective temperature using the relations presented in Faherty et al. (2016). Us- ing the polynomial fit to bona fide and high-likelihood moving group members, the spectral type of β Pic b corresponds to an effective temperature of 1847± 242 K. Including probable moving group members in the poly- nomial fit increases the derived effective temperature to 1888 ± 215 K, while including both probable mov- ing group members and T-dwarf imaged planetary-mass companions decreases it to 1787±240 K. These estimates are consistent with the effective temperature estimated from the evolutionary models in Section 3.1. 3.3. Comparison with atmospheric models Using the spectral data obtained with GPI, we com- puted updated best spectral model fits combining GPI spectral data with previously published photometry of β Pic b (Lagrange et al. 2009; Quanz et al. 2010; Bon- nefoy et al. 2011; Currie et al. 2011; Bonnefoy et al. 2013; Currie et al. 2013; Absil et al. 2013; Males et al. 2014; Morzinski et al. 2015). The SED of β Pic b was compared to publicly available grids of model atmo- 0.00.51.01.5NormalizedFlux(Fλ)2MASSJ03552337+1133437(L5γ/L3vl-g,χ2ν=0.45)Unrestrictedfit1.01.21.41.61.82.02.22.4Wavelength(µm)0.00.51.01.5NormalizedFlux(Fλ)2MASSJ04062677-3812102(L0γ/L1vl-g,χ2ν=1.04)Restritedfit 10 Chilcote et al. Figure 7. The spectrum of β Pic b (black points) fit to the L-dwarf near-infrared low-gravity standards from Allers & Liu (2013) (solid curves) using the unrestricted fit described in Section 3.2.2 (left panel) and the restricted fit described in Section 3.2.3 (right panel). The spectrum of each standard is normalized to the flux at 1.65 µm, and then offset for clarity . spheres: Ames-Dusty7 (Chabrier et al. 2000; Allard et al. 2001), BT-Settl (2015)8 (Allard et al. 2012) and Drift Phoenix9 (Helling et al. 2008; Woitke & Helling 2003; Helling & Woitke 2006). All of these model grids are calculated using the Phoenix atmo- sphere models. The Ames-Dusty grid combines the NASA AMES molecular H2O and TiO line lists and in- cludes the treatment for the condensation of dust within the atmosphere. The BT-Settl (2015) models are part of the BT model family, using updated line lists and re- 7 https://phoenix.ens-lyon.fr/Grids/AMES-Dusty 8 https://phoenix.ens-lyon.fr/Grids/BT-Settl/CIFIST2011 2015 9 http://svo2.cab.inta-csic.es/theory/newov vised solar abundances. BT-Settl uses a detailed cloud model to define the distribution of condensates within the atmosphere. The Drift Phoenix model grids com- bine the Phoenix model with the non-equilibrium cloud model Drift. The fitting procedure was similar to that described in Section 3.2 for the individual spectra in each model grid. The spectra were convolved such that their spectral res- olution matched the spectral resolution in each of the GPI wavelength bands (Larkin et al. 2014). To compute synthetic photometry, the model spectra were integrated over the bandpass using filter curves published for each individual filter and instrument. The (χ2 ν) statistic for each model in comparison to the spectral data was cal- culated using the method described in Section 3.2 and 1.001.251.501.752.002.25Wavelength(µm)0123456NormalizedFlux(Fλ)+Offset2MASSIJ0518461-275645L1,χ2ν=0.752MASSIJ0536199-192039L2,χ2ν=0.552MASSJ15515237+0941148L4,χ2ν=0.862MASSWJ2208136+292121L3,χ2ν=0.702MASSJ22134491-2136079L0,χ2ν=0.892MASSJ22443167+2043433L6,χ2ν=1.441.001.251.501.752.002.25Wavelength(µm)L1,χ2ν=1.30L2,χ2ν=2.13L4,χ2ν=2.85L3,χ2ν=1.10L0,χ2ν=1.52L6,χ2ν=10.79 11 Figure 8. The best fit models within each of the three atmospheric model grids found using only the photometric measure- ments (blue dashed curve) and using both the photometric and spectroscopic measurements (red solid curve) of β Pic b. The photometric measurements of β Pic b compiled by Morzinski et al. (2015) are plotted as black points, while the GPI spectra presented in this study are plotted as light gray points. Synthetic photometry (open blue and red squares) was computed for each model using the filter profiles shown in Figure 2. using Equation (1), where Cj,k(λi) is the flux of model spectrum in the same channel and band. The χ2 ν statis- tic was calculated for each band, and for the spectral bands a punitive factor to account for the uncertainty on the satellite spot ratio in that band was used. We compute this best fit result using only the existing pho- tometry points and for the photometry points combined with the GPI spectrum. The best fit results span a range of grid models from 1700–1800 K, with a log g = 3– 4 [dex] and a R = 1.17–1.41RJup. The best fit to the combined photometry and spectroscopy is found in the Drift Phoenix grid (Teff = 1700 K, log g = 4.0 [dex], R = 1.41RJup, χ2 ν = 1.81). The best fit spectrum for each of the different models for both photometry only and GPI spectrum and photometry are shown in Fig- ure 8 and the results are shown in Table 3. This process was repeated on an interpolated version of each grid, with the points between grid points inter- polated using a quadratic spline in the logarithm of the flux, where the spacing of Teff and log g were reduced to an arbitrarily high resolution of 1 K and 0.005 [dex], respectively. The grids were also interpolated using a bilinear interpolation scheme which produced simi- lar results. We ran the same analysis as above and find that the best fit results span a range of grid mod- els from 1651–1815 K, with a log g = 3.00–4.50 [dex] and a R = 1.18–1.58RJup. Again, the best fit grid is produced by the Drift Phoenix with Teff = 1651 K, log g = 3.00 [dex], R = 1.58RJup and a χ2 ν = 1.21. The χ2 ν surfaces for the interpolated grids are shown in Figure 9, with confidence intervals calculated from the probability p ∝ exp(−χ2/2). As the χ2 does not incor- 0.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)AMESDustyPhoto.onlyTeff=1700Klogg=3.5[dex]R=1.31RJuplogLbol/L(cid:12)=−3.86χ2ν=2.66GPI&Photo.Teff=1800Klogg=3.5[dex]R=1.17RJuplogLbol/L(cid:12)=−3.87χ2ν=3.490.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)BTSettl(2015)Photo.onlyTeff=1800Klogg=3.0[dex]R=1.38RJuplogLbol/L(cid:12)=−3.87χ2ν=2.99GPI&Photo.Teff=1800Klogg=3.5[dex]R=1.22RJuplogLbol/L(cid:12)=−3.87χ2ν=3.1712345Wavelength(µm)0.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)DriftPhoenixPhoto.onlyTeff=1700Klogg=3.5[dex]R=1.41RJuplogLbol/L(cid:12)=−3.81χ2ν=1.55GPI&Photo.Teff=1700Klogg=4.0[dex]R=1.41RJuplogLbol/L(cid:12)=−3.80χ2ν=1.81 12 Chilcote et al. Figure 9. Goodness of fit statistic (χ2 ν ) for the Ames-Dusty, BT-Settl (2015) and Drift Phoenix model grids as a function of effective temperature and surface gravity. Grid points are indicated with light-gray diamonds. The points between model grid points were linearlly interpolated version of the grid, with a spacing of 1 K for Teff and 0.005 [dex] for log g. The best fit model within the original grid is indicated by a large diamond, with the best fit model within the interpolated grid indicated by a circle. The white contours indicate the 68 % (solid), 95 % (dashed), and 99 % (dotted) confidence intervals, calculated using the χ2. porate any model uncertainties, these confidence inter- vals only represent the uncertainty on these parameters for this specific model. 3.4. Comparison with combined evolutionary and atmospheric models The observed SED was also compared with the com- bined evolutionary and atmospheric models of Spiegel & Burrows (2012)10 using the same fitting procedure as with the previous grids. These models differ from the atmosphere-only models in that the grid was computed in terms of the mass and initial entropy of the planet, rather than the effective temperature and surface grav- ity. The Spiegel & Burrows (2012) grid is bound by the canonical low-entropy "cold-start" (8 kB/baryon), and high-entropy "hot-start" (13 kB/baryon) models (c.f. Marley et al. 2007), where kB is Boltzmann's constant. Here, the initial entropy describes how efficiently heat was radiated away from the forming planet during accre- tion; formation through gravitational instability may re- 10 http://www.astro.princeton.edu/~burrows/ sult in a significantly higher initial entropy than forma- tion through core accretion. These evolutionary models were then coupled with an atmospheric model (Hubeny et al. 2003; Burrows et al. 2006) to create synthetic spec- tra at each grid point. The atmospheric model used ei- ther a solar (1×) or super-solar (3×) metallicity, and ei- ther cloud-free or with a linear superposition of cloudy and cloud-free models (hybrid clouds). In total, four grids of synthetic spectra were compared to the SED of β Pic b, spanning this range of atmospheric properties. As the age of β Pic b is well-constrained, the SED of the planet was only fit to the 25 Myr models within each of the four grids. As with the fits to the atmospheric models in Sec- tion 3.3, two fits of the SED of β Pic b were made to each grid. The first using only the photometry presented in Morzinski et al. (2015), and the second combining this with the GPI spectrum presented in this study. The re- sults from these fits are given in Table 4 and the best fit spectrum for each of the different models for both photometry only and GPI spectrum and photometry are shown in Figure 10. The results of the fit to only 3.03.54.04.55.05.56.0logg(dex)Photometryonly1.2RJup1.4RJup1σ2σ3σ3σAMESDustyBestfit(coarse)Bestfit(interp.)1.2RJup1.4RJup1σ2σ3σBTSettl(2015)1.2RJup1.4RJup1σ2σ3σ3σDriftPhoenix1600170018001900Teff(K)3.03.54.04.55.05.56.0logg(dex)GPI&Photometry1.0RJup1.2RJup1.4RJup1.6RJup1600170018001900Teff(K)1.0RJup1.2RJup1.4RJup1.6RJup1.8RJup1600170018001900Teff(K)1.0RJup1.2RJup1.4RJup1.6RJup1.8RJup1.01.52.02.53.03.54.04.55.05.56.0χ2ν 13 Figure 10. The best fit model within each of the four Spiegel & Burrows (2012) grids found using only the photometric measurements (blue dashed curve) and using both the photometric and spectroscopic measurements (red solid curve) of β Pic b. The photometric measurements of β Pic b compiled by Morzinski et al. (2015) are plotted as black points, while the GPI spectra presented in this study are plotted as light gray points. Synthetic photometry (open blue and red squares) was computed for each model using the filter profiles shown in Figure 2. 0.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)Cloud-free(1×solar)Photo.onlyt=25MyrM=14MJupS=9.75kBbaryon−1χ2ν=15.39GPI&Photo.t=25MyrM=15MJupS=9.50kBbaryon−1χ2ν=68.970.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)Cloud-free(3×solar)Photo.onlyt=25MyrM=14MJupS=9.75kBbaryon−1χ2ν=14.46GPI&Photo.t=25MyrM=15MJupS=9.50kBbaryon−1χ2ν=54.720.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)Hybridclouds(1×solar)Photo.onlyt=25MyrM=14MJupS=9.75kBbaryon−1χ2ν=7.17GPI&Photo.t=25MyrM=15MJupS=9.75kBbaryon−1χ2ν=7.7012345Wavelength(µm)0.00.20.40.60.81.0Fλ(×10−14Wm−2µm−1)Hybridclouds(3×solar)Photo.onlyt=25MyrM=14MJupS=9.75kBbaryon−1χ2ν=7.43GPI&Photo.t=25MyrM=15MJupS=9.75kBbaryon−1χ2ν=5.96 14 Chilcote et al. Figure 11. Goodness of fit statistic (χ2 ν ) for the cloud-free and hybrid clouds models with both solar and super solar metallicity (Spiegel & Burrows 2012) fit to the luminosity (top row), photometry (middle row), and photometry and spectroscopy (bottom row) of β Pic b. Model fluxes were interpolated between the grid points (small grey diamonds) to a resolution of 0.01MJup in mass and 0.0025 kB/baryon in initial entropy. The best fit model within the original grid is indicated by a large diamond, with the best fit model within the interpolated grid indicated by a circle. The solid, dashed, and dotted contours indicates the 1, 2, and 3 σ confidence interval derived from the χ2 surface. The poor quality of the fit of these models to the GPI spectrum (bottom row) leads to extremely small confidence intervals as the χ2 does not incorporate any model uncertainty. 8910111213InitialEntropy(kBbaryon−1)LuminosityCloud-free(1×solar)Bestfit(coarse)Bestfit(interp.)Cloud-free(3×solar)Hybridclouds(1×solar)Hybridclouds(3×solar)8910111213InitialEntropy(kBbaryon−1)Photometry12131415Mass(MJup)8910111213InitialEntropy(kBbaryon−1)GPI&Photometry12131415Mass(MJup)12131415Mass(MJup)12131415Mass(MJup)01020304050607080χ2ν the photometry of β Pic b are consistent with that of Morzinski et al. (2015), with a best fit model at a mass of 14.0MJup and an initial entropy of 9.75 kB/baryon for each grid. Including the GPI spectrum slightly changed the best fit in each case, to a lower initial entropy for the cloud-free models, and to a higher mass for both the cloud-free and hybrid cloud models. In general, the quality of the fit to the spectrum was poor, with a min- imum χ2 ν of 6.0 (Figure 10), compared with a minimum χ2 ν of 1.8 for the model atmosphere fits presented in Section 3.3. Including the GPI spectrum in the fit sig- nificantly increases the χ2 ν for the cloud-free models, and as such they are not discussed further. This process was repeated on an interpolated version of the grid to explore the effects of the finitely sam- pled grid on the results. As in Section 3.3, a χ2 ν sur- face was calculated for each model grid for both of the data sets. These surfaces, shown in Figure 11, sug- gest that the global minimum may have been missed by Morzinski et al. (2015) due to the spacing of the grid points in mass. Using only the photometry, we find a minimum at a significantly higher initial entropy of 13 kB/baryon and a lower mass of 13.5MJup, compared with 9.75 kB/baryon and 14.0MJup reported by Morzin- ski et al. (2015). The 1 σ confidence interval extends between 10–13 kB/baryon, but is tightly constrained in terms of mass. The fits to the hybrid cloud models including the GPI spectrum are consistent with those from the coarse grid described previously, however the χ2 ν surface is similar to that from the photometry-only fit. While the minimum is at an intermediate entropy (9.75 kB/baryon) and high mass (15MJup), this extends to lower masses (13.5MJup) at a range of entropies (10–13 kB/baryon). This complex minimum is also seen when fitting the empirical luminosity of β Pic b given in Section 3.1 to the luminosity of each model grid cal- culated by integrating the synthetic spectra (Figure 11, top row). These higher initial entropies are consistent with previous comparisons to evolutionary models show- ing that the initial entropy is higher than 10.5 kB/baryon at the 95 % confidence level (Bonnefoy et al. 2014). 4. CONCLUSION We present the spectrum of β Pic b in Y, J, H, and K bands as observed with the Gemini Planet Imager between 2013 and 2016 using images which were taken as part of the verification and commissioning process, as part of an astrometric monitoring program, and as part of a Gemini Large and Long Program using GPI. Not all of the presented data was originally intended to be used for spectral extraction of the planet, but it is of sufficient quality and is valuable as it improves our understanding of the emission spectrum of β Pic b. Using the standard GPI data reduction pipeline and KLIP-FM to extract 15 the spectrum, we recover the planet at a high SNR in Y , J, H, and K bands allowing a nearly full sample across the near-IR. We compare the spectral energy distribution of β Pic b to that of young, cool, low-surface gravity brown dwarfs, and to several grids of model atmospheres that are valid over the temperature and surface gravity range expected for these objects. Compared with the near-infrared spectra of brown dwarfs in young moving groups and the field, we find that the best fit spectra are those of young low-gravity objects. Of all the objects com- pared the spectrum of β Pic b best matches that of 2MASS J03552337+1133437, a confirmed member of the 149+51−19 Myr AB Doradus moving group that exhibits strong indicators of low surface gravity (Faherty et al. 2013; Liu et al. 2013; Gagn´e et al. 2015). Based on our fits to the low-gravity standards of Allers & Liu (2013), we adopt a spectral type and gravity classification of L2γ ± 1 for β Pic b. Combining the GPI spectrum with literature pho- tometry spanning from YS (0.985 µm) to M(cid:48) (4.72 µm), we directly measure the bolometric luminosity of the planet to be log Lbol/L(cid:12) = −3.76 ± 0.02 [dex], con- sistent with previous estimates derived from photom- etry alone (Morzinski et al. 2015). Comparing to "hot- start" evolutionary models Baraffe et al. (2003) yields model-dependent estimates for the physical properties of β Pic b of M = 12.9 ± 0.2MJup, Teff = 1724 ± 15 K, R = 1.46± 0.01RJup, and log g = 4.18± 0.01 [dex]. The full SED of β Pic b was also compared to atmospheric and evolutionary model grids spanning a range of at- mospheric properties and formation scenarios. The best atmospheric fits we find are to a Drift Phoenix model atmosphere with an Teff = 1700 K, log g = 4.0 [dex], and R = 1.41RJup(χ2 ν = 1.81). These values are consistent with those derived from the bolometric luminosity, and with empirical spectral type to effective temperature re- lations derived for young low-gravity brown dwarfs (e.g., Faherty et al. 2016). Comparing to the combined atmospheric and evolu- tionary models of Spiegel & Burrows (2012) yielded a best fit at a mass of 15MJupand an intermediate en- tropy of 9.75 kB/baryon, with models including a cloudy atmosphere being strongly preferred over those with a clear atmosphere. While the best fit was found at an initial entropy that is intermediate to the predictions of the "cold-start" and "hot-start" formation scenarios, the χ2 ν surface for the interpolated version of the grid has a complex structure with a minimum extending to lower masses (∼ 13.5MJup) at a range of initial entropy val- ues between between ∼10 and 13 kB/baryon, the higher value being similar to that predicted by the "hot-start" formation scenarios. Although the points on the finer grid are based on an interpolation of the coarse grid, 16 Chilcote et al. this analysis suggests that the choice of grid point loca- tion and spacing may significantly impact the resulting best fit. If the grid were sampled more finely or shifted by 0.5MJup, and assuming the interpolated points are a reasonable representation of the "true" model with those parameters, Morzinski et al. (2015) would have reported a higher entropy as the best fit model. The empirical bolometric luminosity presented here combined with the dynamical mass constraints from La- grange et al. (2012b), and the comparison to the atmo- spheric and evolutionary models of Spiegel & Burrows (2012) both suggest a "hot-start" high-entropy forma- tion scenario for β Pic b, and are consistent with the prediction that "cold-start" low-entropy formation is an unlikely formation mechanism for wide-orbit giant plan- ets (e.g., Marleau et al. 2017). As β Pic b heads towards maximum elongation in 2023 it will become separated enough from its host star to be resolved by the near- and mid-IR instruments on the upcoming James Webb Space Telescope. Combining the measurements presented here with mid-IR spectroscopy would provide further insight into the atmospheric properties and evolutionary his- tory of the planet. Interpretation of a well-sampled SED spanning over a decade in wavelength would be ex- tremely well-suited for retrieval techniques (e.g., Burn- ingham et al. 2017) rather than by fitting to finitely- spaced model grids. The Gemini Planet Imager collaboration would like to acknowledge the memory of Leslie Saddlemyer of Canada's National Research Council, who passed away in January of 2017. Les served as the systems engineer and opto-mechanical lead for GPI during its design, con- struction, and he deserves great credit for the capabili- ties of our instrument and team. We thank the anonymous referee for the helpful comments that improved the quality of this work. The authors would like to acknowledge the financial support of the Gemini Observatory, the Dunlap In- stitute, University of Toronto, the NSF Center for Adaptive Optics at UC Santa Cruz, the NSF (AST- 0909188; AST-1211562, AST-1405505), NASA Ori- gins (NNX11AD21G; NNX10AH31G, NNX14AC21G, NNX15AC89G), and NASA NExSS (NNX15AD95G), the University of California Office of the President (LFRP-118057), and the Science and Technology Facil- ities Council (ST/H002707/1). Portions of this work were performed under the auspices of the U.S. Depart- ment of Energy by Lawrence Livermore National Labo- ratory under Contract DE-AC52-07NA27344 and under contract with the California Institute of Technology/Jet Propulsion Laboratory funded by NASA through the Sagan Fellowship Program executed by the NASA Ex- oplanet Science Institute. This work is supported by the NASA Exoplanets Research Program (XRP) by co- operative agreement NNX16AD44G. Support for this work was provided by NASA through Hubble Fellowship grant 51378.01-A awarded by the Space Telescope Sci- ence Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS5-26555. This work benefited from NASA's Nexus for Exoplanet System Science (NExSS) research coordination network sponsored by NASA's Science Mission Directorate. This research has bene- fited from the SpeX Prism Library (and/or SpeX Prism Library Analysis Toolkit), maintained by Adam Bur- gasser at http://www.browndwarfs.org/spexprism, the IRTF Spectral Library, maintained by Michael Cushing, and the Montreal Brown Dwarf and Exoplanet Spectral Library, maintained by Jonathan Gagn´e. Facility: Gemini South (GPI). REFERENCES Absil, O., Milli, J., Mawet, D., et al. 2013, A&A, 559, L12 Allard, F., Hauschildt, P. H., Alexander, D. R., Tamanai, A., & Bonnefoy, M., Lagrange, A.-M., Boccaletti, A., et al. 2011, A&A, 528, L15 Schweitzer, A. 2001, ApJ, 556, 357 Bonnefoy, M., Boccaletti, A., Lagrange, A.-M., et al. 2013, Allard, F., Homeier, D., & Freytag, B. 2012, Royal Society of Astronomy and Astrophysics, 555, A107 London Philosophical Transactions Series A, 370, 2765 Bonnefoy, M., Marleau, G.-D., Galicher, R., et al. 2014, A&A, Allers, K. N., & Liu, M. C. 2013, Astrophysical Journal, 772, 79 Augereau, J. C., Nelson, R. P., Lagrange, A. M., Papaloizou, 567, L9 Burgasser, A. J. 2014, in Astronomical Society of India J. C. B., & Mouillet, D. 2001, Astronomy and Astrophysics, 370, 447 Conference Series, Vol. 11, Astronomical Society of India Conference Series Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Burgasser, A. J., Cruz, K. L., Cushing, M., et al. 2010, ApJ, 710, Hauschildt, P. H. 2003, A&A, 402, 701 1142 Baudino, J.-L., B´ezard, B., Boccaletti, A., et al. 2015, A&A, 582, A83 Bell, C. P. M., Mamajek, E. E., & Naylor, T. 2015, MNRAS, Burgasser, A. J., Geballe, T. R., Leggett, S. K., Kirkpatrick, J. D., & Golimowski, D. A. 2006, Astrophys. J., 637, 1067 Burgasser, A. J., Lopez, M. A., Mamajek, E. E., et al. 2016, 454, 593 ApJ, 820, 32 Boccaletti, A., Lagrange, A.-M., Bonnefoy, M., Galicher, R., & Burningham, B., Marley, M. S., Line, M. R., et al. 2017, ArXiv Chauvin, G. 2013, Astronomy and Astrophysics, 551, L14 e-prints, arXiv:1701.01257 Burrows, A., Saumon, D., Guillot, T., Hubbard, W. B., & Lunine, J. I. 1995, Nature, 375, 299 Burrows, A., Sudarsky, D., & Hubeny, I. 2006, ApJ, 640, 1063 Chabrier, G., Baraffe, I., Allard, F., & Hauschildt, P. 2000, Astrophysical Journal, 542, 464 -. 2012b, Astronomy and Astrophysics, 542, A40 Larkin, J. E., Chilcote, J. K., Aliado, T., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 91471K Leggett, S. K., Hawarden, T. G., Currie, M. J., et al. 2003, Chauvin, G., Lagrange, A.-M., Dumas, C., et al. 2005, MNRAS, 345, 144 17 Astronomy and Astrophysics, 438, L25 Chauvin, G., Lagrange, A.-M., Beust, H., et al. 2012, Astronomy and Astrophysics, 542, A41 Chilcote, J., Barman, T., Fitzgerald, M. P., et al. 2015, ApJL, 798, L3 Chilcote, J. K., Larkin, J. E., Maire, J., et al. 2012, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8446, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Crepp, J. R., Rice, E. L., Veicht, A., et al. 2015, ApJL, 798, L43 Cruz, K. L., Kirkpatrick, J. D., & Burgasser, A. J. 2009, AJ, 137, 3345 Currie, T., Burrows, A., Itoh, Y., et al. 2011, Astrophysical Journal, 729, 128 Currie, T., Burrows, A., Madhusudhan, N., et al. 2013, Astrophysical Journal, 776, 15 Cushing, M. C., Rayner, J. T., & Vacca, W. D. 2005, ApJ, 623, 1115 Delorme, P., Gagn´e, J., Malo, L., et al. 2012, A&A, 548, A26 Faherty, J. K., Rice, E. L., Cruz, K. L., Mamajek, E. E., & N´unez, A. 2013, AJ, 145, 2 Faherty, J. K., Riedel, A. R., Cruz, K. L., et al. 2016, ApJS, 225, 10 Filippazzo, J. C., Rice, E. L., Faherty, J., et al. 2015, ApJ, 810, 158 Gagn´e, J., Lafreni`ere, D., Doyon, R., Malo, L., & Artigau, ´E. 2014, ApJ, 783, 121 Gagn´e, J., Faherty, J. K., Cruz, K. L., et al. 2015, ApJS, 219, 33 Golimowski, D. A., Ardila, D. R., Krist, J. E., et al. 2006, Astronomical Journal, 131, 3109 Graham, J. R., Macintosh, B., Doyon, R., et al. 2007, ArXiv e-prints, arXiv:0704.1454 Gray, R. O., Corbally, C. J., Garrison, R. F., et al. 2006, Astronomical Journal, 132, 161 Heap, S. R., Lindler, D. J., Lanz, T. M., et al. 2000, Astrophysical Journal, 539, 435 Helling, C., Dehn, M., Woitke, P., & Hauschildt, P. H. 2008, ApJL, 675, L105 Helling, C., & Woitke, P. 2006, A&A, 455, 325 Hinkley, S., Pueyo, L., Faherty, J. K., et al. 2013, ApJ, 779, 153 Hubeny, I., Burrows, A., & Sudarsky, D. 2003, ApJ, 594, 1011 Ingraham, P., Marley, M. S., Saumon, D., et al. 2014, ApJL, 794, L15 Kalas, P., & Jewitt, D. 1995, Astronomical Journal, 110, 794 Kalas, P., Graham, J. R., Chiang, E., et al. 2008, Science, 322, 1345 Kirkpatrick, J. D. 2005, ARA&A, 43, 195 Kirkpatrick, J. D., Barman, T. S., Burgasser, A. J., et al. 2006, ApJ, 639, 1120 Kirkpatrick, J. D., Cruz, K. L., Barman, T. S., et al. 2008, Astrophys. J., 689, 1295 Kirkpatrick, J. D., Looper, D. L., Burgasser, A. J., et al. 2010, ApJS, 190, 100 Liu, M. C., Dupuy, T. J., & Allers, K. N. 2013, Astronomische Nachrichten, 334, 85 -. 2016, ApJ, 833, 96 Macintosh, B., Graham, J., Palmer, D., et al. 2006, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 6272, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Macintosh, B., Graham, J. R., Ingraham, P., et al. 2014, Proceedings of the National Academy of Science, 111, 12661 Macintosh, B., Graham, J. R., Barman, T., et al. 2015, Science, 350, 64 Maire, J., Ingraham, P. J., De Rosa, R. J., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 914785 Males, J. R., Close, L. M., Morzinski, K. M., et al. 2014, Astrophysical Journal, 786, 32 Marleau, G.-D., Klahr, H., Kuiper, R., & Mordasini, C. 2017, ArXiv e-prints, arXiv:1701.02747 Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2007, ApJ, 655, 541 Marois, C., Lafreni`ere, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, Astrophysical Journal, 641, 556 Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322, 1348 Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., & Barman, T. 2010, Nature, 468, 1080 Mayor, M., & Queloz, D. 1995, Nature, 378, 355 Millar-Blanchaer, M. A., Graham, J. R., Pueyo, L., et al. 2015, ApJ, 811, 18 Morzinski, K. M., Males, J. R., Skemer, A. J., et al. 2015, ApJ, 815, 108 Mouillet, D., Larwood, J. D., Papaloizou, J. C. B., & Lagrange, A. M. 1997, Monthly Notices of the RAS, 292, 896 Nielsen, E. L., Liu, M. C., Wahhaj, Z., et al. 2014, ApJ, 794, 158 Oppenheimer, B. R., Baranec, C., Beichman, C., et al. 2013, Astrophysical Journal, 768, 24 Perrin, M. D., Maire, J., Ingraham, P., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 91473J Pueyo, L. 2016, ApJ, 824, 117 Quanz, S. P., Meyer, M. R., Kenworthy, M. A., et al. 2010, ApJL, 722, L49 Rameau, J., Chauvin, G., Lagrange, A.-M., et al. 2013, Astrophysical Journal, Letters, 772, L15 Robert, J., Gagn´e, J., Artigau, ´E., et al. 2016, ArXiv e-prints, arXiv:1607.06117 Schneider, A. C., Cushing, M. C., Kirkpatrick, J. D., et al. 2014, AJ, 147, 34 Schneider, A. C., Windsor, J., Cushing, M. C., Kirkpatrick, J. D., & Wright, E. L. 2016, Astrophys. J., 822, L1 Smith, B. A., & Terrile, R. J. 1984, Science, 226, 1421 Soummer, R., Pueyo, L., & Larkin, J. 2012, Astrophysical Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk, M. H. 2010, Journal, Letters, 755, L28 Astrophysical Journal, 719, 497 Sparks, W. B., & Ford, H. C. 2002, Astrophysical Journal, 578, Lagrange, A.-M., Gratadour, D., Chauvin, G., et al. 2009, 543 Astronomy and Astrophysics, 493, L21 Lagrange, A.-M., Bonnefoy, M., Chauvin, G., et al. 2010, Spiegel, D. S., & Burrows, A. 2012, ApJ, 745, 174 Tokunaga, A. T., Simons, D. A., & Vacca, W. D. 2002, PASP, Science, 329, 57 114, 180 Lagrange, A.-M., Boccaletti, A., Milli, J., et al. 2012a, A&A, 542, A40 van Leeuwen, F. 2007, Astronomy and Astrophysics, 474, 653 Wagner, K., Apai, D., Kasper, M., et al. 2016, Science, 353, 673 18 Chilcote et al. Wang, J. J., Rajan, A., Graham, J. R., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 914755 Wang, J. J., Graham, J. R., Pueyo, L., et al. 2016, ArXiv e-prints, arXiv:1607.05272 Woitke, P., & Helling, C. 2003, A&A, 399, 297 19 Table 1. GPI observations of β Pic Date Observing Mode Exposure Time (s) Parallactic ◦ Rotation ( ) DIMM (cid:48)(cid:48) Seeing ( ) K1-Spec. 2013-11-16d,f 2013-11-16d,f K2-Spec. 2013-11-18a,c,d,e,f H-Spec. 2013-12-10d,f H-Spec. 2013-12-10b,d,e,f 2013-12-11d,f 2014-03-23d,f 2014-03-26f 2014-11-08d 2015-12-05h K2-Spec. H-Spec. Y-Spec. J-Spec. H-Spec. K1-Spec. 1789 1253 2446 1312 1597 556 1133 2923 2147 2002 26 18 32 38 18 64 47 26 25 37 1.09 0.93 0.68 0.77 0.70 0.46 0.47 0.86 0.77 1.12 aThis data set was originally astrometrically published by Macintosh et al. (2014) b This data set was origninally published by Bonnefoy et al. (2014) c This data set for a spectrum was published by Chilcote et al. (2015) dThis data set was originally astrometrically published by Millar-Blanchaer et al. (2015) e CCR power state was changed during observations f Observations taken during GPI verification & commitioning tests g AO performance parameters adjusted during GPI verification & commition- ing tests hData part of Gemini's Large and Long program (GS-2015B-LP-6) 20 Chilcote et al. Table 2. Measured β Pic b Parameters Reference Currie et al. (2013) Bonnefoy et al. (2013) Bonnefoy et al. (2014) Chilcote et al. (2015) Baudino et al. (2015) Morzinski et al. (2015) This work (Bolometric) † This work (Spectrophotometry) Teff log g (K) (cgs) 3.8±0.2 4.0±0.5 ∗ 1575 1700±100 1650±150 <4.7 1600–1700 1550±150 1708±23 1724±15 1700 3.5–4.5 3.5±1 4.2 4.18±0.01 4.0 Radius (RJup) 1.65±0.06 ∗ 1.5–1.6 1.5±0.2 ··· 1.76±0.24 1.45±0.02 1.46±0.01 1.41 Mass (MJup) 6.9 ∗ 9–10 ∗ <20 ··· 4.0 12.7±0.3 12.9±0.2 15.0 ∗ † Value calculated in Morzinski et al. (2015). Best fit from Drift-Phoenix and Spiegel & Burrows (2012) models Init. Spec. Entropy −1) (kB baryon ··· ≥9.3 >10.5 ··· ··· 9.75 9.75 21 Table 3. Best-fit atmospheric models Grid Name Data Used Ames-Dusty Photometry Only Teff (K) 1700 GPI Spectrum & Photometry 1800 BT-Settl (2015) Photometry Only 1800 GPI Spectrum & Photometry 1800 Drift Phoenix Photometry Only 1700 GPI Spectrum & Photometry 1700 Grid Points log g ([dex]) Radius (RJup) 3.5 3.5 3.0 3.5 3.5 4.0 1.31 1.17 1.38 1.22 1.41 1.41 χ2 ν 2.66 3.49 2.99 3.17 1.55 1.81 Teff (K) 1704 1706 1781 1815 1708 1651 Interpolated Grid log g ([dex]) Radius (RJup) 3.50 4.50 3.26 3.29 3.66 3.00 1.31 1.18 1.34 1.25 1.41 1.58 χ2 ν 2.66 3.45 2.63 3.04 1.54 1.21 Table 4. Best-fit combined evolutionary and atmospheric models Grid Name Data Used Mass (MJup) Grid Points Initial Entropy χ2 ν (kB/baryon) Interpolated Grid Mass (MJup) Initial Entropy χ2 ν (kB/baryon) Cloud-free (1× solar) Cloud-free (3× solar) Photometry Only Spectrum & Photometry Photometry Only Hybrid clouds (1× solar) Photometry Only Spectrum & Photometry Hybrid clouds (3× solar) Photometry Only Spectrum & Photometry Spectrum & Photometry 14.0 15.0 14.0 15.0 14.0 15.0 14.0 15.0 9.75 9.50 9.75 9.50 9.75 9.75 9.75 9.75 15.39 68.97 14.46 54.72 7.17 7.70 7.43 5.96 13.53 15.00 13.48 15.00 13.56 14.97 13.52 14.99 13.00 9.55 12.98 9.56 12.73 9.77 12.94 9.79 15.18 68.45 13.91 53.52 7.00 7.02 7.17 4.87
1911.12759
1
1911
2019-11-28T15:56:48
RefPlanets: Search for reflected light from extra-solar planets with SPHERE/ZIMPOL
[ "astro-ph.EP", "astro-ph.IM", "astro-ph.SR" ]
RefPlanets is a guaranteed time observation (GTO) programme that uses the Zurich IMaging POLarimeter (ZIMPOL) of SPHERE/VLT for a blind search for exoplanets in wavelengths from 600-900 nm. The goals of this study are the characterization of the unprecedented high polarimetic contrast and polarimetric precision capabilities of ZIMPOL for bright targets, the search for polarized reflected light around some of the closest bright stars to the Sun and potentially the direct detection of an evolved cold exoplanet for the first time. For our observations of Alpha Cen A and B, Sirius A, Altair, Eps Eri and Tau Ceti we used the polarimetric differential imaging (PDI) mode of ZIMPOL which removes the speckle noise down to the photon noise limit for angular separations >0.6". We describe some of the instrumental effects that dominate the noise for smaller separations and explain how to remove these additional noise effects in post-processing. We then combine PDI with angular differential imaging (ADI) as a final layer of post-processing to further improve the contrast limits of our data at these separations. For good observing conditions we achieve polarimetric contrast limits of 15.0-16.3 mag at the effective inner working angle of about 0.13", 16.3-18.3 mag at 0.5" and 18.8-20.4 mag at 1.5". The contrast limits closer in (<0.6") depend significantly on the observing conditions, while in the photon noise dominated regime (>0.6"), the limits mainly depend on the brightness of the star and the total integration time. We compare our results with contrast limits from other surveys and review the exoplanet detection limits obtained with different detection methods. For all our targets we achieve unprecedented contrast limits. Despite the high polarimetric contrasts we are not able to find any additional companions or extended polarized light sources in the data that has been taken so far.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. paper_ZIMPOL_contrast_sh05 December 2, 2019 c(cid:13)ESO 2019 RefPlanets(cid:63): Search for reflected light from extra-solar planets with SPHERE / ZIMPOL S. Hunziker1, H.M. Schmid1, D. Mouillet3, 4, J. Milli5, A. Zurlo18, 19, 16, P. Delorme3, L. Abe12, H. Avenhaus14, 1, A. Baruffolo15, A. Bazzon1, A. Boccaletti10, P. Baudoz10, J.L. Beuzit16, M. Carbillet12, G. Chauvin3, 17, R. Claudi15, A. Costille16, J.-B. Daban12, S. Desidera15, K. Dohlen16, C. Dominik9, M. Downing20, N. Engler1, M. Feldt14, T. Fusco16, 21, C. Ginski11, D. Gisler7, 8, J.H. Girard5, R. Gratton15, Th. Henning14, N. Hubin20, M. Kasper20, C.U. Keller11, M. Langlois22, 16, E. Lagadec12, P. Martinez12, A.L. Maire14, 25, F. Menard3, 4, M.R. Meyer26, A. Pavlov14, J. Pragt2, P. Puget3, S.P. Quanz1, E. Rickman24, R. Roelfsema2, B. Salasnich15, J.-F. Sauvage16, 21, R. Siebenmorgen20, E. Sissa15, F. Snik11, M. Suarez20, J. Szulágyi6, Ch. Thalmann1, M. Turatto15, S. Udry24, R.G. van Holstein11, A. Vigan16, and F. Wildi24 (Affiliations can be found after the references) Received -- ; accepted -- ABSTRACT Aims. RefPlanets is a guaranteed time observation (GTO) programme that uses the Zurich IMaging POLarimeter (ZIMPOL) of SPHERE/VLT for a blind search for exoplanets in wavelengths from 600-900 nm. The goals of this study are the characterization of the unprecedented high polarimetic contrast and polarimetric precision capabilities of ZIMPOL for bright targets, the search for polarized reflected light around some of the closest bright stars to the Sun and potentially the direct detection of an evolved cold exoplanet for the first time. Methods. For our observations of α Cen A and B, Sirius A, Altair,  Eri and τ Ceti we used the polarimetric differential imaging (PDI) mode of ZIMPOL which removes the speckle noise down to the photon noise limit for angular separations (cid:39)0.6(cid:48)(cid:48). We describe some of the instrumental effects that dominate the noise for smaller separations and explain how to remove these additional noise effects in post-processing. We then combine PDI with angular differential imaging (ADI) as a final layer of post-processing to further improve the contrast limits of our data at these separations. Results. For good observing conditions we achieve polarimetric contrast limits of 15.0 -- 16.3 mag at the effective inner working angle of ∼0.13(cid:48)(cid:48), 16.3 -- 18.3 mag at 0.5(cid:48)(cid:48)and 18.8 -- 20.4 mag at 1.5(cid:48)(cid:48). The contrast limits closer in ((cid:47)0.6(cid:48)(cid:48)) depend significantly on the observing conditions, while in the photon noise dominated regime ((cid:39)0.6(cid:48)(cid:48)), the limits mainly depend on the brightness of the star and the total integration time. We compare our results with contrast limits from other surveys and review the exoplanet detection limits obtained with different detection methods. For all our targets we achieve unprecedented contrast limits. Despite the high polarimetric contrasts we are not able to find any additional companions or extended polarized light sources in the data that has been taken so far. Key words. Techniques: polarimetric -- Planets and satellites: detection Instrumentation: high angular resolution -- Methods: data analysis -- Methods: observational -- Techniques: image processing -- 1. Introduction High-contrast imaging is a key technique for the search and clas- sification of extra-solar planets which is one of the primary goals in modern astronomy. However, the technical requirements are very challenging and up to now only about a dozen young, giant planets have been directly imaged (e.g. Macintosh et al. 2015; Bowler 2016; Schmidt et al. 2016; Chauvin et al. 2017; Keppler et al. 2018). Young, self-contracting giant planets are hot with temperatures of T ≈ 1000 − 2000 K (e.g. Baraffe et al. 2003; Spiegel & Burrows 2012), therefore they are bright in the near- infrared (NIR) and the required contrast C = Fpl/Fstar ≈ 10−5±1 is within reach of modern extreme adaptive optics (AO) systems, like SPHERE at the VLT (Beuzit et al. 2008), GPI at Gemini (Macintosh et al. 2014), the NGS AO system at Keck (van Dam et al. 2004) or SCExAO at Subaru (Jovanovic et al. 2015). Un- fortunately, young stars with planets are rare in the solar neigh- (cid:63) Based on observations made with ESO Telescopes at the La Silla Paranal Observatory under programme IDs: 095.C-0312(B), 096.C- 0326(A), 097.C-0524(A), 097.C-0524(B), 098.C-0197(A), 099.C- 0127(A), 099.C-0127(B), 0102.C-0435(A) bourhood. Furthermore, for the young stars in the nearest star forming regions at d ≈ 150 pc the expected angular separations of planets tend to already be quite small and hence they are ob- servationally challenging to detect. Most old planets, including all habitable planets, are cold and therefore produce only scattered light in the visual to NIR (<2 µm) wavelength range (Sudarsky et al. 2003). Light- scattering by the planets' atmosphere produces a polarization signal which can be distinguished from the unpolarized light of the much brighter central star (Seager et al. 2000; Stam et al. 2004; Buenzli & Schmid 2009). The contrast of this reflected light from extra-solar planets with respect to the brightness of their host stars is very challenging (C (cid:47) 10−7), but polarimetric differential imaging (PDI) has been shown to be a very effective technique to reveal faint reflected light signals. For these rea- sons the SPHERE "planet finder" instrument includes the Zurich IMaging POLarimeter (ZIMPOL, Schmid et al. 2018) which was designed for the search of light from reflecting planets in the vi- sual wavelength range using innovative polarimetric techniques (Schmid et al. 2006a; Thalmann et al. 2008). Article number, page 1 of 23 9 1 0 2 v o N 8 2 . ] P E h p - o r t s a [ 1 v 9 5 7 2 1 . 1 1 9 1 : v i X r a A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 We investigate in this paper the achievable contrast of SPHERE/ZIMPOL for a first series of deep observations of promising targets obtained within the RefPlanets project, which is a part of the guaranteed time observation (GTO) program of the SPHERE consortium. An important goal of this work is a better understanding of the limitations of this instrument in order to optimize the SPHERE/ZIMPOL observing strategy for high- contrast targets, and possibly to conceive upgrades for this in- strument or improve concepts for future instruments, for exam- ple for the Extremely Large Telescope (ELT, Kasper et al. 2010; Keller et al. 2010). Pushing the limits of high-contrast imaging polarimetry should be useful for the future investigation of many types of planets around the nearest stars, including Earth twins. The following subsections describe the expected polariza- tion signal from reflecting planets and the search strategy us- ing SPHERE/ZIMPOL. The GTO observations are presented in Section 2, and Section 3 discusses our standard data reduction procedures for ZIMPOL polarimetry. Section 4 provides the de- scription of the angular differential imaging method that we ap- plied to our data and the metric for the assessment of the point- source contrast. Section 5 shows our detailed search results for α Cen A. Section 6 discusses in more detail the physical mean- ing of the contrast limits and Section 7 presents our conclusions. In Appendix A and B we present the advanced data reduction steps necessary to reach the best possible polarimetric contrast limits with ZIMPOL and in Appendix C we present and discuss the detection limits for all other targets of our survey. 1.1. The polarization of the reflected light from planets The expected polarization signal from reflecting planets has been described with simple models (Seager et al. 2000), with de- tailed calculations for e.g. Jupiter and Earth-like planets (Stam et al. 2004; Stam 2008), or for a parameter grid of planets with Rayleigh scattering atmospheres (Buenzli & Schmid 2009; Bai- ley et al. 2018). The intensity and polarized intensity phase func- tions depending on the orbital phase angle φ and the planet-star- observer scattering angle α for one such model is illustrated in Fig. 1. These models are guided by polarimetric observations of Solar System objects, for which the typical fractional polar- ization is quite high p(α) > 10 % for visible wavelengths and scattering angles in the range α ≈ 60◦− 120◦ (e.g. Schmid et al. 2006a). Observations of individual objects have shown that for Rayleigh scattering atmospheres like Uranus and Neptune (Schmid et al. 2006b) the fractional polarization can be substan- tially higher than this value (p(90◦) > 20 %). For mostly haze scattering atmospheres as found on Titan (Tomasko & Smith 1982; Bazzon et al. 2014) or in the polar regions of Jupiter (Smith & Tomasko 1984; Schmid et al. 2011; McLean et al. 2017) the fractional polarization can even reach values up to p(90◦) ≈ 50 %. On the other hand, the Mie scattering process in the clouds that dominate the atmospheres of Venus, Saturn or the equatorial regions of Jupiter produces a lower polarization in the visual wavelengths < 10 % (Smith & Tomasko 1984; Hansen & Hovenier 1974). And for larger objects without any significant atmosphere like Mercury, Moon, Mars and other rocky bodies (e.g. Dollfus 1985) the polarization of the reflected light is some- where in between p(90◦) ≈ 5−20 %. Finally, for the polarization of Earth Bazzon et al. (2013) determined fractional polarizations of about 19 % in V-band and 13 % in R-band mainly caused by Rayleigh scattering in the atmosphere. For Rayleigh scattering, haze scattering, and the reflec- tion from solid planet surfaces, the resulting polarization for Article number, page 2 of 23 α ≈ 30◦− 150◦ is perpendicular to the scattering plane, just like illustrated in Fig. 1(a). This means that for extra-solar planets the polarization is usually positive in perpendicular direction to the line connecting star and planet as projected onto the sky. The polarization, however, can be negative for the reflection from clouds as observed for Venus (Hansen & Hovenier 1974), or for reflections with small scattering angles (α (cid:47) 25◦) on rocky or icy surfaces (Dollfus 1985). 1.2. The signal from extra-solar planets The signal of a reflecting planet depends on the surface prop- erties, which define the reflectivity I(α) and the fractional po- larization p(α) of the planet, as well as the planet size and its separation from the central star. The reflectivity and polarization depend on the scattering angle α given by the orbital phase φ and orbit inclination i as sketched in Fig. 1(a). We set the phase φ = 0 in conjunction, when the planet illumination as seen by the observer is maximal. For circular orbits the dependence is α = arccos(sin i · cos φ) (1) and the scattering angle varies between a minimum and max- imum value αmin and αmax as indicated in Fig. 1(a). For edge-on orbits (i = 90◦), Eq. (1) simplifies to α = φ for φ = −180◦ to 180◦, and for pole-on systems (i = 0◦) we see one single scat- tering angle α = 90◦ during the whole orbit. Small and large scattering angles α (cid:47) 30◦ and α (cid:39) 150◦ are only observable for strongly inclined orbits i > 60◦, but at the corresponding phase angles, planets are typically faint in polarized flux (see Fig. 1(b)), in addition, the angular separation is small and therefore a suc- cessful detection will be particularly difficult (e.g. Schworer & Tuthill 2015). For Rayleigh-like scattering the fractional polarization p(α) is highest around α ≈ 90◦ while the reflectivity I(α) is increas- ing for α → 0◦. Therefore the maximum polarized intensity p(α) I(α) is expected for a scattering angle α ≈ 60◦. The full dependence of the normalized intensity I(φ) and polarized inten- sity p(φ) I(φ) as function of orbital phase for a Rayleigh scatter- ing planet is illustrated in Fig. 1(b). The figure shows simulated phase functions for planets on circular orbits with inclinations of i = 0◦, 30◦, 60◦, and 90◦. The model in Fig. 1(b) was selected from the model grid of Rayleigh scattering atmospheres derived in Buenzli & Schmid (2009). We use it as a reference case for the reflected intensity and polarization of a planetary atmosphere. The model is simi- lar to Uranus and Neptune which are quite favourable cases for a polarimetric search for planets. A giant planet might have a thinner Rayleigh scattering layer τsc < 1 on top of a cloud layer, resulting in a lower polarization fraction (see Buenzli & Schmid 2009). This is because the reflection from a cloud layer produces significantly less polarization than the reflection from a thick (τsc (cid:39) 1) Rayleigh scattering layer. The model shown in Fig. 1(b) has an optical depth of τsc = 2 for the Rayleigh scattering layer, with a single scattering albedo of ω = 0.95 above a cloud layer approximated by a Lambertian surface with an albedo of AS = 1. This model yields for quadrature phase α = 90◦ a reflectivity of I(90◦) = 0.131 and a corresponding polarized signal amplitude of p(90◦)·I(90◦) = 0.055. The parameter p(90◦) is a good way of characterizing the polarization of an extra-solar planet because planets at all inclinations will pass through this phase at least twice. In this phase , the fractional polarization is expected to be close to the maximum and the apparent separation from the star is maximized for planets on circular orbits (see Fig. 1(b), 2). Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL (a) (b) Fig. 1. (a) Diagram showing the essential planes and angles needed to characterize the reflected light intensity: the orbital phase of the planet φ, the inclination of the orbital plane with respect to the sky plane i and the scattering angle α. The scattering angle α is measured along the scattering plane and our definition of the direction of a positive polarization p(α) is perpendicular to this plane. In special cases the polarization of the reflected light could be negative, this would correspond to a direction of p(α) perpendicular to the red arrows. (b) Normalized intensity and polarized intensity of the reflected light as function of the orbital phase φ, calculated with the reference planet atmosphere model used in this paper: a Rayleigh scattering atmosphere from (Buenzli & Schmid 2009) with an optical depth of τsc = 2, a single scattering albedo of ω = 0.90, and a ground surface (= cloud) albedo of AS = 1. The different colors show the phase functions of planets on circular orbits seen at four different inclinations: 0◦ (black), 30◦ (red), 60◦ (green), 90◦ (blue). The polarization p(α) refers to the amplitude of the polar- ization but our raw data consists of independent measurements of the Stokes Q and U parameters. Since we can assume that the reflected light from a planet is polarized along the axis per- pendicular to the connecting line between star and planet, we use the transformation into polar coordinates from Schmid et al. (2006b) to derive Qφ and Uφ. In the dominating single scatter- ing scenario, the tangential polarization Qφ should contain all the polarized intensity of the reflected light, while Uφ should be zero everywhere. Because of this relationship we will refer to Qφ as the polarized intensity throughout this work. The key parameter for the polarimetric search of reflecting extra-solar planets is the polarization contrast Cpol, this is the polarized flux Qφ from the planet relative to the total flux from the central star: Cpol(α) = p(α) · Cflux(α) = p(α) · I(α) R2 p d2 p , (2) where Rp is the radius of the planet, dp the physical sepa- ration between planet and star and I(α) and p(α) the reflectivity and fractional scattering polarization for a given scattering angle, respectively. In this notation, the reflectivity I(α = 0) is equiva- lent to the geometric albedo Ag of a planet. The ratio R2 p for a Jupiter-sized planet with radius RJ at a separation dp = 1 AU J/AU2 = 2.3 · 10−7 and the total polarization contrast of a is R2 planet with our reference model with p(90◦) · I(90◦) = 0.055 would be of order Cpol ≈ 10−8. A Neptune-sized planet would p/d2 have to be located at about 0.5 AU to produce the same polariza- tion contrast. With increasing physical separation dp the contrast decreases rapidly with 1/d2 p (see Eq. (2)). With increasing dis- tance to the star, the angular separation ρ of a planet at a constant dp also decreases. Thus moving it closer to the star where high contrasts cannot be maintained. The combination of both effects limits the sample of possible targets for a search of reflected light to the most nearby stars. In addition to that, the sample is limited to only the brightest stars because photon noise increases like F with the lower photon flux F of stars that are fainter in 1/ the visible wavelengths. √ 1.3. Targets for the search of extra-solar planets The detection space for our SPHERE/ZIMPOL high-contrast ob- servations starts at about ρ ≈ 0.1(cid:48)(cid:48), and the current polarimet- ric contrast limits after post-processing are of the order 10−7 for ρ < 0.5(cid:48)(cid:48) and 10−8 for ρ > 0.5(cid:48)(cid:48). Therefore only the nearest stars within about 5 pc can have a bright enough reflecting planet with Rp ≈ RJ and a contrast of Cpol (cid:39) 10−8 with a sufficiently large angular separation ρ > 0.1(cid:48)(cid:48) for a successful detection. Based on these criteria, some of the best stellar systems for the search of a Jupiter-sized planet in reflected light with SPHERE/ZIMPOL are α Cen A and B, Sirius A,  Eri, τ Cet, Altair and a few others as determined by Thalmann et al. (2008). No extra-solar planet is known to exist around these high priority stars which would fulfil the above detection limit cri- teria. There is strong evidence from radial velocity and astro- metric studies for the presence of a giant planet in  Eri (e.g. Article number, page 3 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 pends on the orbital parameters and the distance of the systems and our example α Cen A system would show for a planet the fastest angular orbital motion for a given orbital separation be- cause of its proximity. Without going into details, already the α Cen A example in Fig. 2 illustrates, that planets on inclined orbits have phases with large separation when they are relatively bright and easy to de- tect, and phases where they are close to the star and faint and challenging to detect. Therefore, a blind search provides only planet detection limits valid for that observing date. One should also notice that data taken during different nights cannot simply be coadded for the search of extra-solar planets due to the ex- pected short orbital periods. Instead it would be necessary to use a tool like K-Stacker (Nowak et al. 2018) that combines the re- sults from multiple epochs while considering the orbital motion of a planet. 2. Observations 2.1. The SPHERE/ZIMPOL instrument The polarimetric survey for extra-solar planets was carried out with the SPHERE "Planet Finder" instrument (Beuzit et al. 2008, 2019) on VLT Unit Telescope 3 (UT3) of the European Southern Observatory. SPHERE is an extreme adaptive optics system with a fast tip-tilt mirror and a fast high-order deformable mirror with 41x41 actuators and a Shack-Hartman wave-front sensor (e.g. Fusco et al. 2006). The system includes an image de-rotator, at- mospheric dispersion correctors, calibration components and the IRDIS (Dohlen et al. 2008), IFS (Claudi et al. 2008) and ZIM- POL focal plane instruments for high-contrast imaging. This program was carried out with ZIMPOL which was specifically designed for the polarimetric search of reflected light from extra-solar planets around the nearest, bright stars in the spectral range 500-900 nm. The SPHERE/ZIMPOL system is described in detail in Schmid et al. (2018) and we highlight here some of the important properties for high-contrast imaging of reflected light from planets: -- the polarimetric mode is based on a fast modulation - demod- ulation technique which reaches a polarimetric sensitivity1 of ∆p < 10−4 (Schmid et al. 2018) in the light halo of a bright star. This is possible because the used modulation frequency of 968 Hz is faster than the seeing variations and therefore the speckle noise suppression for PDI is particularly good as long as the coherence time τ0 is greater than about 2 ms. This condition was usually satisfied during the RefPlanets obser- vations (see Table 1). -- ZIMPOL polarimetry can be combined with coronagraphy for the suppression of the diffraction limited PSF peak of the bright star, for a sensitive search of faint point-sources in the light halo of a bright star. -- ZIMPOL has a small pixel scale of 3.6 mas/pix, a detector mode with a high pixel gain of 10.5 e− ADU−1 and a full well capacity of 640 ke− pix−1. This allows to search for very faint polarized signals in coronagraphic images of very bright stars mR < 4m with broad-band filters by "just" push- ing the photon noise limit thanks to the photon collecting power of the VLT telescope. The combination of high-contrast imaging using AO and coronagraphy provides for point-sources a raw contrast at a level 1 Degree of suppression of the light by the polarimetry Fig. 2. Apparent positions of a model planet on a circular 80◦ inclined orbit with r = 1 AU around α Cen A in a typical coronagraphic intensity (left) and polarization frame (right) at 10 day intervals. The brightness of the planet signal with respect to α Cen A is exaggerated by a factor of 104 for the intensity and 103 for the polarization. The relative brightness of the point for different phases is according to the model presented in Fig. 1. Hatzes et al. 2000; Mawet et al. 2019), but the derived separa- tion is 3 AU and therefore the expected signal is at the level of only Cpol ≈ 10−9. For τ Cet, the presence of planets has been proposed based on radial velocity data (Feng et al. 2017), but none is expected to produce a contrast Cpol (cid:39) 10−9. The radial velocity constraints for the A-stars Sirius A and Altair are very loose because their spectra are not well suited for sensitive radial velocity searches, and undetected giant planets at 1 AU may be present. The radial velocity limits for planets are very stringent for α Cen B (Zhao et al. 2018), but less well constrained for α Cen A (Zhao et al. 2018). However, the simple calculation of the reflected light contrast does not consider the possibility that a planet could be exceptionally bright due to certain reasons, e.g. an extensive ring system surrounding the planet (e.g. Arnold & Schneider 2004). Because of the absence of obvious targets, we decided to carry out an exploratory blind search for "unexpect- edly" bright companions, with the additional aim to investigate the detection limits of this instrument and to define the best ob- serving strategies for possible future searches. For such a survey, one needs to consider that planets around the nearest stars are moving fast through our field-of-view (FOV). This is illustrated in Fig. 2, which simulates the orbit of a planet with a circular orbit with a separation of 1 AU around α Cen A on top of single coronagraphic intensity or polarimet- ric frames. The individual points are the orbital positions of this model planet separated by 10 days. The relative brightness of the points are calculated for an orbit inclination of i = 80◦ coplanar with the α Cen binary (Kervella et al. 2016) and using the same Rayleigh scattering atmosphere model as in Fig. 1, but with the brightness upscaled by a factor of 104 for the intensity and 103 for the polarization to make the dots visible on top of a single coronagraphic observation. Of course, the angular motion de- Article number, page 4 of 23 Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL 10−4 − 10−5, while polarimetry in combination with angular dif- ferential imaging (ADI) yields a further contrast improvement for polarimetric differential imaging of about 10−3, so that a to- tal contrast of Cpol ≈ 10−8 is reachable with sufficiently long integrations. 2.2. Observations In Table 1 we list all observations which were carried out so far for the RefPlanets GTO program. We observed six of the most favourable targets in the solar neighbourhood identified by Thalmann et al. (2008) as ideal targets for the search of planets in reflected light. All data were taken with the fast modulation polarimetry mode, which is the mode of choice for high flux applications. The first observations in 2015 were made with different filters in camera 1 and camera 2 of ZIMPOL. But it was noticed that some disturbing polarimetric residuals can be corrected if the simulta- neous camera 1 and camera 2 frames are taken with the same filter passband, because the residuals have opposite signs and compensate when camera 1 and camera 2 frames taken with the same filter are combined. Of course, the contrast also improves with the combination of data from both cameras because of the lower photon noise limit. From 2016 onwards we took for each hour of coronagraphic observations one or two short polarimet- ric cycles with the star offset from the focal plane mask for the calibration of the flux, the point-spread function (PSF), and the polarimetric beam shift (Schmid et al. 2018). These PSFs were taken with neutral density (ND) filters to avoid detector satura- tion. The main criterion for the filter selection is a high photon throughput. Filters with broader passbands provide more pho- tons and stars with mR > 1m were observed usually in the VBB filter (λc,VBB = 735 nm, ∆λVBB = 291 nm). For Sir- ius A, α Cen A and Altair we used filters with smaller band widths to avoid detector saturation with the minimum detector integration time of 1.1 s available for ZIMPOL, namely, the R_PRIM (λc,R_PRIM = 626 nm, ∆λR_PRIM = 149 nm), N_R (λc,N_R = 646 nm, ∆λN_R = 57 nm) and N_I (λc,N_I = 817 nm, ∆λN_I = 81 nm) filters. Only for τ Ceti we deviated from this strategy and chose the R_PRIM filter instead of the VBB filter because we noticed that certain disturbing wavelength depen- dent instrumental effects (instrumental polarization, beam shift) are easier to correct during the data reduction for data narrower passbands. Almost all objects were observed with SPHERE/ZIMPOL in P1-mode, in which the image de-rotator is fixed. In this mode the sky rotates as a function of the telescope parallactic angle and altitude allowing for ADI (Marois et al. 2008) in combina- tion with PDI because most of the strong aberrations -- mainly caused by the deformable mirror (DM) -- are fixed with respect to the detector. The P1-mode stabilizes the instrument polariza- tion after the HWP2-switch, but does not stabilize the telescope pupil, which still rotates with the telescope altitude. Therefore, speckles related to the telescope pupil cannot be suppressed with ADI. We observed only  Eri in the field-stabilized polarimet- ric P2-mode to make use of the improved capability of the in- strument to detect weak extended scattering polarization from circumstellar dust which could be detectable with our FOV of 3.6(cid:48)(cid:48) × 3.6(cid:48)(cid:48) (e.g. Backman et al. 2009; Greaves et al. 2014). For all observations we used the medium sized classical Lyot coronagraph CLC-MT-WF with a dark focal plane mask spot deposited on a plate with a radius corresponding to 77.5 mas (Schmid et al. 2018), however, the effective inner working an- gle (IWA) of the reduced data is generally larger and depends on the star centering accuracy and stability. The spot in this coron- agraphic mask has a transparency of about 0.1 % (Schmid et al. 2018) and during good conditions and with good centering the star is visible behind the coronagraph so that an accurate center- ing of the frames in possible. Our usual observing strategy for deep coronagraphic obser- vations consists of one-hour blocks with about five to ten polari- metric cycles. Each cycle consists of observations with all four half-wave plate orientations (Q+, Q−, U+, U−). Between these blocks we took short non-coronagraphic cycles with a neutral density filter, by offsetting the star from the coronagraphic mask, to acquire samples of the unsaturated PSF for image quality as- sessments, flux calibrations, and the measurement of the beam shift effect. 3. Basic data reduction The data reduction is mainly carried out with the IDL-based sz-software (SPHERE/ZIMPOL) pipeline developed at ETH Zurich. Basic data preprocessing, reduction and calibration steps are essentially identical to the ESO Data Reduction and Han- dling (DRH) software package developed for SPHERE (Pavlov et al. 2008). The basic steps are described briefly in this subsec- tion and more technical information is available in Schmid et al. (2012, 2018). In addition to that, we describe in the appendix the more advanced sz-pipeline routines and additional data re- duction procedures required especially for high-contrast imaging and polarimetry. The fast modulation and on-chip demodulation imaging po- larimetry of ZIMPOL produces raw frames where the simulta- neous I⊥ and I(cid:107) polarization signals are registered on alternating rows of the CCD detectors. Basically, the ZIMPOL raw polar- ization signal QZ is the difference of the "even-row" I⊥ and the "odd-row" I(cid:107) subframes QZ = I⊥ − I(cid:107). The raw intensity signal is derived from adding the two subframes IZ = I⊥ + I(cid:107). Just like for any other CCD detector data, the basic data re- duction steps include image extraction, frame flips for the cor- rect image orientation, a first bias subtraction based on the pre- and overscan pixel level, bias frame subtraction for fixed pattern noise removal, and flat-fielding. Special steps for the ZIMPOL- system are the differential polarimetric combination of the sub- frames, taking into account the alternating modulation phases for the CCD pixel charge trap correction (Gisler et al. 2004; Schmid et al. 2012), and calibrating the polarimetric efficiency pol or modulation-demodulation efficiency. The polarimetric combina- tion of the frames of a polarimetric cycle Q+, Q−, U +, U− taken with the four half-wave plate orientations is again done in a stan- dard way. For non-field stabilized observations, the data combi- nation must also consider the image rotation. As basic data prod- uct of one polarimetric cycle one obtains four frames IQ, Q, IU, and U, which can be combined with the frames from many other cycles for higher signal-to-noise ratio (SNR) results. The basic PDI data reduction steps listed above are not suf- ficient for reaching the very high polarimetric contrast required for the search of reflecting planets. Especially at smaller sepa- rations (cid:47) 0.6(cid:48)(cid:48) the noise is still dominated by residuals of order 10−6 in terms of contrast compared to the brightness of the star (see Fig. 3). This is why we additionally apply more advanced calibration steps described in Appendix A and B. The steps in- clude: -- Frame transfer smearing correction -- Telescope polarization correction -- Correction of the differential polarimetric beam shift Article number, page 5 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Table 1. Summary of all RefPlanets observations completed until the end of 2018. For each observation we also list observing conditions (seeing in arc seconds and coherence time τ0 in ms) and the total field rotation (relevant for the efficiency of angular differential imaging). Date (UT) Object 2015/05/01 2015/05/02 2016/02/17 2016/02/20 2016/04/18 2016/04/21 2016/06/22 2016/07/21 2016/07/22 2016/10/10 2016/10/11 2016/10/12 2017/04/30 2017/05/01 2017/06/19 2018/10/14 2018/10/15 2018/10/16 2018/10/19 α Cen A α Cen B Sirius A Sirius A α Cen A α Cen A α Cen B Altair Altair  Eri  Eri  Eri α Cen A α Cen A α Cen B τ Ceti τ Ceti τ Ceti τ Ceti mR (mag) -0.5 1.0 -1.5 -1.5 -0.5 -0.5 1.0 0.6 0.6 3.0 3.0 3.0 -0.5 -0.5 1.0 2.9 2.9 2.9 2.9 Filters DIT (sec) cam2 cam1 N_R 1.2 N_I R_PRIM 1.2 VBB N_I 1.2 N_I 1.2 N_I N_I 1.2 N_R N_R 1.2 N_R N_R VBB VBB 1.1 R_PRIM R_PRIM 1.2 R_PRIM R_PRIM 1.2 3.0 VBB 3.0 VBB 5.0 VBB 1.2 N_R N_R 1.2 VBB 1.1 R_PRIM R_PRIM 14 R_PRIM R_PRIM 14 R_PRIM R_PRIM 14 R_PRIM R_PRIM 14 VBB VBB VBB N_R N_R VBB # of pol. cycles 66 90 34 73 40 80 74 63 30 42 48 15 84 39 141 30 24 32 29 a texp 2h 38.4min 3h 36min 1h 21.6min 2h 55.2min 1h 36min 3h 12min 2h 42.8min 2h 31.2min 1h 12min 2h 48min 3h 12min 1h 40min 3h 21.6min 1h 33.6min 3h 26.8min 2h 48min 2h 14.4min 2h 59min 2h 42.4min Seeing ((cid:48)(cid:48)) 0.6 -- 1.0 0.6 -- 1.1 0.7 -- 2.0 1.0 -- 2.0 1.1 -- 1.5 0.7 -- 1.7 0.3 -- 0.8 0.4 -- 0.8 0.4 -- 0.7 0.5 -- 0.8 0.5 -- 0.8 1.0 -- 1.8 0.5 -- 0.7 0.8 -- 1.3 0.3 -- 1.0 0.4 -- 0.7 0.6 -- 1.6 0.6 -- 1.0 0.6 -- 1.4 τ0 (ms) 1.7 -- 2.5 1.5 -- 2.2 1.6 -- 2.5 1.8 -- 3.5 1.8 -- 2.4 2.0 -- 4.0 3.5 -- 6.0 4.5 -- 7.0 3.0 -- 5.0 4.5 -- 6.7 3.3 -- 6.5 1.8 -- 2.1 3.0 -- 4.4 2.0 -- 2.5 4.5 -- 9.5 5.0 -- 10 2.0 -- 4.0 2.5 -- 3.7 2.6 -- 5.3 Air mass 1.24 -- 1.47 1.24 -- 1.46 1.01 -- 1.13 1.01 -- 1.39 1.26 -- 1.48 1.24 -- 1.71 1.24 -- 1.60 1.20 -- 1.44 1.20 -- 1.32 1.05 -- 1.37 1.04 -- 1.18 1.04 -- 1.21 1.24 -- 2.21 1.37 -- 1.84 1.24 -- 1.52 1.01 -- 1.12 1.01 -- 1.12 1.01 -- 1.28 1.01 -- 1.60 Field rotation (◦) 90.4 97.1 101.1 113.2 46.7 107.9 106.6 67.2 27.6 0b 0b 0b 121.5 47.8 115.6 130.5 112.1 104.9 107.5 Notes. The datasets with the deepest limits for each target are marked in bold font. (a) The total exposure time per camera. (b)  Eri was observed in the field stabilized ZIMPOL P2-polarimetry mode. Fig. 3. The 1σ radial contrast levels for the data shown after the different major data reduction steps -- basic PDI by ZIMPOL, correction of the polarimetric beam shift and subtraction of instrument polarization (IP) -- for the polarized intensity Q and for the corresponding intensity I of one single combined zero-phase and π-phase (2 × 1.2 s) exposure of α Cen A in the N_R filter. 4. Post-processing and the determination of the contrast limits There is still a landscape of residual noise visible after the basic data reduction, beam shift and frame transfer smearing correc- tion, and the subtraction of the residual instrument polarization. This can be seen for example in bottom panel in Fig. A.3. We show this quantitatively with 1σ noise levels for a series of short 2.4 s exposures measured after the different reduction steps in Fig. 3. After the full data reduction, the residual noise at small separations < 0.6(cid:48)(cid:48) still dominates the photon noise by a fac- tor of about 2 -- 5 in this particular example. For larger separa- tions > 0.6(cid:48)(cid:48) the residual noise is close to the photon noise limit. In a effort to further reduce the noise at small separations, we used a principle component analysis (PCA) (Amara & Quanz Article number, page 6 of 23 Fig. 4. Total intensity (Stokes I) and polarized intensity (Stokes Qφ) for the complete dataset of α Cen A in the N_R filter. The frames in the bottom row show a closer look at the speckle-dominated region closer to the star after injecting artificial point-sources (black circles) and ap- plying PCA-ADI with 20 PCs. 2012) based ADI algorithm to model the fixed and slowly vary- ing residual noise features. 4.1. PCA based ADI The top row in Fig. 4 shows the coronagraphic total intensity and polarized intensity data of α Cen A from June 2017 after de-rotating and combining all frames. The intensity exhibits a Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL 4.2. Polarimetric point-source contrast The contrast limits after the PCA-ADI step were calculated using artificial point-sources arranged in a spiral pattern around the star which we tried to recover with SNR=5. We determined the SNR according to the methods derived in Mawet et al. (2014), includ- ing the correction for small sample statistics. The artificial planet PSF was simulated with a non-coronagraphic PSF from one of the beam shift measurements, upscaled with the mean value of the transmission curve for the neutral density filter that was used to avoid saturation. We visually selected the non-coronagraphic PSF that best fits the shape of the coronagraphic PSF of the com- bined intensity image at separations > 0.3(cid:48)(cid:48). This ensures that we do not severely over- or underestimate the aperture flux that a point-source would have in our data, and therefore ensures ac- curate contrast limit estimations. For the α Cen A data, we show the radial profiles of both PSFs in Fig. 5, normalized to the num- ber of counts on the detector per second and pixel. The aperture radius rap used for the contrast estimation and SNR calculation was optimized for high SNR under the assump- tion that the searched point-source is weak compared to the PSF of the central star and read-out noise is negligible. With increas- ing rap, the number of counts from a faint source increases, how- ever, a larger aperture also has an increased background noise σbck ∝ rap. We derived an optimized rap ≈ λ/D, corresponding to about 4-6 pixels (14-22 mas) depending on the observed wave- length. The flux in each aperture was background subtracted in- dividually using the mean value of the pixels in a two pixel wide concentric annulus around the aperture, because the point-source contrast should not be affected by residual, non-axisymmetric, large scale structures in the image (e.g. stray light from α Cen A in the observations of α Cen B). We also calculated raw contrast curves for both Stokes I and Qφ without PCA-ADI to investigate how the other advanced data reduction steps improve the contrast limits at different separa- tions. The calculated raw contrast curves do not require the in- sertion of fake signals and are independent on the field rotation. Therefore, the raw contrast is more suitable for assessing the quality of small subsets of the data or even single exposures. For the raw contrast we also used methods derived in Mawet et al. (2014) to calculate the noise at different separations to the star and turn this into the signal aperture flux required for a de- tection. The detection threshold was set to a constant false posi- tive fraction (FPF) corresponding to the FPF of an Nσ detection with Gaussian distributed noise. The required aperture flux was then turned into a contrast limit estimation by dividing it through the aperture flux of the unsaturated stellar PSF. In order to apply the signal detection method described in Mawet et al. (2014), the underlying distribution of noise aperture fluxes has to be approximately Gaussian. We applied a Shapiro- Wilk test and found that this condition is satisfied for all separa- tions. 5. Results for α Cen A We present a detailed analysis of the results from the deepest ob- servations of α Cen A. We derive contrast limits and analyse the properties of the noise at different separations and for different total DITs. This detailed analysis shows the outstanding perfor- mance of ZIMPOL in terms of speckle suppression in PDI mode. In addition to that, we present and discuss the best results for all other targets of our survey in Appendix C. Article number, page 7 of 23 Fig. 5. The coronagraphic PSF with the CLC-MT-WF coronagraph (blue) compared to the non-coronagraphic PSF of α Cen A in the N_R filter (red). The coronagraphic PSF was upscaled by a factor of ∼ 8· 103 to account for the use of a neutral density filter during the measurement. PSF speckle halo with a strong radial gradient over two orders of magnitude. The differential polarization shows arc-like patterns in the de-rotated and combined image which originate from the de-rotated fixed residual noise pattern. ADI can be used to ef- ficiently model and subtract such large scale patterns before de- rotating and combining the images. The PCA-ADI approach was used successfully before by van Holstein et al. (2017) to improve the contrast limits of SPHERE/IRDIS polarimetry data. We used a customized version of the core code from the Pyn- Point pipeline (Amara & Quanz 2012; Stolker et al. 2019) for the ADI process. The complete speckle subtraction process was applied to the stacks of Q+, Q−, U+, U− and intensity frames separately after preprocessing and centering the frames. For the polarized intensity frames we applied PCA in an annulus around the star from 0.1(cid:48)(cid:48)to 1(cid:48)(cid:48)in order to cover the speckle-dominated region. For the total intensity frames we increased the outer ra- dius to 1.8(cid:48)(cid:48)since the whole FOV is dominated by speckles and other fixed pattern noise. We used a fixed number of 20 principle components (PCs), or 10 PCs in the case of the τ Ceti polarime- try, to model and subtract the residual noise patterns because this seemed to be the sweet-spot that produced deep contrast limits at most separations. In the bottom row of Fig. 4 we show an exam- ple for the result after removing 20 PCs from the intensity and polarized intensity frames. In both cases the contrast improved significantly. Typical contrast limit improvements for PCA-ADI were between a factor of 5 -- 10 for the total intensity and up to a factor of 3 for the polarized intensity. In Fig. 4, the resulting im- ages after PCA-ADI also contain a number of artificial planets with SNR≈5, they were introduced for estimating the contrast limits after PCA-ADI. For the de-rotation and combination of the frames, we ap- plied the noise-weighted algorithm as described in Bottom et al. (2017). This algorithm is simple to implement in a direct imag- ing data reduction pipeline and it often improved the SNR of the artificial planets significantly, with typical SNR gains of about 8% and a maximum gain up to 26% for our α Cen A test dataset. 5.1. Total intensity and polarized intensity 5.3. Companion size limit A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 The deepest observations of α Cen A were carried out in the N_R filter and in P1 polarimetry mode during a single half-night with good observing conditions (see Table 1). For the full data reduc- tion we used the best 84 out of 88 polarimetric cycles with a total texp of 201.6 min for each camera. This is our longest exposure time with a narrow-band filter. We combined the results of both √ cameras to improve the photon noise limit by an additional fac- tor of around 2. The resulting de-rotated and combined images are shown in Fig. 4. Just as described in Sec. 1.2, we transformed the polarized intensity frames Q, U into the Qφ, Uφ basis. We ex- pect a positive Qφ and no Uφ signal from the reflected light of a companion. The bottom panels in Fig. 4 show the inner, speckle- dominated region after inserting four artificial point-sources in the lower left corner and subsequently removing 20 PCs modes. The final image for Qφ is clean and shows no disturbing resid- uals except for a few very close to the coronagraph. However, the total intensity shows some strong disturbing features that are extended in the radial direction. These features are residu- als from the diffraction pattern of the rotating telescope spiders. The residuals are unpolarized and hence mostly cancelled in the Qφ result. 5.2. Contrast curve Figure 6(a) shows the 1σ and 5σ contrast limits for polarized intensity Qφ together with the 1σ photon noise limit. The con- trast is limited by speckle noise when the photon noise is lower than the measured 1σ point-source contrast, which is the case for separations (cid:47) 0.6(cid:48)(cid:48), corresponding to (cid:47) 1 AU for α Cen A. The solid green line shows the 5σ contrast limits after applying the basic data reduction steps without beam shift correction and residual instrument polarization subtraction, the solid red line includes both additional corrections. The symbols show the cor- responding contrast improvements after additional PCA speckle subtraction. The additional corrections -- including PCA-ADI -- improve the contrast limits mostly in the speckle noise dominated region close to the star at separations (cid:47) 0.5(cid:48)(cid:48). The contrast can be im- proved to about 2 -- 5 times the fundamental limit due to photon noise for these separations. For separations (cid:39) 0.5(cid:48)(cid:48) the improve- ment for the polarized intensity is zero but the limits are already close to the photon noise and ADI could only make it worse. This is why we have chosen to apply ADI in combination with PDI only in an annulus instead of applying it to the whole frame. √ The solid blue line in Fig. 6(b) is the contrast limit for the total intensity. The corresponding photon noise limit for the in- tensity is a factor of 2 lower than the photon noise limit for the polarization shown in Fig. 6(a) because only 50% of the photons contribute to the polarized signal Qφ. For the total intensity con- trast we also applied PCA-ADI and calculated the resulting con- trast limits inside 1(cid:48)(cid:48)and at 1.5(cid:48)(cid:48). The results show that speckle noise dominates at all separations. The PCA-ADI procedure can be used to improve the limits but they still exceed the photon noise limit by factors of about 100-1000. However, the detec- tion limits for the total intensity could be further improved with the ZIMPOL pupil stabilized imaging mode without polarimetry. This should produce better contrast limits for the same exposure time. Article number, page 8 of 23 The detection limits can be turned into size upper limits for a planet with some assumptions about its reflective properties and orbital phase. We adopt again the reference model from Sec. 1.2 with Q(90◦) = p(90◦)· I(90◦) = 0.055 and use the contrast curve from Sec. 5.2 to calculate the upper radius limits for a compan- ion that would still be detectable with polarimetry. The 5σ lim- its shown in Fig. 7 result in sizes smaller than 1 RJ for small separations ∼ 0.2 AU (0.15(cid:48)(cid:48)) and stay between 1 -- 1.5 RJ within the whole FOV. The sensitivity improves considerably towards smaller separations (short period planets) because the brightness of the planet scales with d−2 P . Companions larger than the calcu- lated limits should be detectable with an average SNR of at least 5. For comparison we also show what size the 1σ photon noise limit corresponds to. The radius limits in Fig. 7 are proportional to (p(α) · I(α))−1/2, therefore improving for planets with higher reflectivity and fractional polarization. 5.4. Contrast gain through longer integration One simple way of improving the achievable contrast limits is through longer integration texp. Especially if photon noise dom- inates, the detections limits should be proportional to t−1/2 exp . At small separations from the star the noise is dominated by the noise residuals that were not eliminated perfectly in the PDI step. This can be seen for example in the bottom frame of Fig. A.2. Some of the aberrations -- especially the ones to the right and left of the coronagraph caused by the deformable mirror (DM) -- are quasi-static throughout the observation (Cantalloube et al. 2019). This changes the statistics of the noise for smaller separa- tions and can ultimately prevent the detection of a point-source signal with a reasonable texp. In Fig. 8 we show how the polarimetric contrast evolves at different separations if we combine more and more polarimetric cycles. The points at texp = 0.04 min show the contrast in a sin- gle zero-phase and π-phase combined 2 × 1.2 second exposure just like the bottom frame of Fig. A.2. All other points show the polarimetric contrast in Stokes Q from one single camera after combining the exposures of multiple polarimetric cycles. PCA-ADI was not applied because the procedure requires a cer- tain amount of field rotation to be effective, and therefore would make it difficult to directly compare the results for different total exposure times. The data show that the noise for separations (cid:39) 0.6(cid:48)(cid:48) is pro- portional to t−1/2 exp , just as expected in the photon noise dominated regime, all the way from the shortest to the longest texp, totally in agreement with what we see in the corresponding contrast curve (Fig. 6). This indicates for these separations that longer integra- tions would certainly improve the achievable contrast to a deeper level. For small separations (cid:47) 0.6(cid:48)(cid:48) and short integration times, the contrast first barely improves with increasing texp. Towards longer integration times, however, it also changes to a t−1/2 exp scal- ing. The transition from a flat curve to a square-root scaling happens later for smaller separations. This can be explained be- cause at small separations the noise is dominated by quasi-static aberrations, therefore angular averaging by the field rotation in- creases the SNR of a point-source, however the efficiency of this process depends on the separation and the speed of the field ro- tation. This explanation is supported by Fig. 8 where the transi- tions for the four separations ρ = 0.15(cid:48)(cid:48)...0.45(cid:48)(cid:48) happen when the field rotation leads to an azimuthal shift of about 1.5 − 2.5λ/D Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL (a) (b) Fig. 6. Radial contrast limits as a function of separation for the deepest ZIMPOL high-contrast dataset of α Cen A in the N_R filter. (a) The plot shows 1σ and 5σ limits for the polarized intensity as well as the 1σ photon noise limit for the polarized intensity. The basic data reduction (green line) was done without beam shift correction and residual instrument polarization subtraction, the complete reduction includes both steps. The diamond symbols show the improvement of the contrast limits after applying PCA-ADI. (b) The plot shows 5σ limits for the intensity as well as the 1σ photon noise limit for the intensity. For both contrast limits -- polarized intensity and intensity -- we also show the corresponding contrast of our reference model planet. Fig. 7. Polarized intensity contrast limits for α Cen A, turned into the minimum size of a planet that could be observed at each apparent sep- aration. We assume that planets are at maximum apparent separation (α = 90◦) and we adopt our reference model from Sec. 1.2 for the re- flective properties of the light. Fig. 8. Contrast limits at different texp, calculated for a range of different separations from the star (indicated by different symbols). The dashed lines are proportional to t−1/2 exp , therefore emphasising the expected be- haviour of the noise in the photon noise limited case. at the corresponding separation. This corresponds to about the characteristic size of a speckle. As a reference: In the data used for this study, the speed of the field rotation during the relevant time period is ∼26 mas/10 min or 1.6 λ/D/10 min at the observed wavelength. 5.5. Detection limits The α Cen A/B system is a close binary with semimajor axis of 23.5 AU, which restricts the range of stable planetary orbits around the individual components. Wiegert & Holman (1997) and Quarles & Lissauer (2016) found that orbits around α Cen A are stable for semimajor axes up to ∼3 AU. Stable orbits would preferably be coplanar to the binary orbital plane with inclination i = 79.2◦ but deviations up to ±45◦ are not unlikely from a stabil- ity point of view. There are also reports of other massive planets around one component in close binary systems with a separa- tion smaller than 25 AU (e.g. HD 196885 (Correia et al. 2007), Gliese 86 (Queloz et al. 2000; Lagrange et al. 2006), γ Cep (Hatzes et al. 2003; Neuhäuser et al. 2007) and HD 41004 A (Zucker et al. 2004)). The radial velocity limits for α Cen A (e.g. Zhao et al. 2018) exclude the presence of massive planets with Article number, page 9 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 M sin(i) > 53 MEarth for the classically defined habitable zone from about 1 to 2 AU with even more stringent limits for smaller separations. This evidence is not in favour of a planet around α Cen A with a mass larger or comparable to Jupiter but planets up to almost 100 MEarth cannot be excluded. Depending on the exact composition, formation history and age, gas giants with masses like that could already be close to Jupiter sized (e.g. Swift et al. 2012). With our radius limits in Fig. 7 we show that we might not be far from being able to detect a planet of this size around α Cen A. An important unknown factor in the radius limits are the reflec- tive properties of the planet. For the limits in Fig. 7 we assumed a model with Q(90◦) = 0.055 for the polarized reflectivity of the reflected light. This is optimistic for the reflection of stellar light by the atmosphere of a giant planet, however, some models predict even larger values. The combination of reflection and po- larization could also be larger due to other reasons. Calculations from Arnold & Schneider (2004) have shown that a planet with a Saturn-like unresolved ring could have an exceptionally high brightness in reflected light. It is not unreasonable to assume that α Cen A could har- bour a still undetected companion that could be observed with SPHERE/ZIMPOL in reflected visible light. Our best detection limits based on one single half-night show no evidence for a Jupiter sized planet with exceptionally high fraction of polar- ized reflectivity. However, there is a temporal aspect to the de- tection limits because of the strong dependence of the reflected light intensity and polarization fraction on the phase angle α (see Fig. 2), even a Jupiter sized planet with exceptionally high reflec- tion and polarization would be faint for a large range of phase angles. Therefore, only a series of multiple observations could verify the absence of such a planet. Alternatively, one can carry out a detailed combined analysis of the detection limits and pos- sible companion orbits for an estimate on the likelihood of ob- serving a companion. We did an investigation like this for α Cen A and discuss the procedure and the results in Sec. 6.2. As far as we know, there has not been a direct imaging search comparable to our study for planetary companions in reflected light around α Cen A. Kervella et al. (2006) performed an ex- tensive direct imaging search for faint comoving companions around α Cen A/B with NACO at the VLT in J-,H- and K- band observations. But their results are difficult to compare to ours because the IWA of their contrast limits is larger than the FOV of ZIMPOL. Schroeder et al. (2000) conducted a survey for low mass stellar and sub-stellar companions with the Hub- ble Space Telescope (HST) for some of the brightest stars clos- est to the Sun. Their contrast limits for α Cen A in a range of separations 0.5(cid:48)(cid:48)-1.5(cid:48)(cid:48)are about 7.5-8.5 mag at a wavelength of ∼1.02 µm. Our much deeper contrast limits in intensity are about 13.7-17.4 mag and in polarized intensity 18.3-20.4 mag but with an effective IWA of only ∼0.13(cid:48)(cid:48)for the R-band. 6. Discussion We have shown the exceptional capability of SPHERE/ZIMPOL polarimetry for the search of reflected light from extra-solar planets on our prime target α Cen A in Sec. 5 and our additional targets in Appendix C. The combination of high resolution and polarimetric sensitivity of our observations is far beyond of any other instrument. For α Cen A, B and Altair we derive polarimet- ric contrast limits better than 20 mag at separations >1(cid:48)(cid:48). Even at the effective coronagraphic IWA of 0.13(cid:48)(cid:48)the polarimetric con- trast limits can be around 16 mag. The same performance would also be possible for Sirius A during better observing conditions. Article number, page 10 of 23 A summary of the resulting 5σ contrast limits for all targets can be found in Table 2. For the less bright objects  Eri and τ Ceti we still see polarimetric contrast limits better than 18.9 mag and 18.2 mag at separations >1(cid:48)(cid:48), respectively, with 16 mag close to the effective IWA for τ Ceti. Photon noise limited polarimetric contrasts can be achieved already at separations as small as 0.6(cid:48)(cid:48). 6.1. Comparison to thermal infrared imaging Only a small number of other high-contrast direct imaging searches for planetary companions are published for our targets (Schroeder et al. 2000; Kervella et al. 2006; Thalmann et al. 2011; Vigan et al. 2015; Mizuki et al. 2016; Boehle et al. 2019; Mawet et al. 2019). The observations were typically carried out with available near-IR high-contrast imagers and the aim was usually a search for thermal light from brown dwarfs or very massive, self-luminous planets. The detection of such objects around the nearest stars would have been possible, but is quite unexpected. For these near-IR observations, the expected signal for the reflected light from a planet is far out of reach, but the obtained results represent the best limits achieved so far. The most sensitive limits were obtained with a combination of SDI and ADI for Sirius A with the SPHERE/IRDIFS mode (Vigan et al. 2015). Our observation of this object suffered from bad ob- serving conditions, however, for PDI and ADI observations of similarly bright targets, our reported contrast limits show an im- provement of 2 -- 3 mag at all separations up to 1.7(cid:48)(cid:48). However, much improved sensitivity is severely needed to detect a planet in reflected light. The only targets where a detection seems to be possible in a single night are α Cen A and B. The lower bright- ness of the other targets decreases the sensitivity at a given an- gular separation and the larger distance to them increases the contrast of companions for the same angular separations. The physical meaning of the contrast limits for the reflected light is different compared to the limits from IR-surveys for the thermal emission from the planet. The contrast limits in the in- frared probe the intrinsic luminosity and surface temperature and can be transformed into upper limits for the planet mass with models for planet formation and evolution (e.g. Baraffe et al. 2003; Spiegel & Burrows 2012) if the age of the system is known and if the irradiation from the star can be neglected. Evolved planets are usually close to or at equilibrium tempera- ture and emit for separations of ∼1 AU or larger at longer wave- lengths (∼10 µm) where reaching high-contrast is difficult with current ground-based observations. The intrinsic flux of planets drops off exponentially towards visible wavelengths. For exam- ple, assuming perfect black body spectra and a solar-like host star, even a self-luminous 800 K Jupiter-sized planet would only have a contrast of order 3 · 10−11 in the visual I-band. While the contrast of the reflected light would be around 3 · 10−8 for dp < 1 AU and I(90◦) = 0.131 (for a discussion with wavelength dependent reflectivity see Sudarsky et al. (2003)). Planets with dp ≈ 1 AU around α Cen A/B would have to be at temperatures above ∼1000 K to be brighter in thermal emission compared to reflected light at visible wavelengths. This is the reason why we can only probe reflected stellar light in the visible wavelengths for all our targets and we do not expect any contribution from thermal emission. The recently launched NEAR (Kasper et al. 2017; Käufl et al. 2018) survey using the VISIR instrument at the VLT aims to achieve high contrasts at 10 µm for α Cen A/B and possibly de- tect evolved planets in the habitable zone of this binary system. The results of the NEAR campaign will be especially interesting for our survey since it will at least provide exceptionally deep de- Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL Table 2. Summary of 5σ contrast limits for the intensity Cflux and polarized intensity Cpol at some key separations for each target Object mR Filters a texp Sirius A Altair  Eri α Cen A α Cen B τ Ceti -1.5 0.6 3.0 -0.5 1.0 2.9 N_I 2h 55.2min R_PRIM 2h 31.2min VBB N_R VBB R_PRIM 2h 48min 3h 12min 3h 21.6min 3h 26.8min Cpol (mag) Inside AO contr. rad.b (cid:47) 0.35(cid:48)(cid:48) (cid:47) 0.45(cid:48)(cid:48) 15.0 16.8 15.8 16.8 17.1 15.8 0.5(cid:48)(cid:48) 15.8 17.9 16.2 18.3 18.5 16.7 1.5(cid:48)(cid:48) 18.8 20.4 19.6 20.4 20.4 18.8 Cflux (mag) Inside AO contr. rad.b (cid:47) 0.35(cid:48)(cid:48) (cid:47) 0.45(cid:48)(cid:48) 11.0 12.7 10.4 12.8 13.0 12.4 0.5(cid:48)(cid:48) 11.8 14.6 11.3 14.0 14.1 13.7 1.5(cid:48)(cid:48) 15.7 19.5 16.3 18.9 18.5 18.2 Notes. (a) The combined total exposure time (b) Average value for the contrast limit at separations inside the AO control radius ((cid:47) 20λ/D) tection limits in the IR that can be directly compared to our own limits for α Cen A/B in reflected light at visible wavelengths. 6.2. Interpreting the contrast limits In contrast to the thermal emission, the polarized intensity of reflected stellar light depends strongly on the planet radius RP, the planet-star separation dp, the reflective properties of the at- mosphere and the phase of the planet (see Eq. (2)). Therefore, contrast limits yield -- for a given physical separation, orbital phase and reflective properties -- an upper limit for the planet radius. This means that the upper limits for the size of a com- panion as presented for the α Cen A data come with a set of critical assumptions. The contrast limits are determined for the apparent separation between star and planet ρ. The reflected light brightness of the planet, however, depends on the physical sep- aration dp and planets located at apparent separation ρ can have any physical separation dp ≥ ρ. This introduces a degeneracy into the calculation of physical parameters that cannot be lifted without further assumptions. Because of this, we assumed for the radius upper limits in Fig. 7 that the physical separation corre- sponds to the apparent separation, in addition to fixing the scat- tering model. This assumption can be justified for a blind search for planets with a Monte-Carlo (MC) simulation of apparent sep- arations and contrasts for a random sample of planets. We sim- ulated 5 000 000 Jupiter sized planets on circular orbits around α Cen A with randomly distributed semi-major axes and incli- nations and the Rayleigh scattering atmosphere model discussed in Sec. 1.2. We used a flat prior distribution for the orbital phase angles in the interval [0, 2π] and for the semi-major axes in the interval [0.01, 3] AU. The inner boundary for the semi-major axis has a negligible effect on the final result as long as it is smaller than 0.18 AU (the effective IWA of our data). Planets with larger semi-major axes would be unstable due to the close binary. For the inclination we assumed a Gaussian prior with a standard deviation of 45◦, centred on the inclination of the binary orbit. Large mutual inclinations of binary and planetary orbits are unlikely due to stability reasons (Quarles & Lissauer 2016). We chose α Cen A as our example because it has some of the best detection limits. Panel (a) in Fig. 9 shows the likelihood of one of the simulated planets having a certain apparent separation and contrast. The likelihood was calculated by dividing the num- ber of MC-samples in each contrast-separation bin by the total number of sampled planets. The likelihood drops to zero towards the upper right corner because planets at large separations have an upper limit for their reflected light intensity determined by their size and reflective properties. The dividing line with the strongly increased likelihood in the center is ∝ ρ−2, representing planets at maximum elongation, corresponding to orbital phase angles close to 90◦ and 270◦. It is more likely for a planet to be located around this line independent from the inclination of its orbit. For orbits close to edge-on the apparent movement of the planet is slower at these phase angles, this naturally increases the likelihood of it being observed during this phase. For or- bits closer to face-on the apparent separation of the planet will not change much during the orbit, this also increases the likeli- hood of the planet being observed during maximum elongation. Around 66% of all sampled planets end up inside the parame- ter space shown in Fig. 9 and 24% end up in close proximity (∆m ≈ 0.4 mag) to the line with maximum separation and con- trast. This is a large fraction considering that we did not assume any prior knowledge about the orbital phase of the sample plan- ets. We compare the likelihood to the completeness or the per- formance map (see Jensen-Clem et al. 2018) of our observation in panel (b) of Fig. 9, adopting the previously shown contrast curve for α Cen A (Fig. 6(b)) and assuming a Gaussian noise distribution. The full performance map in panel (b) is drawn for a detection threshold τ = 5σ, additionally we show the 50% completeness contour for τ = 3σ. The completeness can be un- derstood as the fraction of true positives given τ = 3σ or 5σ. We multiply the performance map and likelihood in panels (c) and (d) in Fig. 9 to show the expected fraction of detectable planets for both detection thresholds and calculate the total integrated fraction of observed planets. Only about 1.5% of the samples would produce a signal with SNR=5 in our data but the num- ber increases by almost a factor of 10 to about 13.5% for signals with SNR=3. This happens because the shape of the contrast curve resembles the ∝ ρ−2 shape of the parameter space where the likelihood is strongly increased. If both curves are on a simi- lar level in terms of contrast, just like in our case with α Cen A, a small contrast improvement can considerably increase the possi- bility of a detection. The same happens if we lower the detection threshold but this simultaneously increases the probability for a false detection (false-alarm probability) significantly. For Gaus- sian distributed noise and a 1024×1024 px2 detector the expected number of random events exceeding >5σ is smaller than one, but the number increases to ∼ 1000 for >3σ. Therefore, the 5σ threshold should definitely be respected in a blind search. How- ever, a detection between 3-5σ could be enough if there were multiple independent such detections with ZIMPOL that could be combined into one single, more significant detection. The MC-simulation shows that the reflected light from a Jupiter sized planet around α Cen A could be detected as a 5σ signal in a single half-night, when it is located relatively close (0.13(cid:48)(cid:48)-0.3(cid:48)(cid:48)) to the star. It should also be possible to detect Jupiter sized planets at any other separation as 3σ signals and multiple 3σ detections could be combined to a 5σ result. Alter- natively, a less significant detection could be considered suffi- cient if the position of the planet is known from another high- contrast detection or from the astrometric reflex motion of the star. Article number, page 11 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Fig. 9. The results of the Monte-Carlo sampling of 5 000 000 Jupiter sized planets on circular orbits around α Cen A. (a) The fraction of samples (in percent) that end up at a certain apparent separation and polarized intensity contrast at all times. (b) The full performance map for a certain detection threshold Nτ. It shows the probability to detect an existing planet for all separations and contrasts with an SNR of N. (c) and (d) The combination of the likelihood and the performance map gives the fraction of planets at each separation and contrast that are detected as either SNR=5 or SNR=3 signals depending on the detection threshold. 6.3. Improving the contrast limits There are multiple ways to improve the detection limits with ZIMPOL with future observations. Different strategies are re- quired for blind searches when compared to follow up observa- tions of already known planets. For blind searches the most ef- fective way is to just increase the total integration time. We have shown in Sec. 5.4 that the contrast improves with the square-root of the integration time. The observations should be done in P1 polarimetry mode to enable ADI for improving the contrast at smaller separations. For longer total integration times it will be necessary to combine the data from multiple observing nights. This is not straight forward for our targets because the apparent orbital motion is large. The most extreme case is α Cen A for which a planet on a face-on circular orbit would move 40 mas or ∼ 2λ/D per day at the IWA of 0.13(cid:48)(cid:48)and 10 mas or ∼ 0.5λ/D at 1.7(cid:48)(cid:48). For the combination of data from different, even con- secutive nights it will be necessary to consider the Keplerian motion of planets. This is possible with data analysis tools like K-Stacker (Nowak et al. 2018). K-Stacker was developed espe- cially for finding weak planet signals in a time series of images when they move on Keplerian orbits. For a time series spanning weeks it would also be necessary to additionally consider the change of the reflected polarized intensity as function of the or- bital phase (Fig. 2). The orbital motion of planets around nearby stars could also be used as an advantage to further improve the contrast limits. Males et al. (2015) developed the concept of Or- bital Differential Imaging (ODI) that exploits the orbital motion of a planet in multi-epoch data to remove the stellar PSF, while minimizing the subtraction of the planet signal. Article number, page 12 of 23 Follow-up observations of a known planet would have major advantages over a blind search because the prior knowledge of orbital phase or orbit location from RV or astrometric measure- ments can be exploited for optimizing the observing strategy and simplify the analysis of the data. Currently, the best planets for a successful follow up with ZIMPOL are the giant planets  Eri b and GJ 876 b and the terrestrial planet Proxima Centauri b. The planet around  Eri can be observed at the favourable photon noise limited apparent separation of ∼0.8(cid:48)(cid:48)with ZIMPOL. How- ever, it is expected to be rather faint in reflected light because of its large semi-major axis of ∼3 AU. The polarimetric contrasts of both Proxima Centauri b and GJ 876 b are expected to be less de- manding but the expected maximum separation of only ∼0.04(cid:48)(cid:48), corresponding to ∼2λ/D in the visible, is very challenging. This requires a specialized instrumental setup for SPHERE/ZIMPOL for example an optimized pupil mask developed to suppress the first Airy ring at 2λ/D as proposed by Patapis et al. (2018). For companions with known separation the selection of the ZIMPOL instrument mode can also be optimized. The P1 po- larimetry mode should be used for companions close to or inside the AO control ring <0.7(cid:48)(cid:48)because it allows the use of ADI for additional speckle noise suppression. ADI also helps to reduce static noise induced by the instrument itself. However, for larger separations ADI is not necessary and the field stabilized P2 po- larimetry mode could be used. This would allow to use longer DIT without diluting the planet signal due to the field rotation during exposures. For planets with well known orbital parameters like semi- major axis, inclination and orbital phase, it would be possible to plan observations to be executed at the right time when the Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL √ reflected intensity and apparent separation are optimal. And fi- nally, if also the position angle of the orbit is known, it would be possible to align the polarimetric Q-direction of ZIMPOL with the expected orientation of the polarized signal from the planet. This would allow to only observe in a rotated Q polarization co- ordinate system without spending half of the time observing U, which is expected to be zero. The observation time would be cut in half for the same detection sensitivity or the contrast limit would be improved by a factor of 2 in the same amount of telescope time. There are certainly other ways to improve the detection limits which were not sufficiently investigated yet. The use of narrow-band versus broad-band filters could be beneficial be- cause instrumental effects like beam shift and instrumental polar- ization are wavelength dependent and the post-processing cannot fully account for this. Therefore, the applied corrections are not optimal for observations taken with broad-band filters and would provide better results for narrow-band filters. Another way to im- prove the detection limits is frame selection. The gain both of the mentioned techniques is difficult to quantify because we did not find any point-sources in our data. Adding more data, even data of bad quality, generally improved the calculated detection lim- its because it decreased the noise level of the data. However, data with bad quality also lowers the signal of a point-source but this effect can only be studied properly if a real signal is present in the data because for deep coronagraphic observations we do not know the exact PSF shape for each image. 7. Conclusion We have observed α Cen A and B, Sirius A,  Eri and τ Cet using SPHERE/ZIMPOL in polarimetry mode. The target list for the search of reflected light from extra-solar planets with di- rect imaging is short and the targets were selected for achiev- ing deep detection limits within a few hours of observation. We were not able to detect a polarized intensity signal above the detection threshold from any of our targets, however, our data provide some of the deepest contrast limits for direct imaging to date. The achieved limits for our brightest targets show that the detection of polarized reflected light from a 1 RJ sized object would be possible in a single night under good observing con- ditions (Seeing (cid:47) 0.8(cid:48)(cid:48), τ0 (cid:39) 4 ms) for our nearest neighbours α Cen A/B with a realistic model for a reflecting atmosphere. Unfortunately, our null result is not constraining for the occur- rence rate of giant planets because of the strong time dependence of the reflected light intensity and given the low frequency of gas giants with 1-10 Jupiter masses between 0.3-3 AU is expected to be only about 4% (Cumming et al. 2008; Fernandes et al. 2019), slightly higher for A-stars (Johnson et al. 2010), but lower for intermediate separation binaries (Kraus et al. 2016). Our results show the capability of ZIMPOL to remove the unpolarized stellar PSF and they deliver the deepest contrast lim- its for direct imaging at visible wavelengths from 600-900 nm. The performance is close to the photon noise limit and this al- lows to scale the contrast limits for different total integration times and for targets with different brightnesses. This will be useful in the future for planning further observations in particu- lar for larger programs with deeper observations of the surround- ings of the nearest stars by combining the results of many nights. Due to the strong phase dependence the search for reflected light is especially well suited as potential follow up observation of targets with known orbital phases, already determined with dif- ferent methods (e.g. RV, astrometry). Another use of the highly sensitive polarimetry with ZIMPOL could be the determination of the linear polarization of the thermal light of low mass com- panions. This measurement has been tried before for a few differ- ent targets at infrared wavelengths (e.g. Jensen-Clem et al. 2016; van Holstein et al. 2017). The main difficulty with brown dwarf companions is that the linear polarization degree for the thermal light is expected to be <1% (Stolker et al. 2017). Another prob- lem for ZIMPOL polarimetry is the low luminosity of L and T dwarfs in the visible wavelengths. Another important aspect of this work are our investigations on the limitations of SPHERE/ZIMPOL at the VLT. We have in- vestigated and corrected the residual instrument polarization and most importantly the polarimetric beam shift effect. The beam shift effect is well known in optics but ZIMPOL is the first as- tronomical instrument where this effect is apparent in the data because of its high spatial resolution and polarimetric sensitiv- ity. Despite all the calibrations and corrections applied in this work, there remain substantial speckle residuals in the differen- tial polarimetry which currently limit the contrast performance at small separations (<0.6(cid:48)(cid:48)). However, the achieved polarimet- ric contrast Cpol is in most cases more than 10 times deeper for separations <1(cid:48)(cid:48)and up to ∼50 times deeper at the smallest sep- arations compared to the imaging contrast Cflux achieved with classical PCA-ADI processing. Over a larger range of separa- tions, deep polarimetric contrast limits and even photon noise limited performance is achieved without additional PCA-ADI. Therefore, polarimetry is a most attractive method to push the detection limits for reflecting planets with future high-contrast instruments (Kasper et al. 2010). In particular, the speckle sup- pression can be further improved by better avoiding color depen- dent disturbing effects (see Appendix A) or by taking advantage of the much improved light gathering power of the upcoming generation of 30-40 meter ELTs. With larger telescopes it will be possible to achieve the same contrast limits for fainter stars at higher resolution, significantly increasing the sample size of nearby targets and possibly allow imaging the planets in the hab- itable zone of nearby M dwarfs like Proxima Centauri. The ex- perience gained with the SPHERE/ZIMPOL RefPlanets survey described in this work should therefore be helpful for the trade- off studies and description of the design of such instruments. Acknowledgements. SH and HMS acknowledge the financial support by the Swiss National Science Foundation through grant 200020_162630/1. SPHERE is an instrument designed and built by a consortium consisting of IPAG (Greno- ble, France), MPIA (Heidelberg, Germany), LAM (Marseille, France), LESIA (Paris, France), Laboratoire Lagrange (Nice, France), INAF -- Osservatorio di Padova (Italy), Observatoire de Genève (Switzerland), ETH Zurich (Switzer- land), NOVA (Netherlands), ONERA (France) and ASTRON (Netherlands) in collaboration with ESO. SPHERE was funded by ESO, with additional contri- butions from CNRS (France), MPIA (Germany), INAF (Italy), FINES (Switzer- land) and NOVA (Netherlands). SPHERE also received funding from the Eu- ropean Commission Sixth and Seventh Framework Programmes as part of the Optical Infrared Coordination Network for Astronomy (OPTICON) under grant number RII3-Ct-2004-001566 for FP6 (2004 -- 2008), grant number 226604 for FP7 (2009 -- 2012) and grant number 312430 for FP7 (2013 -- 2016). We also acknowledge financial support from the Programme National de Planétologie (PNP) and the Programme National de Physique Stellaire (PNPS) of CNRS- INSU in France. This work has also been supported by a grant from the French Labex OSUG@2020 (Investissements d'avenir -- ANR10 LABX56). The project is supported by CNRS, by the Agence Nationale de la Recherche (ANR-14- CE33-0018). It has also been carried out within the frame of the National Centre for Competence in Research PlanetS supported by the Swiss National Science Foundation (SNSF). MRM and SPQ are pleased to acknowledge this financial support of the SNSF. Finally, this work has made use of the the SPHERE Data Centre, jointly operated by OSUG/IPAG (Grenoble), PYTHEAS/LAM/CESAM (Marseille), OCA/Lagrange (Nice), Observatoire de Paris/LESIA (Paris), and Observatoire de Lyon, also supported by a grant from Labex OSUG@2020 (In- vestissements d'avenir -- ANR10 LABX56). We thank P. Delorme and E. La- gadec (SPHERE Data Centre) for their efficient help during the data reduction process. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. This work has been partially supported by the project PRIN- Article number, page 13 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 INAF 2016 The Cradle of Life- GENESIS-SKA (General Conditions in Early Planetary Systems for the rise of life with SKA). We also acknowledge support from INAF/Frontiera (Fostering high ResolutiON Technology and Innovation for Exoplanets and Research in Astrophysics) through the "Progetti Premiali" fund- ing scheme of the Italian Ministry of Education, University, and Research. T.H. acknowledges support from the European Research Council under the Horizon 2020 Framework Program via the ERC Advanced Grant Origins 83 24 28. References Amara, A. & Quanz, S. P. 2012, MNRAS, 427, 948 Arnold, L. & Schneider, J. 2004, A&A, 420, 1153 Backman, D., Marengo, M., Stapelfeldt, K., et al. 2009, ApJ, 690, 1522 Bailey, J., Kedziora-Chudczer, L., & Bott, K. 2018, MNRAS, 480, 1613 Bailey, J., Lucas, P. W., & Hough, J. H. 2010, MNRAS, 405, 2570 Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, A&A, 402, 701 Bazzon, A., Gisler, D., Roelfsema, R., et al. 2012, in Proc. SPIE, Vol. 8446, Ground-based and Airborne Instrumentation for Astronomy IV, 844693 Bazzon, A., Schmid, H. M., & Buenzli, E. 2014, A&A, 572, A6 Bazzon, A., Schmid, H. M., & Gisler, D. 2013, A&A, 556, A117 Beuzit, J.-L., Feldt, M., Dohlen, K., et al. 2008, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 7014, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 18 Beuzit, J. L., Vigan, A., Mouillet, D., et al. 2019, arXiv e-prints, arXiv:1902.04080 Boehle, A., Quanz, S. P., Lovis, C., et al. 2019, arXiv e-prints, arXiv:1907.04334 Bond, H. E., Schaefer, G. H., Gilliland, R. L., et al. 2017, ApJ, 840, 70 Booth, M., Dent, W. R. F., Jordán, A., et al. 2017, MNRAS, 469, 3200 Bottom, M., Ruane, G., & Mawet, D. 2017, Research Notes of the American Astronomical Society, 1, 30 Bowler, B. P. 2016, PASP, 128, 102001 Buenzli, E. & Schmid, H. M. 2009, A&A, 504, 259 Cantalloube, F., Dohlen, K., Milli, J., Brandner, W., & Vigan, A. 2019, The Mes- senger, 176, 25 Chauvin, G., Desidera, S., Lagrange, A. M., et al. 2017, A&A, 605, L9 Claudi, R. U., Turatto, M., Gratton, R. G., et al. 2008, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 7014, Ground- based and Airborne Instrumentation for Astronomy II, 70143E Correia, A. C. M., Udry, S., Mayor, M., et al. 2007, arXiv e-prints [arXiv:0711.3343] Cotton, D. V., Bailey, J., Kedziora-Chudczer, L., et al. 2016, MNRAS, 455, 1607 Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP, 120, 531 Dohlen, K., Langlois, M., Saisse, M., et al. 2008, in Society of Photo-Optical In- strumentation Engineers (SPIE) Conference Series, Vol. 7014, Ground-based and Airborne Instrumentation for Astronomy II, 70143L Dollfus, A. 1985, Advances in Space Research, 5, 47 Feng, F., Tuomi, M., Jones, H. R. A., et al. 2017, AJ, 154, 135 Fernandes, R. B., Mulders, G. D., Pascucci, I., Mordasini, C., & Emsenhuber, A. 2019, ApJ, 874, 81 Fusco, T., Petit, C., Rousset, G., et al. 2006, in Society of Photo-Optical Instru- mentation Engineers (SPIE) Conference Series, Vol. 6272, 62720K Gisler, D., Schmid, H. M., Thalmann, C., et al. 2004, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 5492, Ground- based Instrumentation for Astronomy, ed. A. F. M. Moorwood & M. Iye, 463 -- 474 Greaves, J. S., Sibthorpe, B., Acke, B., et al. 2014, ApJ, 791, L11 Greaves, J. S., Wyatt, M. C., Holland, W. S., & Dent, W. R. F. 2004, MNRAS, 351, L54 905 Hansen, J. E. & Hovenier, J. W. 1974, Journal of Atmospheric Sciences, 31, 1137 Hatzes, A. P., Cochran, W. D., Endl, M., et al. 2003, ApJ, 599, 1383 Hatzes, A. P., Cochran, W. D., McArthur, B., et al. 2000, ApJ, 544, L145 Holman, M. J. & Wiegert, P. A. 1999, AJ, 117, 621 Jensen-Clem, R., Mawet, D., Gomez Gonzalez, C. A., et al. 2018, AJ, 155, 19 Jensen-Clem, R., Millar-Blanchaer, M., Mawet, D., et al. 2016, ApJ, 820, 111 Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R. 2010, PASP, 122, Jovanovic, N., Martinache, F., Guyon, O., et al. 2015, PASP, 127, 890 Kasper, M., Arsenault, R., Käufl, H. U., et al. 2017, The Messenger, 169, 16 Kasper, M., Beuzit, J.-L., Verinaud, C., et al. 2010, in Proc. SPIE, Vol. 7735, Ground-based and Airborne Instrumentation for Astronomy III, 77352E -- 77352E -- 9 Käufl, H.-U., Kasper, M., Arsenault, R., et al. 2018, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 10702, Proc. SPIE, 107020D Article number, page 14 of 23 270 669 152, 8 459, 955 495, 335 Keller, C. U., Schmid, H. M., Venema, L. B., et al. 2010, in Proc. SPIE, Vol. 7735, Ground-based and Airborne Instrumentation for Astronomy III, 77356G Kemp, J. C., Henson, G. D., Steiner, C. T., & Powell, E. R. 1987, Nature, 326, Keppler, M., Benisty, M., Müller, A., et al. 2018, A&A, 617, A44 Kervella, P., Arenou, F., Mignard, F., & Thévenin, F. 2019, A&A, 623, A72 Kervella, P., Mignard, F., Mérand, A., & Thévenin, F. 2016, A&A, 594, A107 Kervella, P., Thévenin, F., Coudé du Foresto, V., & Mignard, F. 2006, A&A, 459, Kraus, A. L., Ireland, M. J., Huber, D., Mann, A. W., & Dupuy, T. J. 2016, AJ, Lagrange, A. M., Beust, H., Udry, S., Chauvin, G., & Mayor, M. 2006, A&A, Lagrange, A.-M., Desort, M., Galland, F., Udry, S., & Mayor, M. 2009, A&A, Lawler, S. M., Di Francesco, J., Kennedy, G. M., et al. 2014, MNRAS, 444, 2665 Macintosh, B., Graham, J. R., Barman, T., et al. 2015, Science, 350, 64 Macintosh, B., Graham, J. R., Ingraham, P., et al. 2014, Proceedings of the Na- tional Academy of Science, 111, 12661 Males, J. R., Belikov, R., & Bendek, E. 2015, in Society of Photo-Optical In- strumentation Engineers (SPIE) Conference Series, Vol. 9605, Proc. SPIE, 960518 Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322, 1348 Mawet, D., Hirsch, L., Lee, E. J., et al. 2019, AJ, 157, 33 Mawet, D., Milli, J., Wahhaj, Z., et al. 2014, ApJ, 792, 97 McLean, W., Stam, D. M., Bagnulo, S., et al. 2017, A&A, 601, A142 Milli, J., Mouillet, D., Mawet, D., et al. 2013, A&A, 556, A64 Mizuki, T., Yamada, T., Carson, J. C., et al. 2016, A&A, 595, A79 Monnier, J. D., Zhao, M., Pedretti, E., et al. 2007, Science, 317, 342 Neuhäuser, R., Mugrauer, M., Fukagawa, M., Torres, G., & Schmidt, T. 2007, A&A, 462, 777 Nowak, M., Le Coroller, H., Arnold, L., et al. 2018, A&A, 615, A144 Patapis, P., Kühn, J., & Schmid, H. M. 2018, in Society of Photo-Optical In- strumentation Engineers (SPIE) Conference Series, Vol. 10706, Advances in Optical and Mechanical Technologies for Telescopes and Instrumentation III, 107065J Pavlov, A., Möller-Nilsson, O., Feldt, M., et al. 2008, in Proc. SPIE, Vol. 7019, Advanced Software and Control for Astronomy II, 701939 Quarles, B. & Lissauer, J. J. 2016, AJ, 151, 111 Queloz, D., Mayor, M., Weber, L., et al. 2000, A&A, 354, 99 Schmid, H. M., Bazzon, A., Roelfsema, R., et al. 2018, A&A, 619, A9 Schmid, H. M., Beuzit, J.-L., Feldt, M., et al. 2006a, in IAU Colloq. 200: Direct Imaging of Exoplanets: Science & Techniques, ed. C. Aime & F. Vakili, 165 -- 170 Schmid, H.-M., Downing, M., Roelfsema, R., et al. 2012, in Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8446, So- ciety of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 8 Schmid, H. M., Joos, F., Buenzli, E., & Gisler, D. 2011, Icarus, 212, 701 Schmid, H. M., Joos, F., & Tschan, D. 2006b, A&A, 452, 657 Schmidt, T. O. B., Neuhäuser, R., Briceño, C., et al. 2016, A&A, 593, A75 Schroeder, D. J., Golimowski, D. A., Brukardt, R. A., et al. 2000, AJ, 119, 906 Schworer, G. & Tuthill, P. G. 2015, A&A, 578, A59 Seager, S., Whitney, B. A., & Sasselov, D. D. 2000, ApJ, 540, 504 Smith, P. H. & Tomasko, M. G. 1984, Icarus, 58, 35 Spiegel, D. S. & Burrows, A. 2012, ApJ, 745, 174 Stam, D. M. 2008, A&A, 482, 989 Stam, D. M., Hovenier, J. W., & Waters, L. B. F. M. 2004, A&A, 428, 663 Stolker, T., Bonse, M. J., Quanz, S. P., et al. 2019, A&A, 621, A59 Stolker, T., Min, M., Stam, D. M., et al. 2017, A&A, 607, A42 Sudarsky, D., Burrows, A., & Hubeny, I. 2003, ApJ, 588, 1121 Swift, D. C., Eggert, J. H., Hicks, D. G., et al. 2012, ApJ, 744, 59 Thalmann, C., Janson, M., Buenzli, E., et al. 2011, ApJ, 743, L6 Thalmann, C., Schmid, H. M., Boccaletti, A., et al. 2008, in Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 7014, So- ciety of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 3 Tomasko, M. G. & Smith, P. H. 1982, Icarus, 51, 65 Tuomi, M., Jones, H. R. A., Jenkins, J. S., et al. 2013, A&A, 551, A79 van Dam, M. A., Le Mignant, D., & Macintosh, B. A. 2004, Appl. Opt., 43, 5458 van Holstein, R. G., Snik, F., Girard, J. H., et al. 2017, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 10400, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 1040015 Vigan, A., Gry, C., Salter, G., et al. 2015, MNRAS, 454, 129 Wiegert, P. A. & Holman, M. J. 1997, AJ, 113, 1445 Zhao, L., Fischer, D. A., Brewer, J., Giguere, M., & Rojas-Ayala, B. 2018, AJ, Zucker, S., Mazeh, T., Santos, N. C., Udry, S., & Mayor, M. 2004, A&A, 426, 155, 24 695 Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL 1 ETH Zurich, Institute for Particle Physics and Astrophysics, Wolfgang-Pauli-Strasse 27, CH-8093 Zurich, Switzerland 2 NOVA Optical Infrared Instrumentation Group at ASTRON, Oude Hoogeveensedijk 4, 7991 PD Dwingeloo, The Netherlands 3 Université Grenoble Alpes, IPAG, 38000 Grenoble, France 4 CNRS, IPAG, 38000 Grenoble, France 5 European Southern Observatory, Alonso de Cordova 3107, Casilla 19001 Vitacura, Santiago 19, Chile 6 Institute for Computational Science, University of Zürich, Win- terthurerstrasse 190, 8057 Zürich, Switzerland 7 Istituto Ricerche Solari Locarno, Via Patocchi 57, 6605 Locarno 8 Kiepenheuer-Institut für Sonnenphysik, Schneckstr. 6, D-79104 Monti, Switzerland Freiburg, Germany 9 Anton Pannekoek Institute for Astronomy, University of Amster- dam, PO Box 94249, 1090 GE Amsterdam, The Netherlands 10 LESIA, CNRS, Observatoire de Paris, Université Paris Diderot, UPMC, 5 place J. Janssen, 92190 Meudon, France 11 Leiden Observatory, Leiden University, P.O. Box 9513, 2300 RA Leiden, The Netherlands 12 Laboratoire Lagrange, UMR7293, Université de Nice Sophia- Antipolis, CNRS, Observatoire de la Côte d'Azur, Boulevard de l'Observatoire, 06304 Nice, Cedex 4, France 13 INAF - Osservatorio Astronomico di Roma, via Frascati 33, I-00087 14 Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidel- Monte Porzio Catone, Italy berg, Germany 15 INAF -- Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 5, 35122 Padova, Italy 16 Aix Marseille Université, CNRS, CNES, LAM (Laboratoire d'Astrophysique de Marseille) UMR 7326, 13388, Marseille, France 17 Unidad Mixta International Franco-Chilena de Astronomia, CNRS/INSU UMI 3386 and Departemento de Astronomia, Univer- sidad de Chile, Casilla 36-D, Santiago, Chile 18 Nucleo de Astronomia, Facultad de Ingenieria y Ciencias, Universi- dad Diego Portales, Av. Ejercito 441, Santiago, Chile 19 Escuela de Ingenieria Industrial, Facultad de Ingenieria y Ciencias, Universidad Diego Portales, Av. Ejercito 441, Santiago, Chile 20 European Southern Observatory, Karl Schwarzschild St, 2, 85748 Garching, Germany 21 ONERA, The French Aerospace Lab BP72, 29 avenue de la Division Leclerc, 92322 Châtillon Cedex, France 22 Centre de Recherche Astrophysique de Lyon, CNRS/ENSL Univer- sité Lyon 1, 9 av. Ch. André, 69561 Saint-Genis-Laval, France 23 INAF - Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, I- 50125 Firenze, Italy 24 Geneva Observatory, University of Geneva, Chemin des Mailettes 51, 1290 Versoix, Switzerland 25 STAR Institute, Université de Liège, Allée du Six Août 19c, 4000 26 Department of Astronomy, University of Michigan, Ann Arbor, MI Liège, Belgium 48109, USA Article number, page 15 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Appendix A: Advanced data reduction steps Appendix A.1: Frame transfer smearing correction ZIMPOL uses frame transfer CCDs which shift each frame at the end of an illumination from the image area to a covered read-out area of the detector. There the previously illuminated frame is read-out during integration of the next frame. The detector is also illuminated during the fast frame transfer, which lasts 56 ms for fast polarimetry, and this causes a frame transfer smearing of the frame in the column direction. The smearing amounts to a maxi- mum of 5 % for the shortest integration time of 1.1 s and less for longer integrations (see Schmid et al. 2018, 2012). The smearing corrects itself in the polarization signal QZ but is for short expo- sures apparent in the intensity signal IZ. We correct this frame transfer smearing in each intensity image with the subtraction of a correctly scaled mean row profile from every row in the bias subtracted image. = p2 tel,Q(θpar) + p2 Appendix A.2: Telescope polarization correction The SPHERE/ZIMPOL instrument uses a half wave plate (HWP) polarization switch to select opposite polarization modes Q+ and Q−, or U+ and U− to compensate the instrumental po- larization (Bazzon et al. 2012). However, there is a remaining residual telescope polarization ptel from the optical components located in front of the HWP switch. The magnitude of the resid- ual telescope polarization p2 tel,U(θpar) is wave- tel length dependent but otherwise essentially constant. The orien- tation of ptel rotates in the Stokes Q-U plane as a function of the parallactic angle of the telescope θpar. The measured Stokes parameters Qobs and Uobs are therefore a combination of the as- trophysical polarization of the target Q, U and the induced tele- scope polarization: Qobs(x, y) = Q(x, y) + ptel,Q(θpar) · I(x, y) Uobs(x, y) = U(x, y) + ptel,U(θpar) · I(x, y) To first order, the telescope polarization does not depend on the image coordinates (x, y). Therefore, we can estimate the fractional telescope polarization for unpolarized sky sources (Q(x, y) ≈ 0 and U(x, y) ≈ 0) by measuring the mean2 fractional polarization Qobs/I and Uobs/I: ptel,Q ≈ (cid:104)Qobs(x, y)(cid:105)/(cid:104)I(x, y)(cid:105) ptel,U ≈ (cid:104)Uobs(x, y)(cid:105)/(cid:104)I(x, y)(cid:105) Schmid et al. (2018) used measurements of unpolarized standard stars for the determination of the amplitude ptel in several differ- ent filters for SPHERE/ZIMPOL, which can be described by ptel,Q(θpar) = ptel · cos ptel,U(θpar) = ptel · sin where δtel is a wavelength dependent offset angle for the tele- scope polarization. (cid:17)(cid:17) (cid:17)(cid:17) θpar + δtel θpar + δtel (A.1) (cid:16) (cid:16) (cid:16) (cid:16) 2 2 (A.2) (A.3) , , 2 Average over pixels in a ring around the star with inner radius r larger than the coronagraph and outer radius R: (cid:80) (cid:80) f (x, y) 1 r2<x2+y2<R2 (cid:104) f (x, y)(cid:105) = r2<x2+y2<R2 Article number, page 16 of 23 Our targets are not zero-polarization standard stars, but it is expected that regular main-sequence stars are highly spherically symmetric and are hot enough to not have any clouds/hazes, therefore α Cen A and B, Sirius A,  Eri or τ Cet are expected to show only very little intrinsic integrated polarization in broad- band filters. For example for the Sun, an upper limit on the inte- grated linear polarization of < 10−6 was determined by Kemp et al. (1987). Altair could be an exception and it may show larger intrinsic linear polarization, because it is a rapidly rotating and therefore ellipsoidal A-star (Monnier et al. 2007). However, existing polarimetry of Altair yield a very low polarization of < 10−5 (Bailey et al. 2010). No interstellar polarization compo- nent is expected for our stars, because of the distance of only a few parsecs, as is confirmed by high precision polarimetry (Bai- ley et al. 2010; Cotton et al. 2016). Because our targets were observed over a large range of par- allactic angles, they show the steady rotation of ptel as illus- trated in Fig. A.1 for the narrow R-band observations of α Cen A taken in April 30, 2017. Each point in the plot shows the aver- age telescope polarization measured in an annulus -- extending from 0.13(cid:48)(cid:48) to 0.72(cid:48)(cid:48) -- in the stellar PSF halo centred on the star. The data can be fit well with the model in Eq. (A.3) with a ptel = 0.12% and δtel = 88.4◦. The residuals indicate that the instrumental second order effects and the intrinsic polarization of α Cen A are of order (cid:47) 0.02% and possibly a systematic shift of the polarization towards positive U values in Fig. A.1. How- ever, fitting a model with a systematic shift as additional free parameter does not significantly improve the residuals, from this we conclude that the majority of the residuals are caused by in- strumental effects. The instrument polarization produces in our Q and U differ- ential polarization image a faint copy of the intensity images at the level of 0.1% (see Fig. A.2, top). We correct this by measur- ing ptel,Q and ptel,U (Eq. (A.2)) and subtracting the scaled inten- sity frames ptel,Q · I, ptel,U · I from the corresponding Q and U frames. Fig. A.2 shows the polarized intensity before (top panel) and after (bottom panel) the subtraction. The instrument polarization is not a dominating source of noise for the search of a localized point-source. Our analysis of the noise after different reduction and calibration steps (see Fig. 3) shows only small improvements for most separations. This is understandable if we separate the noise in a Qobs frame with instrument polarization, as defined in Eq. (A.1), into its two main contributions δQ and ptel · δI. The noise of the polarized intensity δQ originates mainly from residuals of the PDI speckle cancellation at smaller separations and photon noise at larger separations (see Fig. 3, green line). The noise in the total in- tensity δI is dominated by speckles at all separations (see Fig. 3, black line). The small scale noise introduced by the instrument polarization is not significant as long as δQ (cid:29) ptel · δI. In our data this condition is usually satisfied, since ptel is on the order of 10−3 while the noise ratio δI/δQ is only on the order of 10. Even though the improvement in point-source contrast is small, we still do the subtraction of the instrument polarization because we want to reach the best possible contrast and remove as many instrumental effects as possible. For the considerations in this section we always assumed that the astrophysical polarization of the target Q and U are zero when averaged over large portion of the image. In Appendix B we investigate in more detail what happens to a non-zero as- trophysical polarization signal of a faint companion if we apply the instrumental polarization correction as described above. We show that the process of subtracting the instrument polarization Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL Fig. A.1. Median telescope polarization for α Cen A from 2017 in N_R-band measured for both Stokes parameters -- Q and U -- as a function of the telescope parallactic angle θpar. The plots show the measurements for camera 1 (+) and camera 2 ((cid:5)). The measurements have been fit with a model function (solid lines), resulting in a measurement of the telescope polarization amplitude ptel and phase δtel for both cameras. as described above has an insignificant effect on the signal of a faint point-source potentially present in the data. Appendix A.3: Differential polarimetric beam shift The even and odd rows of the detectors in ZIMPOL measure the two opposite linear polarization states I(cid:107) and I⊥ simultane- ously and on the same detector pixels. This would in principle allow for a perfect speckle suppression in the final Q and U po- larization frames. However, it was noticed that I(cid:107) and I⊥ are not perfectly aligned on the detector even though they go through the same optical path in the instrument. This unexpectedly large differential polarization beam shift of up to 0.3 pixels (≈ 1 mas) is caused mostly by reflections on inclined mirrors. The effect is for ZIMPOL described in Schmid et al. (2018) and one exam- ple is shown in Fig. A.3. For high-contrast applications the beam shift must be determined and corrected, otherwise the speckles will not cancel sufficiently in the PDI process and a pattern of positive and negative speckle residuals remains. However, the beam shift and possibly also other differential aberration effects cannot be corrected perfectly. One reason for this is the wavelength dependence of the beam shift. This pro- duces for broad-band observations radially elongated speckles, for which the innermost (shortest wavelength) part suffers a dif- ferent beam shift than the outermost (longest wavelength) part. Tests have shown that the beam shift is similar in R and I-band, but it can be significantly different in V-band. Thus, the wave- length dependence is mainly a problem for observations with the VBB filter because it spans a large portion of the visual spec- trum. Observations with any of the R- and I-band filters should not suffer as much. Despite this, it is important to apply a "mean" beam shift correction for any filter, because it can improve the high-contrast performance significantly. Currently, there exists no comprehensive beam shift model for the SPHERE/ZIMPOL instrument. Therefore, the first step is a beam shift measurement, preferentially based on the science data which need to be corrected. This can be achieved with one of the following methods: -- For data taken with the semi-transparent coronagraph CLC- MT-WF or without coronagraph the beam shift can often be measured in individual images as offset of the PSF peak between even- and odd-row (or I⊥ and I(cid:107)) frames. This is only possible if the PSF peak is not saturated and well de- fined so that the relative PSF-offsets can be determined with high precision. The observing conditions influence the mea- suring precision and a careful selection of good beam shift data is essential. For PSF peaks observed through the semi- transparent coronagraph the determinations are difficult if the peak is close to the "edge" of the coronagraphic flux mini- mum. -- In some cases it is easier to derive the beam shift by cross- correlating the speckle pattern or numerically solving for the differential offset of I(cid:107) and I⊥ that minimizes the residuals in QZ = I⊥−I(cid:107). This method often works well because the beam shift is identical for the whole FOV as shown in Fig. A.3. The minimization works better for shorter exposures and narrow filters because the individual speckles are more numerous and better defined. In long exposures the atmospheric tur- bulences smooth short lived speckles and broad-band obser- vations extend the shape of the speckles strongly in radial di- rection, reducing their signal with respect to the background. Both methods allow the determination of the beam shift with a precision of order 0.01 pixels (≈ 0.04 mas) in a single ex- posure for data with good observing conditions. For a proper beam shift correction the high precision is necessary because the effect is only of order 0.1 pixels, but the difference of the cor- rection is noticeable even in a single exposure. Fig.A.3 shows a short exposure of α Cen A before and after applying the beam shift correction. Most of the residual speckle pattern can be sup- pressed, improving significantly the point-source contrast limit in the speckle-dominated region. In the combined final image the effect of the beam shift can not only be seen as additional speckles on a small scale but also as a disturbing feature on a larger scale. This is because the whole speckle halo of the stellar PSF is beam shifted. In our high-contrast images this artificially produces negatively and positively polarized large scale features that can limit the sensitivity for real large scale polarized signals. Unfortunately, the beam shift correction also introduces some new systematic noise residuals. All intensity features orig- inating from components located downstream of the inclined mirrors -- namely the M3 mirror, the pupil tip-tilt mirror and image de-rotator mirrors -- are not subject to the beam shift ef- fect. Applying the beam shift correction to the polarimetric data, Article number, page 17 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Fig. A.2. Comparison of a coronagraphic exposure before (top) and af- ter (bottom) the subtraction of the telescope polarization ptel. The im- ages show the polarized intensity Q of a combined pair of one single zero-phase and one π-phase (2× 1.2 s) exposure of α Cen A in the N_R filter. The axes are in arc seconds and the color scale is in ADU. Fig. A.3. Comparison of a coronagraphic exposure before (top) and af- ter (bottom) the beam shift correction. The images show the fractional polarization Q/I of a single 0-phase and π-phase combined 2 × 1.2 sec- ond exposure of α Cen A in the N_R filter. The axes are in arc seconds and the color scale is dimensionless. as described above, will introduce spurious residual patterns in the "corrected" polarimetric image. This concerns the intensity edges of the attenuating focal plane mask of the coronagraph, or intensity patterns from the dead actuators of the deformable mir- ror, as well as bad pixels, charge traps and dust on the micro-lens array of the ZIMPOL detector (Schmid et al. 2018). An example in Fig. A.3 is the black and white pattern visible at the edge of the coronagraph because of the applied beam shift correc- tion in vertical direction. This effect increases the effective IWA of the result and it cannot be corrected. The intensity patterns from dust on the micro-lens array can be efficiently removed by flat-fielding, and the pixel scale effects can be strongly reduced by masking, bad-pixel cleaning, dithering or angular differential imaging. The beam shift changes continuously with the telescope pointing direction and depends on both the parallactic angle θpar and altitude angle θalt. Therefore it is advantageous for long ob- servations if the beam shift can be accurately measured from the science data itself without requiring any additional overhead. Article number, page 18 of 23 However, for coronagraphic data or saturated data, the beam shift measurement may fail and therefore we regularly take non- coronagraphic PSF measurements. Appendix B: Subtraction of instrument polarization By measuring the residual polarization in our data for different parallactic angles θpar (see Sec. A.2) we determined that all tar- gets of our survey are only weakly polarized (pst (cid:46) 10−4). At this low level of polarization we cannot distinguish any more between the intrinsic polarization of the target and second order instrumental polarization effects. We have also shown that the residual telescope polarization ptel is of order 10−3. In Sec. A.2 we explain how this polarization offset can be measured and re- moved from the data. This is a common way of removing the in- strument polarization and it does not harm any polarized signals in the vicinity of the star as long as the stellar PSF is only polar- ized due to instrumental or interstellar polarization because both processes only add a constant fractional polarization offset to all Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL sources in the FOV. In the following section we investigate the effect of the instrumental polarization correction on a polarized point-source if the star itself exhibits an intrinsic polarization. We analyse this problem with a model for the observed Stokes Iobs(x, y) and Qobs(x, y) signal, with (x, y) being the im- age coordinates (for Uobs(x, y) the analysis is equivalent). In our model the observed intensity distribution Iobs(x, y) consists of contributions from the PSF of the star itself IPSF,st(x, y) and an offset PSF from the planet IPSF,pl(x, y), scaled with the flux con- trast Cflux(α). The intensities IPSF,st(x, y) and IPSF,pl(x, y) are off- set from each other but otherwise identical. The polarized in- tensity distribution Qobs(x, y) consists of the intrinsic polariza- tion of the star pst,QIPSF,st(x, y), the polarized signal from the planet Cpol,Q(α)IPSF,pl(x, y) and a term that describes the polar- ization offset due to instrumental and/or interstellar polarization ptel,Q(IPSF,st(x, y) + Cflux(α)IPSF,pl(x, y)). For the sake of readabil- ity, we omit the dependencies on scattering angle α and image coordinates (x, y) during the derivations. Iobs = IPSF,st + CfluxIPSF,pl Qobs = pst,QIPSF,st + Cpol,QIPSF,pl + ptel,Q(IPSF,st + CfluxIPSF,pl) (B.1) From Eq. (B.1) we want to extract the signal of the polar- ized planet with all instrumental and (inter-)stellar contributions removed, so that the corrected polarized intensity Q(cid:48) obs(x, y) is: Q(cid:48) obs = Cpol,QIPSF,pl = ppl,QCfluxIPSF,pl (B.2) The contrasts Cflux(α) and Cpol,Q(α) are linked by the frac- tional polarization of the planet ppl,Q(α) = Cpol,Q(α)/Cflux(α). Fractional polarizations have an additional subscript Q that can be either negative or positive because they are entries of the two dimensional vector for the linear polarization p = (pQ, pU) in the Stokes Q−U plane. All of the derivations here are equivalent for the Stokes U measurements. a large aperture (cid:104)Qobs(cid:105)/(cid:104)Iobs(cid:105) is given by Using this model, the flux weighted fractional polarization in (cid:104)Qobs(cid:105) (cid:104)Iobs(cid:105) = (cid:104)pst,QIPSF,st + Cpol,QIPSF,pl(cid:105) (cid:104)IPSF,st + CfluxIPSF,pl(cid:105) + ptel,Q (B.3) We usually do not know the exact intrinsic polarization of the star, except that it is small, and contributions from the planet, ex- cept that they are very faint. Therefore, we correct the telescope polarization by just assuming that the two contributions can be neglected and subtract the scaled intensity Iobs like: Qobs,norm = Qobs − (cid:104)Qobs(cid:105) (cid:104)Iobs(cid:105) Iobs (B.4) We then combine Eq. (B.3) and (B.4) and notice that all terms with ptel,Q have successfully vanished: Qobs,norm =pst,QIPSF,st + Cpol,QIPSF,pl − pst,Q(cid:104)IPSF,st(cid:105) + Cpol,Q(cid:104)IPSF,pl(cid:105) (cid:104)IPSF,st(cid:105) + Cflux(cid:104)IPSF,pl(cid:105) We define fpl,st = (cid:104)IPSF,pl(cid:105)/(cid:104)IPSF,st(cid:105) then simplify the result by factoring out the stellar and planetary PSFs: Qobs,norm = − CfluxIPSF,st fpl,st + CfluxIPSF,pl ppl,Q − pst,Q 1 + Cflux fpl,st ppl,Q − pst,Q 1 + Cflux fpl,st (B.6) Now we can estimate the order of magnitude of the different terms for our particular case: -- fpl,st corresponds to the total number of counts in the plane- tary PSF divided by the number of counts in the stellar PSF measured in the ring of pixels around the star that was used to calculate (cid:104)...(cid:105). We determined that the value of this param- eter for the α Cen A data is (cid:46) 3 if the planetary PSF is inside the ring and it is always smaller than one if the planetary PSF is outside the ring. of order 10−7 or smaller. Considering the simplifications above, we can approximate -- The flux contrast of a reflecting planet Cflux is expected to be (1 + Cflux fpl,st) ≈ 1 and simplify Eq. (B.6) to: (cid:16) (cid:17) (cid:32) (cid:33) Qobs,norm ≈ Cflux ppl,Q − pst,Q IPSF,pl 1 − fpl,s IPSF,st IPSF,pl (B.7) Eq. (B.7) still shows two additional terms compared to the desired result in Eq. (B.2). The expression in the right hand bracket in Eq. (B.7) mainly depends on the separation of the stel- lar and planetary PSFs because, for this term to have a small con- tribution to Qobs,norm, the value of IPSF,st(x, y)/IPSF,pl(x, y) needs to be small at the (x, y)-position where the planet PSF peaks. The PSFs for the α Cen A observation in Fig. 6 show that for any separation larger than the IWA of about 0.15(cid:48)(cid:48)this ratio is smaller than ∼ 10−3 and the whole expression in the right hand bracket does not reduce the Q signal of a planet by more than 0.3%. If we neglect this small correction factor, we finally arrive at the expression Qobs,norm(x, y) ≈ Cflux ppl,Q(α) − pst,Q IPSF,pl(x, y) (B.8) (cid:16) (cid:17) for the telescope polarization corrected (or normalized) Q image, where we re-introduced the correct dependencies of the parameters on scattering angle α and image coordinates (x, y). The result means that, for a high-contrast point-source, our pro- cess of removing the instrument polarization modifies the frac- tional polarization of the planet ppl,Q(α) with the intrinsic frac- tional polarization of the star pst,Q. However, given that we have shown (see Appendix A.2) that the polarization of the star pst is (cid:47) 2·10−4 for all our targets, we expect the change of the observed polarization of the planet to be (cid:47) 0.002 %. This is insignificant compared to the expected ppl,Q(α) ≈ 10% of a reflecting planet, therefore we conclude that our process of removing the instru- ment polarization does not harm the polarized signal of a planet significantly in our specific case. Appendix C: Results and discussion for the additional targets (IPSF,st + CfluxIPSF,pl) (B.5) Appendix C.1: α Centauri B In Fig. C.1 we show the deepest contrast limits for α Cen B, derived from a dataset with a total texp of 206.8 min in the VBB Article number, page 19 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Fig. C.1. Radial contrast limits as a function of separation for the deep- est ZIMPOL high-contrast dataset of α Cen B in the VBB filter. The plot shows the 5σ limits for the intensity and polarized intensity from the same observation as well as the 1σ photon noise limit for the po- larized intensity. The meaning of the colors and symbols is the same as explained in Fig. 6. Fig. C.2. Radial contrast limits as a function of separation for the deep- est ZIMPOL high-contrast dataset of Altair in the R_PRIM filter. The plot shows the 5σ limits for the intensity and polarized intensity from the same observation as well as the 1σ photon noise limit for the po- larized intensity. The meaning of the colors and symbols is the same as explained in Fig. 6. filter during a half-night with excellent observing conditions (see Table 1). α Cen B is about four times fainter than the A compo- nent but using the VBB filter instead of the N_R (or N_I) com- pensates for this, resulting in similar count rates and contrast limits as for component A. The raw polarimetric contrast can be improved considerably with the use of PCA-ADI, even surpass- ing the limits for α Cen A for some separations. Because of the excellent observing conditions, despite using the shortest possi- ble DIT of 1.1 sec, some frames have saturated pixels just at the edge of the coronagraph located within the effective IWA of the data (highlighted by the grey bar in the contrast curve plots). From a stability point-of-view, there is no reason why it should not be possible for α Cen B to Harbor a Jupiter sized planet, just like for α Cen A. However, for the B component of the system, the RV limits are much more stringent than for the A component. The radial velocity limits for α Cen B from Zhao (2018) exclude planets with Msin(i) > 8.4 MEarth for the classi- cally defined habitable zone from about 0.7 to 1.3 AU, and even more stringent limits for smaller separations. For α Cen A we have discussed some arguments why the RV limits would still allow a giant planet to be in orbit around this star and that our deep contrast limits could allow us to ob- serve such a giant planet. Due to the stringent RV limits, how- ever, the arguments cannot be applied to α Cen B. Orbital in- clinations with sin(i) (cid:46) 0.5 are unlikely due to stability argu- ments. The resulting optimistic upper limit for the mass of a po- tential companion of ∼ 20 MEarth, comparable to the mass of Neptune or Uranus, makes it unlikely that α Cen B could Harbor a planet large enough to be detectable with the limits presented in Fig. C.1 for a single half-night. The possibility of detecting a low mass planet around α Cen B with ZIMPOL was discussed by Milli et al. (2013) in the light of the former exoplanet candidate α Cen Bb. They concluded that a detection of the reflected light should be possible. However, the study focused on very close separations and the use of a four-quadrant phase-mask corona- graph (not commissioned for ZIMPOL). Such small separations Article number, page 20 of 23 are mostly inaccessible with the Lyot coronagraph used in our survey. A previous search around α Cen B was again done by Schroeder et al. (2000) with HST. The contrast limits are com- parable to α Cen A, between 0.5(cid:48)(cid:48)-1.5(cid:48)(cid:48)the limits are between 7.5-8.5 mag at a wavelength of ∼1.02 µm. Our contrast limits (see Fig. C.1 and summary in Table 2) push these limits for the R+I-band by a large amount. Appendix C.2: Altair In Fig. C.2 we show the deepest contrast limits for Altair derived from a dataset with a total texp of 151.3 min in the R_PRIM fil- ter. The filter is ∼2.6 times broader than the narrow R-band fil- ter used for α Cen A, but Altair is fainter by about the same factor. This results in similar numbers for the captured photons per second and similar contrast limits. The main difference is the larger distance, and therefore the lower expected signal for given planet parameters and angular separation, as shown for our model planet in Fig. C.2. Altair is an active star and a fast rotator, this makes it difficult to use the RV method to determine precise upper mass limits for possible companions. For example, the survey from Lagrange et al. (2009) shows that the RV limits for planets around a fast rotating early type star like Altair allow only the detection of high-mass and short-period exoplanets. Therefore, a direct imag- ing search is competitive and complementary with respect to the RV studies. The deepest contrast limits for Altair were derived in the HST survey of Schroeder et al. (2000) conducted with the Hub- ble Space Telescope (HST) who achieved for separations 0.5(cid:48)(cid:48)- 1.5(cid:48)(cid:48)about 7.5-8.5 mag for a wavelength of ∼1.02 µm. Our limits in intensity are about 12.8-17.7 mag and in polarized intensity 17.9-20.3 mag but with an effective IWA of only ∼0.13(cid:48)(cid:48). The limits are deep in terms of contrast, however, Altair is with 5 pc the most distant object in our survey. As a result of that, the polarimetric contrast of the reflected light from a Jupiter sized Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL planet with our model atmosphere would be lower than 10−8 for all separations larger than 0.22(cid:48)(cid:48). Reaching such contrast levels is not possible with only a few hours of observation. Appendix C.3: Sirius A In Fig. C.3 we show the deepest contrast limits for Sirius A based on a total texp of 175.2 min. The N_I filter was used to avoid saturation of this very bright, blue star in coronagraphic mode with the shortest possible DIT. Unfortunately, the atmospheric conditions for these observations were poor (see Table 1), re- sulting in a degradation of the resolution and the contrast, which makes it more difficult to perform some of the data reduction steps (e.g. centering, beam shift correction, ...). The resulting contrast limits are much worse than what would be possible for such a bright target. The coronagraphic PSF in the right panel in Fig. C.3 shows well that the level of the PSF halo is enhanced and the non-coronagraphic PSF peak is significantly lowered when compared to the PSFs from the observation of α Cen A shown in Fig. 5. Both effects have a negative impact on the SNR of a point-source and the contrast limit of the data. The radial velocity mass/separation limits for low mass ob- jects are loose for Sirius A because it is an intermediate mass (∼2 M(cid:12)) A1V star with strong intrinsic RV variation (e.g. La- grange et al. 2009). The possibility for stable planetary orbits around Sirius A was investigated by Holman & Wiegert (1999) and Bond et al. (2017), suggesting that stable orbits with peri- ods up to 2.24 yr are possible. This corresponds to a semima- jor axis of 2.2 AU or 0.83(cid:48)(cid:48)in angular separation. Bond et al. (2017) used precise HST astrometry and could not exclude the presence of a third body in the system with a mass smaller than ∼15-25 MJupiter. There have been attempts to find massive companions to Sir- ius A in the infrared by Schroeder et al. (2000) using HST and reaching a contrast limit of about 7.5-8.5 mag between 0.5(cid:48)(cid:48)- 1.5(cid:48)(cid:48)for a wavelength of ∼1.02 µm. Thalmann et al. (2011) used Subaru IRCS and AO188 in the 4.05 µm narrow-band Br α fil- ter. At an IWA of 0.7(cid:48)(cid:48)they were able to achieve a contrast of about 11 mag, and about 14 mag at a separation of 1.5(cid:48)(cid:48). The deepest limits were obtained with the IRDIS and IFS instru- ment of SPHERE/VLT in the near infrared from 0.95 to 2.3 µm. (Vigan et al. 2015) using SDI in combination with ADI with an IWA of only 0.2(cid:48)(cid:48). They report contrasts up to 14.3 mag at 0.2(cid:48)(cid:48)and ∼16.3 mag in the 0.4-1.0(cid:48)(cid:48)range. With our combination of PDI and ADI we achieved slightly better contrasts of about 14.7 mag at 0.2(cid:48)(cid:48)and ∼17.1 mag in the 0.4-1.0(cid:48)(cid:48)range in I-band (λ = 817 nm) with a smaller effective IWA of ∼0.13(cid:48)(cid:48). However, our observation suffers from poor observing conditions and con- trast limits like for the other bright targets in our survey (e.g. α Cen A/B) should be possible for Sirius A as well under good seeing conditions. This would improve our polarized intensity contrast limits by about 3 mag at all separations. Our contrast limits are relatively far away from detecting the Jupiter sized reference model planet when comparing to α Cen A/B. This is partially due to the poor observing conditions and partially due to the distance of 2.64 pc which is about twice as far as α Cen. Because the contrast of a companion scales for a given angular separation like L[pc]−2 with distance L. Thus, the reflected light contrast for a reference planet at the same angular separation to its host star is four times more demanding for Sir- ius A compared to α Cen. For bright stars < 1.5m, the contrast efficiency of ZIMPOL is limited by the frame rate, or the abil- ity to collect as many photons as possible without saturating the coronagraphic images. Therefore, the C ∝ L−2 is increasing the required texp for a detection of a planet in reflected light around Sirius A by factor of 16 when compared to α Cen A/B. However, for smaller separations (cid:47) 0.3(cid:48)(cid:48) the reflected light contrast of a possible Jupiter-sized planet around Sirius A increases to values above 2· 10−8 which could be in reach for ZIMPOL within a few consecutive observing nights, assuming a t−1/2 exp noise scaling. Appendix C.4: τ Ceti In Fig. C.4 we show the deepest contrast limits for a half-night of observing τ Ceti during a time with excellent observing con- ditions (see Table 1) derived from a total texp of 168 min in the R_PRIM filter. We used a long tDIT = 14 sec per exposure to ensure that the contrast is photon noise limited in the whole ZIMPOL FOV. A problem with long exposures in P1 mode is rotational smearing. The field rotation can be quite fast because τ Ceti passes close to the zenith. We could have used the broader VBB filter for this observation to maximize the number of col- lected photons, however we selected R_PRIM, because some of the instrumental effects (e.g. instrument polarization, beam shift) can be corrected more accurately in the data reduction for the narrower filters because of strongly wavelength dependent ef- fects which increase with filter width. This strategy seems to be beneficial for the planet search at small separation < 0.4(cid:48)(cid:48) where the speckle noise dominates, while it is less favourable at larger separation in the photon noise limited region. The resulting con- trast limits for one half-night for τ Ceti are not as deep as for most other targets in our survey because of the resulting photon counts are about 10 times lower than for our brighter targets. The presence of RV planets around τ Ceti has been proposed by Tuomi et al. (2013) and Feng et al. (2017). However, the mea- sured signals indicate masses Msin(i) (cid:46) 6.6 MEarth and such planets would be too faint to be observable in our data. High- mass planets >3 MJupiter are excluded by a separate study based on Gaia and Hipparcos astrometry for the separation range 3 -- 30 AU or ≈ 1 − 10(cid:48)(cid:48) (Kervella et al. 2019). Deep direct imaging contrast limits for τ Ceti are given by Schroeder et al. (2000) who achieved between 9.0-11.5 mag in the separation range from 0.5(cid:48)(cid:48)-1.5(cid:48)(cid:48)with HST at a wavelength of ∼1.02 µm. At longer wavelengths, Boehle et al. (2019) report a limiting contrast of about 11.0-12.0 mag in L'-band with NACO at the VLT in the separation range from 1.0(cid:48)(cid:48)-1.5(cid:48)(cid:48). With one ex- cellent night we were able to achieve contrast limits in intensity of about 12.3-16.9 mag and in polarized intensity 16.7-18.8 mag for the separation range from 0.5(cid:48)(cid:48)-1.5(cid:48)(cid:48)but with an IWA down to ∼0.13(cid:48)(cid:48). The contrast limits are deep but -- as shown in Fig. C.4 -- they are still far above what we calculate for our model Jupiter sized planet, both in intensity and polarized intensity. This is be- cause τ Ceti is, with 3.65 pc, one of the more distant targets and with mR = 2.9 also one of the fainter targets of our survey. As a result, the polarimetric contrast of the reflected light from our Jupiter sized reference planet model would be lower than 10−8 for any separations larger than 0.3(cid:48)(cid:48). It is known that τ Ceti hosts a large debris disk and Lawler et al. (2014) measured with Herschel an inner edge between 1 and 10 AU, an outer edge at about ∼55 AU and an inclination of 35◦ ± 10◦ from face-on. The total mass of the disk is estimated to be only ∼1 MEarth (Greaves et al. 2004) and it is extended over a large range of separations. We did not detect the signal of an extended source around τ Ceti in our data, therefore it is either too faint to be seen directly in our data or the inner edge is located outside of our FOV of about 1.7(cid:48)(cid:48)or about 6 AU at this distance. Article number, page 21 of 23 A&A proofs: manuscript no. paper_ZIMPOL_contrast_sh05 Fig. C.3. Left: Radial contrast limits as a function of separation for the deepest ZIMPOL high-contrast dataset of Sirius A in the N_I filter. The plot shows the 5σ limits for the intensity and polarized intensity from the same observation as well as the 1σ photon noise limit for the polarized intensity. The meaning of the colors and symbols is the same as explained in Fig. 6. Right: The coronagraphic PSF compared to the non-coronagraphic PSF of Sirius A in the N_I filter Fig. C.4. Radial contrast limits as a function of separation for the deep- est ZIMPOL high-contrast dataset of τ Ceti in the R_PRIM filter. The plot shows the 5σ limits for the intensity and polarized intensity from the same observation as well as the 1σ photon noise limit for the polar- ized intensity. The meaning of the colors and symbols is the same as in Fig. 6, except that we only subtracted 10 PCs during the PCA-ADI step for the polarized intensity. For this dataset the resulting contrast limits were significantly better with only 10 instead of the 20 PCs that we used for all other dataset. Appendix C.5:  Eridani In Fig. C.5 we show the deepest contrast limits of  Eri for a sin- gle half-night with good observing conditions (see Table 1) and a total texp of 192 min in the VBB filter. This is the only target observed in the field stabilized P2 polarimetry mode. Without field rotation, we cannot apply ADI to this dataset. The targets  Eri and τ Ceti are almost identical in brightness and therefore it is interesting to compare the contrast limits of the non-ADI  Eri Article number, page 22 of 23 Fig. C.5. Radial contrast limits as a function of separation for the deep- est ZIMPOL high-contrast dataset of  Eri in the VBB filter. The plot shows the 5σ limits for the intensity and polarized intensity from the same observation as well as the 1σ photon noise limit for the polarized intensity. The meaning of the colors and symbols is the same as in Fig. 6 data with the ADI τ Ceti data and the different filters used. The achieved contrast for  Eri is significantly deeper at larger sepa- rations because of the longer total texp and the broader filter, and therefore increased photon counts by a factor of ∼2. At closer separations in the speckle-dominated regime the contrast lim- its for τ Ceti are better, despite the smaller amount of collected photons. The ADI data of τ Ceti profit from reduced quasi-static aberrations because the speckles are averaged and significantly reduced by the PSF subtraction. From this comparison we es- timate that field rotation would improve the contrast limits for  Eri inside the speckle ring by up to a factor of 5 for the polar- Hunziker et al.: RefPlanets: Search for reflected light from extra-solar planets with SPHERE / ZIMPOL ization up to 10 for the intensity, but at the time of the observa- tion the addition of ADI was not yet considered as an option. The deepest direct imaging and radial velocity limits for  Eri are both presented in Mawet et al. (2019). They also present the strongest evidence so far for the existence of  Eri b. A giant planet with a mass of ∼1.2 MJupiter for an orbital inclination of 34◦± 2◦ when assumed to be coplanar with the outer debris disk. The star is also monitored by Gaia but the measured astrometric trends are not yet precise enough to confirm the planet (Kervella et al. 2019). The planet's separation of ∼3.5 AU is well within the FOV of ZIMPOL and in the photon noise dominated regime of the contrast curve (see Fig. C.5). However, we were not able to detect the planet because a Jupiter sized reference model planet would produce at this orbital separation, a polarization contrast below 1.2· 10−9 (22.3 mag). Our 5σ contrast limit with one half- night of observation at this separation is 2.2·10−8 (19.1 mag). We were also not able to spot any extended polarized emission from the disk around  Eri. The well known part of the disk around  Eri as seen by Herschel and ALMA is located between 11 -- 13 AU (Greaves et al. 2014; Booth et al. 2017), which is outside of our FOV with ZIMPOL. The deepest high spatial resolution imaging limits for the thermal emission of  Eri b were obtained with VLT/NACO in Lp-band (Mizuki et al. 2016) and Keck/NIRC2 in Ms-band (Mawet et al. 2019). The 5σ contrast curves presented in Mawet et al. (2019) range from 0.3(cid:48)(cid:48)to 1.5(cid:48)(cid:48)with contrast limits in the range from ∼9.0 -- 11.8 mag and ∼10.3 -- 13.5 mag for Lp-band and Ms-band, respectively. With our best half-night we were able to achieve contrast limits of about 10.0 -- 16.3 mag in reflected inten- sity and 15.0 -- 19.6 mag in polarized intensity for the same range of separations with an effective IWA of ∼0.13(cid:48)(cid:48). Due to the broad passband of the VBB filter the contrast limits for  Eri are deep at the separation where  Eri b is expected to be orbiting but this is still a factor of ∼20 or ∼3.3 mag away from the expected signal for planet b. Article number, page 23 of 23
1011.6125
1
1011
2010-11-29T03:49:03
Volatiles and refratories in solar analogs: no terrestial planet connection
[ "astro-ph.EP" ]
We have analysed very high-quality HARPS and UVES spectra of 95 solar analogs, 24 hosting planets and 71 without detected planets, to search for any possible signature of terrestial planets in the chemical abundances of volatile and refractory elements with respect to the solar abundances. We demonstrate that stars with and without planets in this sample show similar mean abundance ratios, in particular, a sub-sample of 14 planet-host and 14 "single" solar analogs in the metallicity range 0.14<[Fe/H]<0.36. In addition, two of the planetary systems in this sub-sample, containing each of them a super-Earth-like planet with masses in the range ~ 7-11 Earth masses, have different volatile-to-refratory abundance ratios to what would be expected from the presence of a terrestial planets. Finally, we check that after removing the Galactic chemical evolution effects any possible difference in mean abundances, with respect to solar values, of refratory and volatile elements practically dissappears.
astro-ph.EP
astro-ph
The Astrophysics of Planetary Systems: Formation, Structure, and Dynamical Evolution Proceedings IAU Symposium No. 276, 2011 Alessandro Sozzetti, Mario G. Lattanzi & Alan P. Boss, eds. c(cid:13) 2011 International Astronomical Union DOI: 00.0000/X000000000000000X Volatiles and refratories in solar analogs: no terrestial planet connection J. I. Gonz´alez Hern´andez1,2, G. Israelian1, N. C. Santos3,4, S. Sousa3, E. Delgado-Mena1, V. Neves3, and S. Udry5 1Instituto de Astrof´ısica de Canarias, C/ Via L´actea s/n, 38200 La Laguna, Spain email: [email protected] 2Dpto. de Astrof´ısica y Ciencias de la Atm´osfera, Facultad de Ciencias F´ısicas, Universidad Complutense de Madrid, E-28040 Madrid, Spain 3Centro de Astrof´ısica, Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal 4Departamento de F´ısica e Astronomia, Faculdade de Ciencias, Universidade do Porto, Portugal 5Observatoire Astronomique de l'Universit´e de Gen`eve, 51 Ch. des Maillettes, -Sauverny- Ch1290, Versoix, Switzerland1 Abstract. We have analysed very high-quality HARPS and UVES spectra of 95 solar analogs, 24 host- ing planets and 71 without detected planets, to search for any possible signature of terrestial planets in the chemical abundances of volatile and refractory elements with respect to the solar abundances. We demonstrate that stars with and without planets in this sample show similar mean abun- dance ratios, in particular, a sub-sample of 14 planet-host and 14 "single" solar analogs in the metallicity range 0.14 < [Fe/H] < 0.36. In addition, two of the planetary systems in this sub- sample, containing each of them a super-Earth-like planet with masses in the range ∼ 7 − 11 Earth masses, have different volatile-to-refratory abundance ratios to what would be expected from the presence of a terrestial planets. Finally, we check that after removing the Galactic chemical evolution effects any possible difference in mean abundances, with respect to solar values, of refratory and volatile elements practically dissappears. Keywords. stars: abundances -- stars: fundamental parameters -- stars: planetary systems -- stars: planetary systems: formation -- stars: atmospheres 1. Introduction The discovery of more than 400 exoplanets orbiting solar-type stars by the radial velocity technique have provided a substantial amount of high-quality spectroscopic data (see e.g. Neves et al. 2009). Recently, Mel´endez et al. (2009) have obtained a clear trend [X/Fe] versus TC in a sample of 11 solar twins, and claimed (see also Ram´ırez et al. 2009, 2010) that the most likely explanation to this abundance pattern is related to the presence of terrestial planets in the solar planetary system. Here we summarize the analysis of very high-quality HARPS and UVES spectroscopic data of a sample of 95 solar analogs with and without planets (see Gonz´alez Hern´andez et al. 2010), with a resolving power of λ/δλ &85,000 and a mean hS/Ni∼850. The stellar parameters and metallicities of the whole sample of stars were computed using the method described in Sousa et al. (2008). The chemical abundance derived for 1 0 1 0 2 v o N 9 2 . ] P E h p - o r t s a [ 1 v 5 2 1 6 . 1 1 0 1 : v i X r a 2 Gonz´alez Hern´andez et al. each spectral line was computed using the LTE code MOOG (Sneden 1973), and a grid of Kurucz ATLAS9 model atmospheres (Kurucz 1993). 2. Metal-rich solar analogs hosting super-Earth-like planets We find no substantial differences in the abundance patterns of solar analogs with and without planets. In particular, the slopes of the abundance ratios [X/Fe] versus TC in two metal-rich stars, HD 1461 and HD 160691, containing each of them one super-Earth-like planet, with 7-11 Earth masses, have the opposite sign to what one would expect if the amount of refractory metals in the atmospheres of planet hosts would depend only on the amount of terrestial planets. HD 1461 ([Fe/H] = +0.2) HD 1461 ([Fe/H] = +0.2) Zr O O C C S R A T S - N U S ] e F X [ ∆ / 0.2 0.1 0.0 -0.1 -0.2 -0.3 -0.4 HD 160691 ([Fe/H] = +0.3) S Zn S Zn Cu Na Mn Na Cu Mn Ba Ca Sr BaCe Eu Mg Fe Si Cr V Ni Co Nd Zr Ti Y Al Sc Nd Zr Ca Sr Ti Ce V Y Al Sc Eu Fe Cr Si Mg Ni Co O C S R A T S - N U S ] e F X [ ∆ / 0.2 0.1 0.0 -0.1 -0.2 -0.3 -0.4 Zn S Eu Mg Fe Si Co Cr Ni Cu Na Mn Na Ce Ba Sr Ca V Ba Ti Nd Y Al Sc Sr Ca Ce V Y Nd Zr Ti Al Sc O C S Zn Cu Mn Eu Ni Fe Co Cr Si Mg HD 160691 ([Fe/H] = +0.3) 0 500 1000 TC (K) 1500 0 500 1000 TC (K) 1500 Figure 1. Left panel: Abundance differences, ∆[X/Fe]SUN−STARS, between the Sun, and 2 planet hosts with super-Earth-like planets. Linear fits for different TC ranges to the data points weighted with the error bars are also displayed. We note the different slopes derived when choosing the range TC > 1200 K (dashed-dotted line) as in Mel´endez et al. (2009) and Gonz´alez Hern´andez et al. (2010), and TC > 900 K (dashed-three-dotted line) as in Ram´ırez et al. (2009, 2010). An arbitrary shift of -0.25 dex has been applied to the abundances of the planet host HD 160691. Right panel: Same as left panel of this figure but after correcting each element abundance ratio of each star using a linear fit to the Galactic chemical trend of the corresponding element at the metallicity of each star. In left panel of Fig. 1 we display the abundances of these two stars and some linear fits for different TC ranges. The steep positive trend in the linear fit for TC > 900 K is probably affected by chemical evolution effects on Mn, Na and Cu. In right panel of Fig. 1 we have already removed the Galactic chemical evolution effects and both stars do not seem to show any trend. We may conclude that it seems plausible that many of our targets hosts terrestrial planets but this may not affect the volatile-to-refratory abundance ratios in the atmospheres of these stars (see e.g. Udry & Santos 2007). References Gonz´alez Hern´andez, J. I., Israelian, G., Santos, N. C., et al. 2010, ApJ, 720, 1592 Kurucz, R. L. ATLAS9 Stellar Atmospheres Programs and 2 km s−1 Grid, CD-ROM No. 13, Smithsonian Astrophysical Observatory, Cambridge, 1993 Mel´endez, J., Asplund, M., Gustafsson, B., & Yong, D. 2009, ApJ Letters, 704, L66 Neves, V., Santos, N. C., Sousa, S. G., Correia, A. C. M., & Israelian, G. 2009, A&A, 497, 563 Ram´ırez, I., Mel´endez, J., & Asplund, M. 2009, A&A Letters, 508, L17 Ram´ırez, I., Asplund, M., Baumann, P., Mel´endez, J., & Bensby, T. 2010, A&A, 521, A33 Sousa, S. G., et al. 2008, A&A, 487, 373 Sneden, C. 1973, PhD Dissertation, Univ. of Texas, Austin Udry, S., & Santos, N. C. 2007, ARA&A, 45, 397
1306.2220
1
1306
2013-06-10T15:04:38
Mixing and Transport of Short-Lived and Stable Isotopes and Refractory Grains in Protoplanetary Disks
[ "astro-ph.EP", "astro-ph.SR" ]
Analyses of primitive meteorites and cometary samples have shown that the solar nebula must have experienced a phase of large-scale outward transport of small refractory grains as well as homogenization of initially spatially heterogeneous short-lived isotopes. The stable oxygen isotopes, however, were able to remain spatially heterogenous at the $\sim$ 6% level. One promising mechanism for achieving these disparate goals is the mixing and transport associated with a marginally gravitationally unstable (MGU) disk, a likely cause of FU Orionis events in young low-mass stars. Several new sets of MGU models are presented that explore mixing and transport in disks with varied masses (0.016 to 0.13 $M_\odot$) around stars with varied masses (0.1 to 1 $M_\odot$) and varied initial $Q$ stability minima (1.8 to 3.1). The results show that MGU disks are able to rapidly (within $\sim 10^4$ yr) achieve large-scale transport and homogenization of initially spatially heterogeneous distributions of disk grains or gas. In addition, the models show that while single-shot injection heterogeneity is reduced to a relatively low level ($\sim$ 1%), as required for early solar system chronometry, continuous injection of the sort associated with the generation of stable oxygen isotope fractionations by UV photolysis leads to a sustained, relatively high level ($\sim$ 10%) of heterogeneity, in agreement with the oxygen isotope data. These models support the suggestion that the protosun may have experienced at least one FU Orionis-like outburst, which produced several of the signatures left behind in primitive chondrites and comets.
astro-ph.EP
astro-ph
Mixing and Transport of Short-Lived and Stable Isotopes and Refractory Grains in Protoplanetary Disks Department of Terrestrial Magnetism, Carnegie Institution for Science, 5241 Broad Branch Road, NW, Washington, DC 20015-1305 Alan P. Boss [email protected] ABSTRACT Analyses of primitive meteorites and cometary samples have shown that the solar nebula must have experienced a phase of large-scale outward transport of small refractory grains as well as homogenization of initially spatially heteroge- neous short-lived isotopes. The stable oxygen isotopes, however, were able to remain spatially heterogenous at the ∼ 6% level. One promising mechanism for achieving these disparate goals is the mixing and transport associated with a marginally gravitationally unstable (MGU) disk, a likely cause of FU Orionis events in young low-mass stars. Several new sets of MGU models are presented that explore mixing and transport in disks with varied masses (0.016 to 0.13 M⊙) around stars with varied masses (0.1 to 1 M⊙) and varied initial Q stability min- ima (1.8 to 3.1). The results show that MGU disks are able to rapidly (within ∼ 104 yr) achieve large-scale transport and homogenization of initially spatially heterogeneous distributions of disk grains or gas. In addition, the models show that while single-shot injection heterogeneity is reduced to a relatively low level (∼ 1%), as required for early solar system chronometry, continuous injection of the sort associated with the generation of stable oxygen isotope fractionations by UV photolysis leads to a sustained, relatively high level (∼ 10%) of heterogeneity, in agreement with the oxygen isotope data. These models support the suggestion that the protosun may have experienced at least one FU Orionis-like outburst, which produced several of the signatures left behind in primitive chondrites and comets. Subject headings: accretion, accretion disks -- hydrodynamics -- instabilities -- planets and satellites: formation -- 2 -- 1. Introduction The short-lived radioisotope (SLRI) 26Al was alive during the formation of the first refractory solids in the solar nebula, the Ca-, Al-rich inclusions (CAIs) found in primitive chondritic meteorites. This means that at least some of the solar system's SLRIs may have been injected into either the presolar cloud (e.g., Boss & Keiser 2012; Boss 2012) or the solar nebula (Ouellette et al. 2007, 2010; Dauphas & Chaussidon 2011) by a supernova or AGB star shock wave. In either case, injection occurred as a single event that was spatially heterogeneous, which would potentially reduce the usefulness of 26Al as a spatially homogeneous chronometer (Dauphas & Chaussidon 2011) for precise studies of the earliest phases of planet formation (MacPherson et al. 2012; cf. Krot et al. 2012). Previous models (e.g., Boss 2011, 2012) have shown how such initial spatial isotopic heterogeneity can be substantially reduced in a marginally gravitationally unstable (MGU) disk, as a result of the large-scale inward and outward transport and mixing of gas and particles small enough to move with the gas (e.g., Boss et al. 2012). Other elements and their isotopes suggest a similarly well-mixed solar nebula (e.g., Os: Walker 2012; Fe: Wang et al. 2013). The stable oxygen isotopes, on the other hand, appear to have been spatially heterogeneous in the solar nebula during the early phases of planet formation; e.g., small refractory particles from Comet 81P/Wild 2 have normalized 17,18O/16O ratios that span the entire solar system range of ∼ 6% variations (Nakashima et al. 2012). The leading explanation for generating these oxygen anomalies is UV photodissociation of CO molecules at the surface of the outer solar nebula (e.g., Podio et al. 2013), where self-shielding could lead to isotopic fractionation between gas-phase and solid-phase oxygen atoms (e.g., Lyons & Young 2005; Krot et al. 2012). CO self-shielding on the irregular, corrugated outer surface of the disk would also lead to initial spatial heterogeneity, though the process would be continuous in time, rather than a single-shot event like a supernova shock wave. Furthermore, the very existence of refractory particles in Comet 81P/Wild 2 (Brownlee et al. 2006; Simon et al. 2008; Nakamura et al. 2008), which are thought to have formed close to the protosun, implies that these small particles experienced large-scale outward transport from the inner solar nebula to the comet-forming regions of the outer solar nebula. MGU disks offer a means to accomplish this early large-scale transport (e.g., Boss 2008, 2011; Boss et al. 2012). Marginally gravitationally unstable disks are likely to be involved in the FU Orionis outbursts experienced by young solar-type stars (e.g., Zhu et al. 2010b; Vorobyov & Basu 2010; Martin et al. 2012). MGU disk models (e.g., Boss 2011) can easily lead to the high mass accretion rates (∼ 10−5M⊙ yr−1) needed to explain FU Orionis events. FU Orionis outbursts are believed to last for about a hundred years and to occur periodically for all low mass protostars (Hartmann & Kenyon 1996; Miller et al. 2011). MGU models are also capable of offering an alternative mechanism (disk instability) for gas giant planet formation (e.g., -- 3 -- Boss 2010; Meru & Bate 2012; Basu & Vorobyov 2012). However, the magnetorotational instability (MRI) is likely to be involved in FU Orionis outbursts as well (Zhu et al. 2009a), with MRI operating in the ionized innermost disk layers as well as at the disk's surfaces. Zhu et al. (2009c, 2010a,b) have constructed one- and two-dimensional (axisymmetric) models of a coupled MGU-MRI mechanism, with MGU slowly leading to a build-up of mass in the innermost disk, which then triggers a rapid MRI instability and an outburst. Alternatively, MRI may operate in the outermost disk, partially ionized by cosmic rays, leading to a build- up of mass in the dead zone at the intermediate disk midplane, thus triggering a phase of MGU transport. Such a coupled mechanism may be crucial for achieving outbursts in T Tauri disks, where the disk masses are expected to be smaller than at earlier phases of evolution. We present here several new sets of MGU disk models that examine the time evolution of isotopic heterogeneity introduced in either the inner or outer solar nebula, by either a single-shot event or a continuous injection process, for a variety of disk and central protostar masses, including protostars with M dwarf masses. Low mass exoplanets are beginning to be discovered around an increasingly larger fraction of M dwarfs (Bonfils et al. 2013; Dressing & Charbonneau 2013; Kopparapu 2013), with a number of these being potentially habitable exoplanets, elevating the importance of understanding mixing and transport processes in M dwarf disks. 2. Numerical Methods The numerical models were computed with the same three dimensional, gravitational hydrodynamics code that has been employed in previous MGU disk models (e.g., Boss 2011). Complete details about the code and its testing may be found in Boss & Myhill (1992). Briefly, the code performs second-order-accurate (in both space and time) hydrodynamics on a spherical coordinate grid, including radiative transfer in the diffusion approximation. A spherical harmonic (Ylm) expansion of the disk's density distribution is used to compute the self-gravity of the disk, with terms up to and including l = m = 32. The radial grid contains 50 grid points for the 10 AU disk models and 100 grid points for the 40 AU disk models. All models have 256 azimuthal grid points, and effectively 45 theta grid points, given the hemispherical symmetry of the grid. The theta grid is compressed around the disk's midplane to provide enhanced spatial resolution, while the azimuthal grid is uniformly spaced. The Jeans length constraint is used to ensure adequate resolution. The inner boundary absorbs infalling disk gas, which is added to the central protostar, while the outer disk boundary absorbs the momentum of outward-moving disk gas, while retaining the gas on the active -- 4 -- grid. The central protostar wobbles in such a manner as to preserve the center of mass of the entire system. The time evolution of a color field is calculated (e.g., Boss 2011) in order to follow the mixing and transport of isotopes carried by the disk gas or by small particles, which should move along with the disk gas. The equation for the evolution of the color field density ρc (e.g., Boss 2011) is identical to the continuity equation for the disk gas density ρ ∂ρc ∂t + ∇ · (ρcv) = 0, where v is the disk gas velocity and t is the time. The total amount of color is conserved in the same way that the disk mass is conserved, as the hydrodynamic equations are solved in conservation law form (e.g., Boss & Myhill 1992). 3. Initial Conditions The initial disk density distributions are based on the approximation derived by Boss (1993) for a self-gravitating disk orbiting a star with mass Ms −(cid:16) γ − 1 γ (cid:17)h(cid:16)2πGσo(R) K ρ(R, Z)γ−1 = ρo(R)γ−1 K (cid:16) 1 (cid:17)Z + GMs R − 1 (R2 + Z 2)1/2(cid:17)i, where R and Z are cylindrical coordinates, G is the gravitational constant, ρo(R) is the midplane density, σo(R) is the surface density, K = 1.7 × 1017 (cgs units) and γ = 5/3. The initial midplane density is while the initial surface density is ρo(R) = ρoi(cid:16) Roi R (cid:17) 3/2 , σo(R) = σoi(cid:16)Roi R (cid:17) 1/2 . The parameters ρoi and σoi and the reference radius Roi are defined in Table 1 for the various disk models explored in this paper. The total amount of mass in the models does not change -- 5 -- during the evolutions; the initial infalling disk envelope accretes onto the disk, and no further mass is added to the system across the outer disk boundary at Ro. The outer disk surfaces are thus revealed to any potential source of UV irradiation. For the 10 AU outer radius disks listed in Tables 2 and 3, the initial disk temperature profiles (Figures 1 and 2) are based on the Boss (1996) temperature profiles, with variations in the assumed outer disk temperature To, chosen in order to study the effect of varied minimum values of the Q stability parameter. Values of Q > 1.5 indicate marginally gravitationally unstable disks. The inner disks are all highly Q stable, with Q >> 1. For the 40 AU outer radius disks listed in Table 4, the initial disk temperatures are uniform at the specified outer disk temperature To, leading to similar initial Q values throughout the disks. For all of the models, the temperature of the infalling envelope is 50 K. The initial color field is added to the surface of the initial disk in an azimuthal sector spanning either 45 degrees (10 AU outer radius disks) or 90 degrees (40 AU outer radius disks) in a narrow ring of width 1 AU, centered at the injection radii listed in the Tables. These models are intended to represent one-time, single-shot injections of isotopic heterogeneity, such as supernova-induced Rayleigh-Taylor fingers carrying live 26Al (e.g., Boss & Keiser 2012). Table 4 lists both single-shot and continuous injection models, where in the latter case the color is added continuously to the same location on the disk surface throughout the evolution, crudely simulating ongoing photodissociation of CO (e.g., Lyons & Young 2005) possibly leading to stable oxygen isotope fractionation between the gas and solid phases. The color field in the latter case is intended to represent isotopically distinct gas or small particles resulting from the UV photochemistry. Note that in both the single-shot and continuous injection models, the total amount of color added is arbitrary (e.g., the color field in the injection volume is simply set equal to 1), and is intended to be scaled to whatever value is appropriate for the isotope(s) under consideration. The color field is a massless, passive tracer that has no effect on the disk's dynamics, so the total amount of color added is irrelevant for the disk's subsequent evolution. The models seek to follow the deviations from uniformity of the color field, not the absolute amounts of color added; the evolution of the dispersion of the color field about its mean radial value, divided by the mean radial value at each instant of time, is the goal of these models. Observations of the DG Tau disk by Podio et al. (2013) have shown that DG Tau itself irradiates its disk's outer layers from 10 AU to 90 AU with a strong UV flux, sufficient for significant UV photolysis and the formation of observable water vapor. Much higher levels of UV irradiation can occur for protoplanetary disks that form in stellar clusters containing massive stars (e.g., Walsh et al. 2013), an environment that has been suggested for our own solar system (e.g., Dauphas & Chaussidon 2011) in order to explain the evidence for live -- 6 -- SLRIs found in primitive meteorites. The fact that molecular hydrogen constitutes the great majority of a disk's mass, yet cannot be directly detected, except at the star-disk boundary region, means that estimates of disk masses are uncertain at best (e.g., Andrews & Williams 2007), as they are typically based on an assumed ratio between the amount of mm-sized dust grains and the total disk mass. Isella et al. (2009) estimated that low- and intermediate-mass pre-main-sequence stars form with disk masses ranging from 0.05 to 0.4 M⊙. DG Tau's disk has a mass estimated to be as high as 0.1 M⊙ (Podio et al. 2013). Recently, the mass of the TW Hydra disk was revised upward to at least 0.05 M⊙ (Bergin et al. 2013). These and other observations suggest that the MGU disk masses assumed in these models may be achieved in some fraction of protoplanetary disks, and perhaps in the solar nebula as well. In fact, Miller et al. (2011) detected a FU Orionis outburst in the classical T Tauri star LkHα 188-G4. Disk masses are typically thought to be < 0.01M⊙ for such stars. The fact that a FU Orionis event occurred in LkHα 188-G4 shows that even the disks around Class II-type objects can experience instabilities leading to rapid mass accretion, e.g., MGU disk phases. 4. Results We present results for a variety of protostellar and protoplanetary disk masses, varied initial minimum Q stability parameters, and varied injection radii, for disks of two different sizes. 4.1. 10 AU Outer Radius Disks 4.1.1. G Dwarf Disks Table 2 shows the initial conditions for the models with a 0.019M⊙ disk in orbit around a 1.0M⊙ protostar. The disks extend from 1 AU to 10 AU, as in the models by Boss (2008, 2011). The main difference from these previous models is that the disk mass (0.019M⊙) is considerably lower than that of the previous models (0.047M⊙). As a result, the initial minimum Q values are considerably higher than in the previous models, ranging from 2.2 to 3.1, compared to the previous range of 1.4 to 2.5. The present models are thus less gravitationally unstable initially than the disks previously considered, with the goal being to learn whether or not the previous results will change for higher values of min Qi. Figure 1 displays the initial midplane temperature profiles for these models. Only the outermost regions of the disks are cool enough to be gravitationally unstable, but the models show that -- 7 -- this is sufficient to result in qualitatively similar behavior for all of the Table 2 models. Figures 3 and 4 show the equatorial plane distribution of the color/gas ratio (ρc/ρ) for model 1.0-2.6-9. This ratio is plotted, as it is equivalent to the 26Al/27Al and 17,18O/16O ratios measured by cosmochemists, i.e., the abundance of an injected or photolysis product species, divided by that of a species that was prevalent in the pre-injection disk. Figure 3 shows that the initial disk surface injection at 9 AU has resulted in the rapid transport of the color field downward to the disk's midplane, as well as inward to close to the inner disk boundary at 2 AU. The vigorous three dimensional motions of a MGU disk are responsible for this large-scale transport in just 34 yr. At this time, the color/gas ratio is still highly heterogeneous, but Figure 4 shows that only 146 yr later, the color/gas ratio has been strongly homogenized throughout the entire disk midplane. Figure 5 shows the evolution of the dispersion of the ratio of the color density to the gas density for models 1.0-2.6-9 and 1.0-2.6-2 at two times. These models differ only in the injection radius, either 9 AU or 2 AU. The dispersion plotted in Figure 5 is defined to be the square root of the sum of the squares of the color field divided by the gas density, subtracted by the azimuthal average of this ratio at a given orbital radius, divided by the square of the azimuthal average at that radius, normalized by the number of azimuthal grid points, and plotted as a function of radius in the disk midplane. Figure 5 shows that the isotopic dispersion is a strong function of orbital radius and time, with the dispersion initially being relatively large (i.e., at 180 yr, in spite of the apparent homogeneity seen in Figure 4 at the same time) as a result of the isotopes traveling downward and radially inward and outward. However, the dispersion decreases dramatically in the outer disks for both models by 777 yr to a value of ∼ 1% to 2%. In fact, the dispersion in both models 1.0-2.6-9 and 1.0-2.6-2 evolves toward essentially the same radial distribution by this time, showing that the exact location of the injection location has little effect on the long term evolution of the distribution: that is controlled solely by the evolution of the underlying MGU disk, which is identical for these two models (i.e., the color fields are passive tracers, and have no effect on the disk's evolution). Note that any small refractory grains present in the initial disk will be carried along with the disk gas, so that some of the grains that start out at 2 AU will be transported to the outermost disk, in the same manner that some of the gas is transported outward. Most of the gas and dust, however, is accreted by the growing protostar. Figure 6 shows the results for three models with varied min Qi, i.e., models 1.0-2.6-9, 1.0-2.9-9, and 1.0-3.1-9, all after 1370 yr. It can be seen that in spite of the variation in the initial degree of instability, the dispersions in the outermost disks all converge to similar values of ∼ 1% to 2%. This suggests that MGU disk evolutions are not particularly sensitive to the exact choice of the initial Q profile, a result that was also found by Boss (2011) for -- 8 -- somewhat more massive disks. As also found by Boss (2011), the dispersions in the innermost disks are significantly higher (∼ 10% to 20%) than in the outermost disks, a direct result of the stronger mixing associated with the cooler outer disks, in spite of the longer orbital periods in the outer disks. 4.1.2. M Dwarf Disks Table 3 shows the initial conditions for the models with either 0.016M⊙ disks around 0.1M⊙ protostars, or 0.018M⊙ disks around 0.5M⊙ protostars. In either case, the disks extend from 1 AU to 10 AU. These models are of interest for exploring how conditions might vary between disks around G dwarfs and M dwarfs, with possible ramifications for the habitability of any rocky planets that form (e.g., Raymond et al. 2007) around M dwarfs. Figure 2 shows the initial midplane temperature profiles for these models. Figure 7 shows the time evolution of the dispersion for model 0.1-1.8-2, appropriate for a late M dwarf protostar. As in all the models, it can be seen that the initially highly het- erogeneous disk becomes rapidly homogenized, in this case by about 5000 yr. Note that this time scale is considerably longer than that for G dwarf disk mixing and transport processes, as a result of the longer Keplerian orbital periods for lower mass, M dwarf protostars. As in the G dwarf protostar models (e.g., Figure 5), the inner disk dispersion is higher than in the outer disk, though in these models (with a lower initial min Q = 1.8) the inner disk dispersion drops to ∼ 5% to 10%, compared to ∼ 1% to 2% in the outer disk. Figure 8 shows the same behavior for model 0.1-1.8-9, which differs from the previous model shown in Figure 7 only in having the injection occur at 9 AU instead of 2 AU. As in the G dwarf disks, the dispersions for both of these models evolve toward essentially identical radial dis- tributions: the underlying MGU disk evolution determines the outcome for the dispersions. Similar results hold for the models with 0.5M⊙ protostars, i.e., early M dwarf disks. 4.2. 40 AU Outer Radius Disks We now turn to a consideration of the consequence of single-shot versus continuous injection at the surface of much larger outer radius (40 AU) disks than have been considered to date for G dwarf stars; Boss (2007) considered disks extending from 4 AU to 20 AU in radius. Table 4 shows the initial conditions for the models with a 0.13M⊙ disk around a 1.0M⊙ protostar, with the disks extending from 10 AU to 40 AU. Because of the much larger inner and outer disk radii for this set of models, these models can be calculated for times -- 9 -- as long as ∼ 3 × 104 yr (Table 4). Such times are still considerably less than the typical ages (∼ 106 yr) of T Tauri stars, implying that in order for MGU disks to occur at such late phases, a prior phase of coupled MRI-MGU evolution might be required to make the present results relevant. Figures 9 and 10 display the evolution of the dispersions for models 1.0-1.1-40-20 and 1.0- 1.1-40-20c, differing only in that the former model has single-shot injection while the latter model has continuous injection, intended to simulate a disk with ongoing UV photolysis and fractionation at the outer disk surface. For model 1.0-1.1-40-20, it can be seen that the evolution is similar to that of the previous single-shot models: a rapid drop in the dispersion, followed by homogenization to ∼ 1% to 2% away from the inner disk boundary. The higher dispersions seen near the outer disk boundaries (∼ 40 AU) are largely caused by the unphysical pile-up of considerable disk mass at 40 AU and should be discounted. However, for the continuous injection model shown in Figure 10, it can be seen that the dispersions throughout the disk even after ∼ 104 yr can be as high as ∼ 20%, consistent with the much larger variation in stable oxygen isotope ratios, compared to SLRI ratios. In a calculation with finer spatial grid resolution, as well as perhaps sub-grid mixing processes, one might expect even stronger homogenization to occur, so the dispersion levels obtained from the present models should be considered to be upper bounds. The total amount of color added during the continuous injection models is large, compared to single-shot injection models: for model 1.0-1.4-40-20c, for example, after 200 yr, the total amount of color injected has increased by a factor of ∼ 90 compared to the single-shot total, and by another factor of ∼ 60 after 27000 yr. Figures 11 and 12 compare the results for continuous injection at either 20 AU or 30 AU, respectively, i.e., for models 1.0-1.1-40-20c and 1.0-1.1-40-30c. In spite of the different injection radii, Figures 11 and 12 show that even at a relatively early phase (405 yr) of evolution, the midplane color/gas ratios look somewhat similar; as before, the MGU disk evolution is the same for both models, and that is the primary determinant of the long term evolution. Finally, similar results as those shown in Figures 9-12 were obtained for the other models listed in Table 4. These models show that the main factor in determining the radial dispersion profile is whether the injection occurs in a single-shot or continuously; in the latter case, the MGU disk does its best to homogenize the color field, but the fact that spatial heterogeneity is being continuously injected limits the degree to which this heterogeneity can be reduced. -- 10 -- 5. Discussion While dust grains in the interstellar medium are overwhelmingly amorphous, crystalline silicate grains have been found in a late M dwarf (SST-Lup3-1) disk at distances ranging from inside 3 AU to beyond 5 AU, in both the midplane and surface layers (Mer´ın et al. 2007). Such crystalline silicate grains are likely to have been produced by thermal annealing in the hottest regions of the disk, well inside of 1 AU (Sargent et al. 2009). Again, outward transport seems to be required to explain the observations, and the results for the models with a 0.1 M⊙ protostar suggest that MGU phases in low mass M dwarf disks may be responsible for these observations. In fact, crystalline mass fractions in protoplanetary disks do not appear to correlate with stellar mass, luminosity, accretion rate, disk mass, or the disk to star ratio (Watson et al. 2009). These results also appear to be consistent with the results of the present models, which show that MGU disk phases are equally capable of relatively rapid large-scale mixing and transport, regardless of the stellar or disk mass, or the exact value of the Q stability parameter. 6. Conclusions These models have shown a rather robust result, namely that a phase of marginal gravitational instability in disks and stars with a variety of masses and disk temperatures can lead to relatively rapid inward and outward transport of disk gas and small grains, as required to drive the protostellar mass accretion associated with FU Orionis events, as well as to explain the discovery of refractory grains in Comet 81P/Wild 2. A MGU disk phase driving a FU Orionis outburst is astronomically quite likely to have occurred for our protosun, and cosmochemically convenient for explaining the relative homogeneity of 26Al/27Al ratios derived from a supernova injection event, and the range of 17,18O/16O ratios derived from sustained UV self-shielding at the surface of the outer solar nebula. Low-mass stars, from G dwarfs to M dwarfs, may well experience a similar phase of MGU disk mixing and transport. In this context, it is worthwhile to note that FU Orionis itself, the prototype of the FU Orionis outburst phenomenon, has a mass of only ∼ 0.3M⊙ (Zhu et al. 2007, 2009b; Beck & Aspin 2012), i.e., the mass of a M dwarf, suggesting that M dwarf protoplanetary disks may experience evolutions similar to that of the solar nebula, with possible implications for the habitability of any resulting planetary system (e.g., Raymond et al. 2007; Bonfils et al. 2013; Dressing & Charbonneau 2013; Kopparapu 2013). I thank Jeff Cuzzi for his comments, the referee for a number of suggested improve- ments, and Sandy Keiser, Michael Acierno, and Ben Pandit for their support of the cluster -- 11 -- computing environment at DTM. This work was partially supported by the NASA Origins of Solar Systems Program (NNX09AF62G) and is contributed in part to the NASA Astro- biology Institute (NNA09DA81A). Some of the calculations were performed on the Carnegie Alpha Cluster, the purchase of which was partially supported by a NSF Major Research Instrumentation grant (MRI-9976645). REFERENCES Andrews, S. M., & Williams, J. P. 2007, ApJ, 659, 705 Basu, S., & Vorobyov, E. I. 2012, Meteorit. Planet. Sci., 47, 1907 Beck, T. L., & Aspin, C. 2012, AJ, 143, 55 Bergin, E. A., et al. 2013, Nature, 493, 644 Bonfils, X., et al. 2013, A&A, 549, A109 Boss, A. P. 1993, ApJ, 417, 351 Boss, A. P. 1996, ApJ, 469, 906 Boss, A. P. 2007, ApJ, 660, 1707 Boss, A. P. 2008, Earth Planet. Sci. Lett., 268, 102 Boss, A. P. 2010, ApJ, 725, L145 Boss, A. P. 2011, ApJ, 739, 61 Boss, A. P. 2012, Ann. Rev. Earth Planet. Sci., 40, 23 Boss, A. P., & Keiser, S. A. 2012, ApJ, 756, L9 Boss, A. P., & Myhill, E. A. 1992, ApJS, 83, 311 Boss, A. P., Alexander, C. M. O'D., & Podolak, M. 2012, Earth Planet. Sci. Lett., 345-348, 18 Brownlee, D. E. et al. 2006, Science, 314, 1711 Dauphas, N., & Chaussidon, M. 2011, Ann. Rev. Earth Planet. Sci., 39, 351 Dressing, C. D., & Charbonneau, D. 2013, ApJ, 767, 95 Hartmann, L., & Kenyon, S. J. 1996, ARAA, 34, 207 Isella, A., Carpenter, J. M., & Sargent, A. I. 2009, ApJ, 701, 260 Kopparapu, R. K. 2013, ApJL, 767, L8 Krot, A. N., et al. 2012, Meteorit. Planet. Sci., 47, 1948 -- 12 -- Lyons, J. R., & Young, E. D. 2005, Nature, 435, 317 MacPherson, G. J., Kita, N. T., Ushikubo, T., Bullock, E. S., & Davis, A. M. 2012, Earth Planet. Sci. Lett., 331-332, 43 Martin, R. G., Lubow, S. H., Livio, M., & Pringle, J. E. 2012, MNRAS, 423, 2718 Mer´ın, B., et al. 2007, ApJ, 661, 361 Meru, F., & Bate, M. R. 2012, MNRAS, 427, 2022 Miller, A. A., et al. 2011, ApJ, 730, 80 Nakamura, T., et al. 2008, Science, 321, 1664 Nakashima, D., et al. 2012, Earth Planet. Sci. Lett., 357-358, 355 Ouellette, N., Desch, S. J., & Hester, J. J. 2007, ApJ, 662, 1268 Ouellette, N., Desch, S. J., & Hester, J. J. 2010, ApJ, 711, 597 Podio, L., et al. 2013, ApJL, 766, L5 Raymond, S. N., Scalo, J., and Meadows, V. S. 2007, ApJ, 669, 606 Sargent, B. A., et al. 2009, ApJ, 690, 1193 Simon, S. B., et al. 2008, Meteorit. Planet. Sci., 43, 1861 Vorobyov, E. I., & Basu, S. 2010, ApJL, 714, L133 Walker, R. J. 2012, Earth Planet. Sci. Lett., 351-352, 36 Walsh, C., Millar, T. J., & Nomura, H. 2013, ApJL, 766, L23 Wang, K., et al. 2013, Meteorit. Planet. Sci., 48, 354 Watson, D. M., et al. 2009, ApJS, 180, 84 Zhu, Z., Hartmann, L., Calvet, N., et al. 2007, ApJ, 669, 483 Zhu, Z., Hartmann, L., & Gammie, C. 2009a, ApJ, 694, 1045 Zhu, Z., Espaillat, C., Hinkle, K., et al. 2009b, ApJL, 694, L64 Zhu, Z., Hartmann, L., Gammie, C., & McKinney, J. C. 2009c, ApJ, 701, 620 Zhu, Z., Hartmann, L., & Gammie, C. 2010a, ApJ, 713, 1143 Zhu, Z., Hartmann, L., Gammie, C., et al. 2010b, ApJ, 713, 1134 This preprint was prepared with the AAS LATEX macros v5.2. -- 13 -- Table 1. Density (ρoi) and surface density (σoi) parameters at a reference radius (Roi) for varied disk masses (Md) in orbit around varied mass protostars (Ms). Ri and Ro are the inner and outer disk boundaries, respectively. Ms (M⊙) Md (M⊙) ρoi σoi Roi (AU) Ri (AU) Ro (AU) 1.0 0.5 0.1 1.0 0.019 0.018 0.016 0.13 4.0 × 10−10 3.1 × 10−10 2.0 × 10−10 1.0 × 10−10 3.2 × 103 2.9 × 103 2.5 × 103 2.0 × 103 1.0 1.0 1.0 4.0 1.0 1.0 1.0 10.0 10.0 10.0 10.0 40.0 -- 14 -- Table 2. Initial conditions and final times (tf ) for models with a 10 AU outer radius, 0.019M⊙ disk in orbit around a 1.0M⊙ protostar. To is the outer disk temperature, min Qi is the minimum value of the initial Q disk parameter, and rinject is the radius where the color is injected at the disk's surface. model To (K) min Qi rinject (AU) tf (yr) 1.0-2.2-2 1.0-2.2-9 1.0-2.6-2 1.0-2.6-9 1.0-2.9-2 1.0-2.9-9 1.0-3.1-2 1.0-3.1-9 15 15 20 20 25 25 30 30 2.2 2.2 2.6 2.6 2.9 2.9 3.1 3.1 2 9 2 9 2 9 2 9 2520 2520 2043 2043 1200 1400 1100 1300 -- 15 -- Table 3. Initial conditions and final times for models with 10 AU outer radius disks (Md) in orbit around lower mass protostars (Ms), as in Table 2. model Ms (M⊙) Md (M⊙) To (K) min Qi rinject (AU) tf (yr) 0.1-1.8-2 0.1-1.8-9 0.5-2.4-2 0.5-2.4-9 0.1 0.1 0.5 0.5 0.016 0.016 0.018 0.018 40 40 25 25 1.8 1.8 2.4 2.4 2 9 2 9 8040 8040 4660 4660 -- 16 -- Table 4. Initial conditions and final times for models with a 40 AU outer radius, 0.13M⊙ disk in orbit around a 1.0M⊙ protostar, as in Table 2. model To (K) min Qi rinject (AU) injection mode tf (yr) 1.0-1.1-40-20 1.0-1.1-40-30 1.0-1.4-40-20 1.0-1.4-40-30 1.0-1.1-40-20c 1.0-1.1-40-30c 1.0-1.4-40-20c 1.0-1.4-40-30c 30 30 50 50 30 30 50 50 1.1 1.1 1.4 1.4 1.1 1.1 1.4 1.4 20 30 20 30 20 30 20 30 single-shot single-shot single-shot single-shot continuous continuous continuous continuous 25000 24500 24000 15000 19500 19800 27000 27000 -- 17 -- Fig. 1. -- Initial midplane temperature distributions (Boss 1996) for models 1.0-2.2-2 and 1.0-2.2-9 (blue), 1.0-2.6-2 and 1.0-2.6-9 (red), 1.0-2.9-2 and 1.0-2.9-9 (green), and 1.0-3.1-2 and 1.0-3.1-9 (black). -- 18 -- Fig. 2. -- Initial midplane temperature distributions (Boss 1996) for models 0.1-1.8-2 and 0.1-1.8-9 (red) and 0.5-2.4-2 and 0.5-2.4-9 (black). -- 19 -- Fig. 3. -- Logarithm of the ratio of the color field to the gas density (log ρc/ρ) in the disk's midplane (arbitrary units) after 34 yr for model 1.0-2.6-9, with single-shot color injection at the initial disk's surface at 9 AU. Region shown is 20 AU in diameter. A 1.0 M⊙ protostar lies at the center of the MGU disk. The color field has been transported inward to close to the inner boundary at 1 AU. The color to gas ratio is still highly heterogeneous. -- 20 -- Fig. 4. -- Same as Figure 1, but after 180 yr. The color field has been transported throughout the disk and the color to gas ratio is now relatively homogeneous. -- 21 -- Fig. 5. -- Time evolution of the azimuthal dispersion of the color to gas ratio as a function of disk radius for models with single-shot surface injection at either 9 AU or 2 AU, respectively: model 1.0-2.6-9 (red at 180 yr, blue at 777 yr) and model 1.0-2.6-2 (black at 180 yr, green at 777 yr). While initially quite different, the dispersions converge by 777 yr. -- 22 -- 0.4 0.2 0 0 2 4 6 8 10 Fig. 6. -- Comparisons of the azimuthal dispersions after 1370 yr of the color to gas ratio as a function of disk radius, for models with single-shot surface injection at 9 AU, but varied initial minimum Q values of 2.6, 2.9, and 3.1: models 1.0-2.6-9 (red), 1.0-2.9-9 (green), and 1.0-3.1-9 (blue), respectively. The outer disks have become relatively homogenized. -- 23 -- Fig. 7. -- Time evolution of the azimuthal dispersion of the color to gas ratio as a function of disk radius for model 0.1-1.8-2, with single-shot injection at 2 AU and a 0.1 M⊙ protostar, at times: black - 71 yr, red - 320 yr, green - 430 yr, blue - 5300 yr, and cyan - 8040 yr. -- 24 -- Fig. 8. -- Time evolution of the azimuthal dispersion of the color to gas ratio as a function of disk radius for model 0.1-1.8-9, with single-shot injection at 9 AU and a 0.1 M⊙ protostar, at times: black - 140 yr, red - 320 yr, green - 430 yr, blue - 5300 yr, and cyan - 8040 yr. -- 25 -- 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40 Fig. 9. -- Time evolution of the azimuthal dispersion of the color to gas ratio as a function of disk radius for model 1.0-1.1-40-20, with single-shot injection at 20 AU on the surface of a 40 AU radius disk around a 1.0 M⊙ protostar, at times: black - 200 yr, red - 5200 yr, green - 12400 yr, and blue - 19700 yr. -- 26 -- Fig. 10. -- Time evolution of the azimuthal dispersion of the color to gas ratio as a function of disk radius for model 1.0-1.1-40-20c, with continuous injection at 20 AU on the surface of a 40 AU radius disk around a 1.0 M⊙ protostar, at times: black - 200 yr, red - 5200 yr, green - 12600 yr, and blue - 19600 yr. -- 27 -- Fig. 11. -- Log ρc/ρ in the midplane after 405 yr for model 1.0-1.1-40-20c, with continuous color injection at the disk's surface at 20 AU. Region shown is 80 AU in diameter. A 1.0 M⊙ protostar lies at the center of the MGU disk. -- 28 -- Fig. 12. -- Same as Figure 9, except at 405 yr for model 1.0-1.1-40-30c, with continuous color injection at the disk's surface at 30 AU, instead of 20 AU.
1903.01890
1
1903
2019-03-05T15:14:09
Effect of dust size and structure on scattered light images of protoplanetary discs
[ "astro-ph.EP" ]
We study scattered light properties of protoplanetary discs at near-infrared wavelengths for various dust size and structure by performing radiative transfer simulations. We show that different dust structures might be probed by measuring disk polarisation fraction as long as the dust radius is larger than the wavelength. When the radius is larger than the wavelength, disc scattered light will be highly polarised for highly porous dust aggregates, whereas more compact dust structure tends to show low polarisation fraction. Next, roles of monomer radius and fractal dimension for scattered light colours are studied. We find that, outside the Rayleigh regime, as fractal dimension or monomer radius increases, colours of the effective albedo at near-infrared wavelengths vary from blue to red. Our results imply that discs showing grey or slightly blue colours and high polarisation fraction in near-infrared wavelengths might be explained by the presence of large porous aggregates containing sub-microns sized monomers.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 17 (2019) Preprint 6 March 2019 Compiled using MNRAS LATEX style file v3.0 Effect of dust size and structure on scattered-light images of protoplanetary discs Ryo Tazaki,1(cid:63) H. Tanaka,1 T. Muto,2 A. Kataoka,3 and S. Okuzumi4 1Astronomical Institute, Graduate School of Science, Tohoku University, 6-3 Aramaki, Aoba-ku, Sendai 980-8578, Japan 2Division of Liberal Arts, Kogakuin University, 1-24-2 Nishi-Shinjuku, Shinjuku-ku, Tokyo 163-8677, Japan 3National Astronomical Observatory of Japan, Mitaka, Tokyo 181-8588, Japan 4Department of Earth and Planetary Sciences, Tokyo Institute of Technology, Meguro, Tokyo, 152-8551, Japan Accepted XXX. Received YYY; in original form ZZZ ABSTRACT We study scattered light properties of protoplanetary discs at near-infrared wave- lengths for various dust size and structure by performing radiative transfer simula- tions. We show that different dust structures might be probed by measuring disk polarisation fraction as long as the dust radius is larger than the wavelength. When the radius is larger than observing wavelength, disc scattered light will be highly po- larised for highly porous dust aggregates, whereas more compact dust structure tends to show low polarisation fraction. Next, roles of monomer radius and fractal dimension for scattered light colours are studied. We find that, outside the Rayleigh regime, as fractal dimension or monomer radius increases, colours of the effective albedo at near- infrared wavelengths vary from blue to red. Our results imply that discs showing grey or slightly blue colours and high polarisation fraction in near-infrared wavelengths might be explained by the presence of large porous aggregates containing sub-microns sized monomers. Key words: infrared: ISM -- protoplanetary discs -- radiative transfer 1 INTRODUCTION The first step of planet formation is coagulation of dust par- ticles in protoplanetary discs. Initial growth of dust parti- cles yields fractal dust aggregates, whose porosity is higher than 99% (Kempf et al. 1999; Ormel et al. 2007; Okuzumi et al. 2012). According to grain growth studies, porosity of an aggregate in discs is a matter of debates; some stud- ies suggest highly porosity (> 85%) (e.g., Wada et al. 2008; Suyama et al. 2008), others predicts lower porosity (67−85%) (e.g., Blum et al. 2006). In our Solar System, both fractal aggregates and compact particles are found, e.g., in comet 67P/Churyumov-Gerasimenko (Fulle et al. 2015, 2016; Man- nel et al. 2016; Bentley et al. 2016). In a debris disc, the pres- ence of mildly porous particles (porosity of about 60%) is suggested by observations (e.g., Augereau et al. 1999). How- ever, porosity of dust aggregates in a protoplanetary disc is still not clear observationally. Since dust porosity is a key to overcome bouncing barrier (Wada et al. 2011; Kothe et al. 2013; Brisset et al. 2017) and radial drift barrier (Okuzumi et al. 2012; Kataoka et al. 2013) in dust growth, characteri- sation of dust porosity from disc observations may shed light on dust evolution in discs. (cid:63) E-mail: [email protected] © 2019 The Authors Since light scattering properties sensitively depend on size, structure, and composition of dust particles (e.g., Shen et al. 2008, 2009; Paper I; Paper II; Halder et al. 2018; Ysard et al. 2018), disc scattered light may provide information of dust properties. Since disc scattered light has been com- monly detected in optical- and near-infrared-wavelengths, in this study, we investigate a connection between dust prop- erties and near-infrared disc scattered light. Observed disc scattered light colours are thought to reflect wavelength dependence of the albedo of dust par- ticles (e.g., Mulders et al. 2013). In near-infrared wave- lengths, most protoplanetary discs show grey colours in to- tal intensity (e.g., Fukagawa et al. 2010), though some discs have reddish colours (e.g., Mulders et al. 2013; Long et al. 2017). More recently, Avenhaus et al. (2018) found that discs around T-Tauri stars have almost grey colours in polarised intensity. The relation between radii of compact spherical grains and disc colours has so far been studied (e.g., Mulders et al. 2013). Although Min et al. (2012) and Kirchschlager & Wolf (2014) studied how dust structure and grain porosity affect disc scattered light, respectively, these studies did not investigate colours. Hence, a role of dust structure on disc colours has not yet been clarified well. Furthermore, Min et al. (2012) adopted the effective medium theory to com- 2 R. Tazaki et al. pute optical properties of fluffy aggregates; however, once EMT is applied to fluffy aggregates, it significantly overesti- mates the degree of forward scattering (Paper I; Paper II). Hence, more accurate opacity model should be used. Kirch- schlager & Wolf (2014) properly treated optical properties of porous grains; however, they only considered grain porosity up to 60%, whereas fluffy dust aggregates that may form in discs have much higher porosity, e.g., more than 99% (e.g., Kataoka et al. 2013). An aim of this study is to examine a relation between disc scattered light (total intensity, polarised intensity, and colours) and dust properties (size and structure). As a dust model, we particularly focus on highly porous dust aggre- gates (porosity higher than 85%), whose presence has been suggested by grain growth models (Kempf et al. 1999; Ormel et al. 2007; Okuzumi et al. 2012). Disc scattered light im- ages are obtained by performing 3D Monte Carlo radiative transfer simulations and optical properties of porous aggre- gates are obtained by using a rigorous numerical method, the T-Matrix Method (Mackowski & Mishchenko 1996), or a carefully tested approximate method (Paper I; Paper II). This paper is organised as follows. We first explain, in Section 2, our star and disc models used in this paper. We also summarise dust particle models and their optical prop- erties. Then, we perform radiative transfer simulations of protoplanetary discs at near-infrared wavelengths and dis- cuss images (Section 3) and disc scattered light colours (Sec- tion 4). In Section 5, we study scattering properties of dust aggregates with various structure. Finally, we compare our results with disc observations and close with a conclusion section. 2 MODELS 2.1 Star and disc model The central star is assumed to be a T-Tauri star with the effective temperature of 4000 K and the radius of 2R(cid:12). The dust surface density is assumed to be the single power-law distribution, Σd = Σ0(R/Rout)−1, where R is the distance from the central star, Σ0 is the surface density of dust grains at the outer radius Rout. The disc inner and outer radii are trun- cated at 10 au and 100 au, respectively. The total dust disc mass is set as 10−4M(cid:12). All simulations shown in this paper has the same total dust disk mass. The vertical distribution of dust grains is assumed to be the Gaussian distribution with the dust scale height of Hd = 3.3× 10−2(R/1 au)β au. In this paper, we consider a flared disc geometry with β = 1.25 (Kenyon & Hartmann 1987). In Section 4.3, we address how the choice of disc geometry affects our conclusion. We perform radiative transfer simulations using a 3D Monte Carlo radiative transfer code, radmc-3d (Dullemond et al. 2012). The radial mesh is uniformly spaced in a log- arithmic space of [5,105] au with 256 grids. The zenith and azimuthal angles are uniformly spaced in a linear space of [2π/9, 7π/9] and [0, 2π] with 128 and 256 grids, respectively. The number of photon packages for the scattering Monte Carlo simulation is 109. In this paper, we consider three near-infrared wavelengths: λ = 1.1 µm, 1.6 µm, and 2.2 µm. 2.2 Dust models and methods To study how size and internal structure of dust particles affect observational appearance of discs, we consider three types of dust models. In this paper, we basically use the term large or small dust radius when the radius is large or small compared to λ/(2π), where λ is a wavelength of interest. • Single monomer model: 0.1 µm-sized homogeneous spherical grain (= a monomer particle). This particle is used to form porous dust aggregates described below. Optical properties of the monomer are obtained by using the Mie theory (Bohren & Huffman 1983). The monomer radius as- sumed is roughly the same as the constituent particle of cometary dust aggregates (e.g., Kimura et al. 2006). Since we assume that a monomer is homogeneous spherical grain, volume filling factor of the monomer is unity. • Porous dust aggregate model: We consider two types of fractal dust aggregates, ballistic cluster cluster agglomerate (BCCA) and ballistic particle cluster agglomerate (BPCA) (see Figure 1). Fractal dust aggregates obey a relation N = k0(Rg/R0)d f , where N is the number of monomers, Rg is the radius of gyration of the aggregate, R0 is the radius of a monomer, k0 is the fractal prefactor, and df is the fractal dimension. Typically, BCCA and BPCA have fractal dimen- sion and fractal prefactor df = 1.9 and k0 = 1.04 and df = 3.0 and k0 = 0.3, respectively (Paper I). It is convenient to de- fine the characteristic radius by Rc =(cid:112)5/3Rg (Mukai et al. 0/R3 1992) 1. In this paper, volume filling factor of the aggre- gate is defined by f = N R3 c (Mukai et al. 1992) 2. In addition, it is also useful to define volume-equivalent radius RV ≡ R0N1/3 = Rc f 1/3. Porosity of dust aggregates can be defined by P = 1 − f . If the volume filling factor is unity, the characteristic radius is equal to the volume equivalent radius. Optical properties of the BCCA and BPCA models are obtained by using a rigorous numerical method, T-Matrix Method 3 (TMM; Mackowski & Mishchenko 1996) with the Quasi-Monte Carlo orientation averaging method (Okada 2008)4. We also use the modified mean field theory (MMF) developed in Paper II5. Rg =(cid:112)3/5a. Thus, characteristic radius of porous aggregate with 1 The solid sphere of the radius a has the radius of gyration the radius of gyration Rg is defined in this way. 2 A way to define filling factor is not unique, and the choice may somewhat affect porosity values, in particular for BCCA. Our definition of filling factor has been widely used (e.g., Mukai et al. 1992), and hence, in this paper, we follow this conventional definition. 3 The code is available on ftp://ftp.eng.auburn.edu/pub/ dmckwski/scatcodes/. A newer version of this code is available at http://www.eng.auburn.edu/~dmckwski/scatcodes/. 4 More detailed information about TMM calculations of porous dust aggregate model is available in Paper I and Paper II. 5 The Rayleigh -- Gans -- Debye theory studied in Paper I is a single scattering theory, whereas multiple scattering is often important for relatively large compact aggregates with d f > 2. A simple way to implement multiple scattering is to adopt the mean field ap- proximation (Berry & Percival 1986; Botet et al. 1997), although the mean field approximation fails to predict the single scatter- ing albedo, once multiple scattering inside the aggregate becomes dominant (Paper II). Paper II proposed the modified mean field theory in which inaccurate behaviours in the mean field theory are MNRAS 000, 1 -- 17 (2019) Scattered-light images of fluffy dust discs 3 Figure 1. Morphology of porous dust aggregates. Left and right panels correspond to the BCCA and BPCA models, respectively. The number of monomers is 1024 and the monomer radius is set as R0 = 0.1 µm, and hence the characteristic radii of the BCCA and BPCA models are Rc = 4.8 µm and 1.9 µm, respectively. Both dust aggregates have the same volume equivalent radius RV (cid:39) 1 µm. Hence, BCCA and BPCA have porosity of 99% and 85%, respectively. We are mainly interested in porous dust aggregates whose radii are larger than λ/(2π) because small aggregates simply give rise to Rayleigh scattering. For TMM computation, we adopt N = 1024. Although N = 1024 is not a large num- ber, for R0 = 0.1 µm, the characteristic radii of the BCCA and BPCA models are Rc = 4.8 µm and 1.9 µm, respec- tively; thus, the characteristic radii exceed λ/(2π) when λ corresponds to near-infrared wavelengths. Porosity of the BCCA and BPCA models is 99% and 85%, respectively. For further large aggregates, TMM computation becomes time-consuming, and hence we use MMF instead of TMM. Comparison between MMF and TMM is available in Paper I, Paper II, and Appendix A. • Compact dust aggregate model: Single-sized compact dust aggregate, whose porosity is low enough so that the distribution of hollow spheres6 (DHS; Min et al. 2005, 2016) can be applicable outside the Rayleigh domain. We adopt the irregularity parameter fmax = 0.8 (Min et al. 2016). The volume equivalent radius of a compact dust aggregate is var- ied from 0.1 µm to 5.0 µm. In the DHS method, volume filling factor of an aggregate is not necessary to be specified. Among three dust models, we adopt astronomical silicate for the optical constant (Draine & Lee 1984; Laor & Draine 1993). In this case, opacities and scattering matrix elements of the BCCA and BPCA models have already presented in Paper I and Paper II. We discuss the effect of dust compo- sition (Section 4.3), and monomer radius and fractal dimen- sion (Section 5) in more detail. 2.3 Optical properties of out dust models We compute optical properties of dust models given in Sec- tion 2.2. Figure 2 shows dust optical properties at wave- improved by employing geometrical optics approximation (e.g., Bohren & Huffman 1983). 6 In order to avoid strong resonances appeared in scattering prop- erties, we prefer to use the DHS method rather than the Mie theory for compact particles larger than λ/(2π). MNRAS 000, 1 -- 17 (2019) length λ = 1.6 µm for the single monomer model, the porous dust aggregate model (the BCCA and BPCA models with N = 1024 and R0 = 0.1 µm), and the compact dust ag- gregate model (RV = 1 µm). Since BCCA and BPCA has N = 1024 and R0 = 0.1 µm, their volume equivalent radii are RV = R0N1/3 (cid:39) 1.0 µm. Therefore, BCCA, BPCA and compact dust aggregates have the same volume equivalent radii. In this paper, scattering matrix elements are defined by Zi j = Si j/(k2mdust), where Si j is those defined by Bohren & Huffman (1983), k is the wavenumber, and mdust is the mass of dust particle. Since Z11 is a differential cross section per unit mass, it is related to the scattering opacity κsca via Þ κsca = Z11dΩ, (1) where dΩ is the solid angle. For the sake of simplicity, in this paper, we refer Z11 as phase function. As shown in Fig- ure 2, scattering is anisotropic for BCCA, BPCA, and com- pact dust aggregate models. This is because these aggregates have the radii larger than λ/(2π). On the other hand, the monomer model shows isotropic scattering due to Rayleigh scattering. The degree of linear polarisation is given by −Z12/Z11. For small particles, such as a monomer particle, the degree of polarisation is 100% at 90 degree scattering angle due to Rayleigh scattering. On the other hand, for a particle larger than λ/(2π), the degree of polarisation depends on the struc- ture of dust aggregates. Porous dust aggregates tend to show high degree of polarisation with a bell-shaped profile (Fig- ure 2). This is because multiple scattering inside the aggre- gate is suppressed due to its highly porous structure, and then, polarisation properties are determined by those of the monomer (see Eq. (9) in Paper I). The degree of polarisa- tion of compact dust aggregates is lower than porous dust aggregates due to occurrence of multiple scattering inside the particle. Hence, the degree of polarisation provides valu- able information about structure of dust aggregates when the particle radius is larger than wavelength. 4 R. Tazaki et al. Figure 2. Phase function Z11 (left), and the degree of linear polarisation −Z12/Z11 (right). Optical properties of the single monomer model, the BCCA and BPCA models, and the compact dust aggregate model are obtained by using the Mie theory, TMM, and DHS, respectively. A single monomer has the radius of 0.1 µm, whereas the porous (BCCAs and BPCAs) and compact dust aggregate models have the same volume equivalent radii 1.0 µm. Wavelength is set as λ = 1.6 µm. Hatched region indicates a range of scattering angle to be observed for a disc with the flaring index β = 1.25 and the inclination angle i = 60◦. 3 RESULTS OF RADIATIVE TRANSFER SIMULATIONS: IMAGES Table 1. Disc Integrated Polarisation Fraction at λ = 1.6 µm for the our Dust Models We perform radiative transfer simulations using a model de- scribed in Section 2. Optical properties used in the simula- tions are the ones summarised in Section 2.3. Total intensity images (middle panels of Figure 3) can be used as a diagnostics of particle radius. Both the porous and compact aggregate models show the brightness asymme- try, that is, the near side of the disc is brighter than the back- ward side. This is because the dust particle radii assumed are larger than the wavelength (radius (cid:38) λ/2π), and hence strong forward scattering occurs. Meanwhile, for the single monomer model, the brightness asymmetry is weak because the monomer is a Rayleigh scatterer, that is, isotropic scat- tering occurs. Although differences between porous aggregates (BCCA and BPCA) and compact aggregates are less obvious in total intensity images, they clearly differ in images of polarisation fraction. Figure 4 shows polarisation fraction measured at R = 50 au from Figure 3, where scattering angles are estimated by using a scattered-light mapping method (Stolker et al. 2016). As shown in Figure 4, the polarisation fraction of the BCCA and BPCA models is as high as 65% - 75%. Polarisation fraction obtained by simulations is somewhat smaller than the degree of polarisation (Figure 2). This is mainly due to occurrence of multiple scattering at the disc surface because the disk is optically thick. When we compare Figures 2 and 4, it is found that the degree of polarisation is similar between the monomer, the BCCA and BPCA models, whereas highest polarisation fraction observed in the discs for these particles can largely differ. Polarisation fraction of the compact dust aggregate model can be much smaller than those of porous dust aggre- gate models. This is because multiple scattering easily oc- curs for more compact structure of dust aggregates once the Inclination angle monomer BCCA BPCA Compact i = 0◦ i = 15◦ i = 30◦ i = 45◦ i = 60◦ i = 75◦ 84% 80% 68% 53% 42% 37% 63% 57% 46% 32% 18% 14% 46% 42% 31% 19% 8% 8% 22% 19% 21% 13% 8% 6% aggregate radius becomes larger than λ/2π 7. As a result, compact dust aggregates show lower polarisation fraction compared to the porous dust aggregates. Table 1 summarises integrated polarisation fraction of discs for various dust models. For all dust models, more in- clined discs show lower polarisation fraction because wider scattering angles can be observed. The BCCA model pre- dicts that integrated polarisation fraction is as high as 63% for a face-on disc, whereas it is only 14% for an very inclined disc (i = 75◦). On the other hand, more compact dust struc- ture, like BPCA and compact dust models, results in smaller integrated polarisation fraction than those of BCCA. As a result, it is found that disc polarisation fraction is an impor- tant quantity to distinguish different dust structure, where relatively high polarisation fraction of BCCA and monomer could be distinguished by the presence of forward scattered light in total intensity image. The tendency that porous aggregates show high polari- sation fraction is compatible with the previous works by Min et al. (2012), who used EMT, and by Kirchschlager & Wolf 7 When the particle has refractive index close to unity, polari- sation fraction remains high even if the size parameter exceeds unity (see e.g., chapter 6 of Bohren & Huffman 1983). This is due to multiple scattering inside the sphere becomes sub-dominant for such a transparent material. MNRAS 000, 1 -- 17 (2019) 100101102103104105020406080100120140160180Z11[(cm2/g)/str]Scatteringangle[◦]MonomerBCCABPCACompact100101102103104105020406080100120140160180−1−0.500.51020406080100120140160180−Z12/Z11Scatteringangle[◦]MonomerBCCABPCACompact−1−0.500.51020406080100120140160180 Scattered-light images of fluffy dust discs 5 Figure 3. Polarised intensity (top), total intensity (middle), and polarisation fraction (bottom). From left to right, each panel indicates the results for the single monomer model (0.1 µm-sized spherical particles), the BCCA and BPCA models (N = 1024, R0 = 0.1 µm, astronomical silicate), and the compact dust aggregate model with the radius of 1.0 µm. The wavelength is λ = 1.6 µm and the inclination angle is i = 60◦. Optical properties of BCCAs and BPCAs are calculated by a rigorous method, TMM. No star is included in images (perfect coronagraph). For the porous dust aggregate models, the disc shows brightness intensity asymmetry (in total intensity) due to forward scattering as well as high polarisation fraction. (2014), who adopted less porous particle models. However, as shown in Appendix A, EMT largely under/over-estimate the scattered-light intensity and the polarisation fraction for large fluffy dust aggregates. This is due to the fact that phase function obtained by EMT shows extremely strong forward scattering, which suppresses both large angle scattering (re- ducing total intensity) and multiple scattering at disc surface (increasing polarisation fraction). Polarised intensity does not show clear brightness asym- metry because polarisation fraction becomes small along the minor axis. The compact dust aggregate model shows some slits in the image of polarised intensity. In our com- pact dust aggregate model, oscillatory behaviour appears in the angular profile of the degree of polarisation due to res- onances arising from smooth spherical surfaces (see Figure 2). However, realistic compact dust aggregates are thought to have surface roughness, and hence these resonances would be smeared out (e.g., Min et al. 2016). Therefore, these MNRAS 000, 1 -- 17 (2019) slits in the image thought to be unrealistic. The reason why a monomer particle, which also has smooth surface, does not show the oscillatory behaviour is simply because Rayleigh scattering occurs, that is, scattering is coherent. It is also worth mentioning that even if the oscillatory be- haviour is suppressed, polarisation flip at disk backward side may remain for some dust properties because of the effect called negative polarisation branch (e.g., Kirchschlager & Wolf 2014). This effect, for example, has been observed for cometary dust particles (e.g., Kolokolova et al. 2007). 4 RESULTS OF RADIATIVE TRANSFER SIMULATIONS: SCATTERED-LIGHT COLOURS We define the scattered light colour by log(Ldisc,λ2/Lstar,λ2) − log(Ldisc,λ1/Lstar,λ1) η = log(λ2/λ1) , (2) 6 R. Tazaki et al. Figure 4. Polarisation fraction as a function of scattering angles at R = 50 au measured from simulated images in Figure 3. Because of multiple scattering at the disc surface, polarisation fraction is smaller than the degree of polarisation given in Figure 2 (right). Table 2. Disc Scattered Light Colour for Different Aggregate Models Dust model Monomer BCCA (rV = 1.0 µm, Rc = 4.8 µm) BPCA (rV = 1.0 µm, Rc = 1.9 µm) Compact (RV = 1.0 µm) Compact (RV = 1.9 µm) Compact (RV = 4.8 µm) ηI −2.4 −0.51 −0.21 0.030 0.77 0.68 ηPI −2.4 −0.41 −0.27 0.099 0.91 0.053 where Ldisc,λ and Lstar,λ are total luminosity of disc and cen- tral star, respectively, for either total intensity or polarised intensity. We adopt λ1 = 1.1 µm and λ2 = 2.2 µm. In this paper, we classify scattered-light colours into blue, grey, and red when η < −0.5, −0.5 ≤ η ≤ 0.5, and η > 0.5, respectively. 4.1 Comapct aggregates vs. porous dust aggregates Figure 5 and Table 2 show scattered-light colour for both in total intensity (ηI) and in polarised intensity (ηPI) at incli- nation angle i = 60◦. Figure 5 also compares total intensity colours of BCCA (Rc = 4.8 µm) to those of the compact aggregate with the same radius. We find that total intensity colours of BCCA (Rc = 4.8 µm) are grey or slightly blue for both total and polarised intensity. Meanwhile, a compact aggregate with RV = 4.8 µm shows reddish colour (Mulders et al. 2013). BPCA (Rc = 1.9 µm) also show grey colours in total intensity, whereas compact aggregates with 1.9 µm show reddish colours. As a result, porous aggregates large compared to wavelength tends to show grey or slightly blue colours in total intensity, whereas large compact aggregates give rise to reddish colours. Therefore, even if porous and compact dust aggregates have the same radii, their scat- tered light colour can differ. It is worth mentioning that for compact dust aggregates, grey colours can appear when the radius is comparable to the wavelength (see Figure B4 in Appendix B for more detail). 4.2 Scattered light colours of millimetre-sized BCCA Due to strong aerodynamic coupling, highly porous dust ag- gregates at disc surface may be much larger than micron- size, whereas large compact dust aggregates are likely to settle down to the midplane. In Section 4.1, we concluded that a few micron-sized BCCA are shown to yield grey or slightly blue rather than reddish colours. However, one may doubt that the presence of further larger fluffy aggregates than those considered in Section 4.1 may makes the disc reddish. In this section, we show that colours remain almost the same even if the radius of aggregates is increased to millimetre-size. 4.2.1 Phase function Since we are interested in aerodynamically coupled aggre- gates, we consider BCCA, whose mass-to-area ratio is the same as the individual monomer particle. In particular, we study scattering properties of BCCA with radius from Rc = 1 µm to 1 mm. Millimetre-sized BCCA with R0 = 0.1 µm contains ∼ 107 monomers, and hence, TMM computation is time consuming. Thus, we adopt MMF instead of TMM. For MMF, angular dependence of phase matrix elements are reliably computed when the single scattering assumption is validated (Paper II): ∆φ < 1, m − 1 < 2, (3) (4) where ∆φ is the (maximum) phase shift of dust aggregates and m is the complex refractive index. For astronomical sili- cate, the first condition is satisfied for λ ≥ 0.85 µm for BCCA with R0 = 0.1 µm. It should be emphasised that in the case of BCCA, ∆φ does not depend on the aggregate radius. In optical and near-infrared wavelengths, the second condition is also satisfied (see Figure 3 in Paper I). Figure 6 shows phase function Z11 of BCCA with ra- dius from Rc = 1 µm to 1 mm obtained by MMF. As the aggregate radius increases, forward scattering becomes strong; however, intermediate- and backward-scattered in- tensity saturates. This saturation is a natural consequence of single scattering by dust aggregates (Berry & Percival 1986; Paper I). Mechanism of the saturation is illustrated in Fig- ure 7. For large angle scattering (θ (cid:38) (2πRg/λ)−1), scattered intensity is dominated by coherent scattered waves from a pair of monomers separated by the distance smaller than the aggregate radius. On the other hand, a pair of monomers with separation as large as aggregate radius only gives rise to incoherent contribution to the scattered intensity because their optical path of difference is large. Thus, scattered inten- sity at large scattering angles is governed by its small scale structure, e.g., fractal dimension and the monomer's prop- erties, rather than the aggregate radius. As a result, phase function Z11 shown in Figure 6 is insensitive to the charac- teristic radius at intermediate- and back-scattering angles. For the case of small angle scattering (θ (cid:46) (2πRg/λ)−1), large scale structure of aggregates can produce coherent scattered MNRAS 000, 1 -- 17 (2019) 00.10.20.30.40.50.60.70.80.91020406080100120140160180i=60◦PolarizationfractionScatteringangle[deg]MonomerBCCABPCACompact Scattered-light images of fluffy dust discs 7 Figure 5. Fractional luminosity in total intensity (left) and polarised intensity (right) at the inclination angle i = 60◦. Black and red solid lines represents the results for BCCA and BPCA, respectively. Black and red dashed lines indicate the results for compact dust aggregates with 4.8 µm and 1.9 µm radii, respectively. Figure 6. Phase function Z11 of extremely large BCCA obtained by MMF. The wavelength is set as λ = 1.6 µm. Hatched region indicates a range of scattering angle to be observed for a disc with the flaring index β = 1.25 and the inclination angle i = 60◦. light, and therefore, Z11 depends on the aggregate radius as can be seen in Figure 6. 4.2.2 Disc scattered-light colours By using optical properties of BCCA computed in Section 4.2.1, we perform radiative transfer simulations of discs. As shown in Figure 6, millimetre-sized BCCAs show very sharp and intense forward scattering. Hence, a large number of grids will be required in the radiative transfer calculation8. In addition, sharp increase of phase function makes opac- ity integration inaccurate. In order to avoid these problems, we place upper limit on peaking forward scattering, since 8 See also Section 6.5.6 in Manual for radmc-3d Version 0.39, which is available on http://www.ita.uni-heidelberg.de/ ~dullemond/software/radmc-3d/download.html. MNRAS 000, 1 -- 17 (2019) Figure 7. Schematic illustration to explain saturation of Z11 for large angle scattering. At large scattering angles, scattered light is dominated by coherent light scattered from small scale structure of the aggregate, whereas large scale structure, e.g., aggregate radius, only produces incoherent light due to relatively large op- tical path of difference. Hence, the aggregate radius is insensitive to Z11 at intermediate scattering angles. forward scattering outside the observable angle range is not likely to affect observed images. In this paper, we adopt Z11(θ < θc) = Z11(θc), where θc is taken to be 1 degree. Figure 8 shows results of radiative transfer simulations of discs containing BCCA with Rc = 1 µm to 1 mm (Ob- tained images are shown in Figure A5.). It is found that the disc scattered light is insensitive to the characteristic radius of aggregate. This is because scattered light intensity in a range of observable scattering angles does not depend on the characteristic radius (see hatched range of scattering angles in Figure 6). Table 3 shows that colours become shallower as the characteristic radius increases; however, even if the aggregate radius is increased up to millimetre-size, colours do not reach the reddish colour regime. Colours of BCCA with R0 = 0.1µm seem to be slightly blue, but not too blue like Rayleigh scattering particles (see Table 2). Therefore, even if the disc contains BCCA as large as millimetre-size 10−210−110011.21.41.61.822.2LIdisk,λ/Lstar,λλ(µm)BCCA(Rc=4.8µm)Compact(RV=4.8µm)BPCA(Rc=1.9µm)Compact(RV=1.9µm)10−310−210−111.21.41.61.822.2LPIdisk,λ/Lstar,λλ(µm)BCCA(Rc=4.8µm)Compact(RV=4.8µm)BPCA(Rc=1.9µm)Compact(RV=1.9µm)101102103104105106020406080100120140160180Z11[(cm2/g)/str]Scatteringangle[◦]BCCA(Rc=1mm)BCCA(Rc=100µm)BCCA(Rc=10µm)BCCA(Rc=1µm)101102103104105106020406080100120140160180 8 R. Tazaki et al. Table 3. Disc Scattered Light Colour for extremely large BCCA Table 4. Disc Scattered Light Colour for Different Disc and Dust Models Aggregate Radius Rc = 1 µm (MMF) Rc = 10 µm (MMF) Rc = 100 µm (MMF) Rc = 1 mm (MMF) Rc = 1 mm, R0 = 0.01 µm (MMF) Rc = 4.8 µm (TMM) Rc = 4.8 µm (MMF) ηI −1.0 −0.69 −0.67 −0.58 −1.2 −0.51 −0.67 ηPI −0.64 −0.68 −0.66 −0.58 −1.1 −0.41 −0.67 at the surface, reddish colour scattered light is not likely to occur. In Figure 6, we also investigate how the monomer ra- dius affects disc scattered lights. In order to satisfy Equation (3), we consider millimetre-sized BCCA with R0 = 0.01 µm- sized monomers. If the monomer radius is decreased, the disc becomes faint due to small albedo value of dust aggre- gates. Large BCCA of small monomers show bluer colours than those of R0 = 0.1 µm monomers. Therefore, scattered light colours of BCCA depend on the monomer radius (see Section 5 for more detail). Finally, we mention error of scattered light colours due to the use of MMF (see also Appendix A). MMF results show slightly blue colours for large BCCA; however, MMF tends to produce bluer colours compared to rigorous TMM calcu- lations (see Table 3). For Rc = 4.8 µm, the difference of the slope is ∆I ≡ ηI,TMM−ηI,MMF (cid:39) 0.16, where ηI,TMM and ηI,MMF are total intensity colours obtained by TMM and MMF, re- spectively. Similarly, we can define the difference of the slope for polarised intensity, ∆PI ≡ ηPI,TMM−ηPI,MMF (cid:39) 0.26. Hence, TMM results are shallower than MMF results (see Figure A4). These difference are mainly due to approach of ∆φ to unity, which makes MMF inaccurate due to occurrence of multiple scattering. ∆φ increases as the wavelength de- creases; however, for sufficiently large BCCA (e.g., df ≈ 2), ∆φ does not depends on the aggregate radius and depends only on the property of the monomer (refractive index and size parameter) (Berry & Percival 1986; Paper II). Hence, it is reasonable to assume that errors in slope for Rc = 10 µm, 100 µm, and 1 mm are similar to those of Rc = 4.8µm, and therefore, slope of these large BCCA may be shallower than those given in Table 3 by ∆I = 0.16 and ∆PI = 0.26 for total intensity and polarised intensity, respectively. As a result, by considering errors in colours between TMM and MMF, millimetre-sized BCCA with R0 = 0.1 µm are thought not to produce reddish colours at near-infrared wavelengths and more likely to produce grey colours. 4.3 Effect of disc geometry, inclination, and dust composition In Sections 4.1 and 4.2, we studied scattered-light colour by assuming a single disc geometry, inclination angle, and a single dust composition. Here, we address how these param- eters affect scattered light colours. We vary following parameters: flaring index β, disc in- ner and outer truncation radius Rin and Rout, respectively, inclination angle i, and dust composition. Total dust mass is kept the same, that is 10−4M(cid:12). As a dust composition, Dust model ηI −0.51 fiducial i = 15◦ −0.80 −0.43 β = 1.1 −0.61 Rout = 300 au −0.55 Rin = 0.1 au Carbon-rich BCCA −0.60 ηPI −0.41 −0.46 −0.30 −0.47 −0.47 −0.66 we consider the carbon-rich composition, which is a mixture of silicate, iron sulphide, water ice, and amorphous carbon. Fraction of each ingredients are determined by a recipe given by Min et al. (2011) with the carbon partition parameter w = 1. Optical constants of silicate, iron sulphide, and wa- ter ice are taken from Henning & Stognienko (1996) and for amorphous carbon from Zubko et al. (1996). These optical constants are mixed by using the Bruggeman mixing rule (Bruggeman 1935). Optical properties of carbon-rich BCCA (Rc = 4.8 µm and R0 = 0.1 µm) are obtained by using MMF. In Figure 9, we show scattered-light colour of the BCCA model for various parameters. Table 4 summarises scattered- light colours. It is shown that that these parameters affect mainly magnitude of fractal luminosity. Although colours can differ for different disc geometries and dust composition, within a parameter range we studied, these parameters do not significantly change the our results given in Sections 4.1 and 4.2. It should be mentioned that fractional luminosity in to- tal intensity and polarised intensity derived in this paper is higher than the observed values (Fukagawa et al. 2010; Avenhaus et al. 2018). However, as shown in Figure 9, disc scattered light luminosity is sensitive to the disc structure and dust composition, although the colours are less sensitive to them. Hence, we expect observed fractional luminosity in total intensity and polarised intensity might be explained by including these effects, though we need to check it for each object. This is beyond the scope of this paper. 5 EFFECT OF DUST AGGREGATE STRUCTURE ON SCATTERING PROPERTY In Section 4, it is shown that large porous dust aggregates show slightly blue or grey colours in scattered light, while large compact aggregates tends to show reddish colours. In this section, we discuss reasons for this by considering in- trinsic optical properties of dust aggregates. We define the effective albedo by ωeff = κeff sca/(κabs + κeff sca), sca = (1− g)κsca is the effective scattering opacity, κabs where κeff is the absorption opacity, and g is the asymmetry parameter. Since small angle scattered light is hardly observable, the effective albedo is a useful quantities (see also Dullemond & Natta 2003; Min et al. 2010; Mulders et al. 2013). In the case of efficient forward scattering (g ∼ 1), only small amount of incident photons will be scattered toward the observer (supposed to be at θ (cid:44) 0); hence, the effective albedo has small values. On the other hand, if scattering is isotropic (g ∼ 0), the effective albedo gives rise to the single scattering MNRAS 000, 1 -- 17 (2019) Scattered-light images of fluffy dust discs 9 Figure 8. Fractional luminosity of the disc for total intensity (left) and polarised intensity (right). [t] Figure 9. Fractional luminosity for total intensity for various disc and dust composition, where the dust morphology is the BCCA model (Rc = 4.8 µm and R0 = 0.1 µm). The fiducial model has the inclination angle i = 60◦, the flaring index β = 1.25, the inner and outer radii Rin = 10 au and Rout = 100 au, respectively, and the dust composition is astronomical silicate (plus symbol). Each parameter is changed one-by-one from the fiducial model; the fiducial model, but for i = 15◦ (cross), for β = 1.1 (circle), for Rout = 300 au (square), for Rin = 0.1 au (diamond), and for carbon-rich BCCA (triangle). albedo ω. Thus, the effective albedo might be used as a qualitative measure of disc scattered light colours. In order to obtain the effective albedo of dust aggre- gates, we use MMF developed by Paper II. In MMF, the structure of dust aggregates is specified in terms of two-point correlation function of monomers (Berry & Percival 1986; Botet et al. 1997; Paper II). In addition, by using of mean field assumption, multiple scattering inside the aggregate is solved in a self-consistent manner. Hence, it is suitable to study how fractal dimension affect optical cross sections. Figure 10 shows the effective albedo for dust aggre- MNRAS 000, 1 -- 17 (2019) gates with various fractal dimension and monomer radii. It is found that when the characteristic radius is fixed to 2.5 µm, with increasing fractal dimension from 2 to 3, colours of the effective albedo vary from blue to red at near-infrared wavelengths. This can be explained by as follows. For large compact dust aggregates, as wavelength decreases, forward scattering becomes more efficient, that is, g increases. Thus, the effec- tive albedo decreases for short wavelength domain (Mulders et al. 2013). Hence, large compact aggregates shows red- dish colours. Meanwhile, for large highly porous aggregates, asymmetry parameter g becomes almost constant with re- spect to the wavelength due to saturation of scattered light (see Appendix C), and therefore, anisotropic scattering does not reduce the effective albedo. As a result, porous aggre- gates tends to show more bluer colours compared to compact aggregates. Scattered light colours depend also on the monomer ra- dius, in particular for df = 2 (like BCCA). In right panel of Figure 10, we show the effective albedo of fluffy dust aggregates (k0 = 1.0 and df = 2.0) with various monomer radii, where the complex refractive indices is set as m = 1.67 +0.0326i for the sake of simplicity (corresponding to as- tronomical silicate at λ = 1.6 µm). It is found that when the monomer radius is much smaller that the NIR wavelength (cases of R0 = 1 nm and 10 nm), NIR slope of the effective albedo is blue (see also Table 3). On the other hand, for the case of R0 = 1 µm, the single scattering assumption is violated, and hence, the effective albedo becomes red due to occurrence of multiple scattering inside the cluster. Thus, the monomer radius is sensitive to scattered light colours of large fluffy dust aggregates. There are three possibilities to explain observed grey scattered light colours. One possibility is moderately com- pressed aggregates. As shown in Figure 10, scattered light colours of aggregates with df ≈ 2.5 will be almost grey. Nu- merical simulations have shown that mutual aggregate colli- sions can only produces an aggregate with fractal dimension df ≈ 2.5 (Wada et al. 2008; Suyama et al. 2008). Hence, collision compressed aggregates may produce grey scattered light colours at near-infrared wavelength. Another possibil- ity is that the aggregate with df ≈ 2 has the monomer ra- 10−210−110011.21.41.61.822.2LIdisk,λ/Lstar,λλ(µm)Rc=1µmRc=10µmRc=100µmRc=1mmRc=1mm,R0=0.01µm10−310−210−111.21.41.61.822.2LPIdisk,λ/Lstar,λλ(µm)Rc=1µmRc=10µmRc=100µmRc=1mmRc=1mm,R0=0.01µm10−210−110011.21.41.61.822.2LIdisk,λ/Lstar,λλ(µm)fiduciali=15◦β=1.1Rout=300auRin=0.1auCarbon-richBCCA 10 R. Tazaki et al. Figure 10. (Left) The effective albedo for different values of d f and k0, where the characteristic radius and monomer radii are Rc = 2.5 µm and R0 = 0.1 µm, respectively. Refractive indices is astronomical silicate. (Right) The effective albedo of fluffy dust aggregates (k0 = 1.0 and d f = 2) of various monomer radii. The characteristic radius is assumed to be 2.5 µm. Refractive indices is set as m = 1.67 + 0.0326i, which corresponds to those of astronomical silicate at λ = 1.6 µm. dius slightly larger than 0.1 µm but not as large as micron size. This can also produce grey colours. The third possibil- ity is the conventional one, that is, compact particles with the radius comparable to the observing wavelength (Fig- ure B4). Although composite mixture of compact particles and porous aggregates in the same disk may also affect disk colours, this is left for a future task. 6 COMPARISON WITH DISC OBSERVATIONS Near-infrared observations of protoplanetary discs have shown that most discs show grey colours in both total and polarised intensity (Fukagawa et al. 2010; Avenhaus et al. 2018). Previously, these scattered-light properties are inter- preted as the presence of compact grains with the radius comparable to the observing wavelength (see Figure B4). In this case, it is necessary that dust grain radii at disc surface is adjusted to observing wavelength. One adjusting mech- anism could be dust vertical stratification (Duchene et al. 2004; Pinte et al. 2007). This study provides a new insight into interpretation of discs showing grey colours. We found that large porous dust aggregates can also show marginally grey or slightly blue total intensity. One advantage is that in the case of porous aggregates, their radius is not neces- sary to adjust to observing wavelength. In addition, large porous dust aggregates are expected to be stirred higher al- titude of the discs, they are likely to affect disc scattered light. Therefore, grey discs in total intensity might by ex- plained by large fluffy dust aggregates, although more detail modelling is necessary for each object. If a disc contains porous dust aggregates, we predict that polarisation fraction of disk scattered light should be high. For example, as shown in Figure 4, porous aggregates with sub-micron monomers produces polarisation fraction as high as 65%-75% and disc integrated polarisation fraction is about 18% at λ = 1.6 µm at inclination angle i = 60◦ (see also Table 1). Spatially resolved polarisation fraction map is obtained for several discs, and these observations have revealed that disc scattered light in near-infrared wavelengths is often highly polarised. Polarisation fraction of GG Tau and AB Aur discs show polarisation fraction as high as 50% at λ = 1 µm for GG Tau (Silber et al. 2000) and λ = 2 µm for AB Aur (Perrin et al. 2009). More recently, Itoh et al. (2014) reported polarisation fraction of GG Tau is as high as that of Rayleigh scattering function at H-band; nevertheless, strong forward scattering is also observed at H-band. For HD 142527, Canovas et al. (2013) derived polarisation fraction of 10% -- 25% at H-band; however, Avenhaus et al. (2014) re- ported significantly higher polarisation fraction for this ob- ject (20% -- 45%) at H-band. Tanii et al. (2012) show that polarisation fraction of UX Tau is up to 66% at H-band. Fur- thermore, Poteet et al. (2018) have shown that polarisation fraction of TW Hya is as high as 63% ± 9%. Thus, scattered light of these spatially resolved objects is highly polarised, and hence these scattered light might be explained by porous dust aggregates models, although both detail modelling for each object and further observational studies to derive polar- isation fraction are necessary. It should be emphasised that polarisation fraction predicted by our porous dust aggregate model is not very high compared to disc observations. Mulders et al. (2013) pointed out that some protoplane- tary discs show reddish scattered-light colour in total inten- sity. For the case of the outer disc of HD 100546, reddish and faint scattered light are observed, and it can be explained by 2.5 µm sized spherical particles (Mulders et al. 2013). Based on the SED modelling of HD 100546, Mulders et al. (2013) also claimed that the disc scale height is consistent with a particle model of 0.1 µm rather than 2.5 µm for a given turbulent strength and disc gas mass. They suggested a possibility that the presence of extremely fluffy dust ag- gregates containing submicron-sized monomers at the disc surface layer may reconcile this apparent conflict, since such dust aggregate behaves like a small particle in dynamics and like a large particle in light scattering process. However, we found that fluffy aggregates of submicron-sized monomers do not show reddish colour in total intensity. Reddish scattered light is more likely to be observed when the dust aggregates MNRAS 000, 1 -- 17 (2019) 10−310−210−1100110100Rc=2.5µm,R0=0.1µm,astrosilωeffλ(µm)df=2.0,k0=1.0(P=97.6%)df=2.2,k0=1.0(P=95.7%)df=2.4,k0=1.0(P=92.2%)df=2.6,k0=1.0(P=85.8%)df=2.8,k0=1.0(P=74.3%)df=3.0,k0=1.0(P=53.5%)10−310−210−1100110100Rc=2.5µmωeffλ(µm)R0=1.0µm(P=76.0%)R0=0.1µm(P=97.6%)R0=0.01µm(P=99.8%)R0=0.001µm(P=99.98%) have compact structure (see Figures B4 and 10). Therefore, our results imply that fluffy dust aggregates are not respon- sible for the scattered-light properties of the outer disk of HD 100546. Our results predict that polarisation fraction of HD 100546 should not be high. Quanz et al. (2011) show in- tegrated polarisation fraction of this disc is about 14%. Since average inclination angle of HD 100546 is 46◦ (Mulders et al. 2013), observed polarisation fraction is lower than the prediction of our BCCA model, which shows about 32% at i = 45◦ (Table 1). This observed polarisation fraction is more close to our compact dust aggregate model. Recently, Stolker et al. (2016) derived scattering phase function of this object. The derived phase function increases with increasing scatter- ing angles from intermediate to back scattering angles. This enhanced backward scattering can be seen in phase function of compact dust aggregates (Min et al. 2016), and not for fluffy dust aggregate (Paper I, see also Figure 2). Laboratory experiments also support that large compact particles show the enhanced backscattering (Munoz et al. 2017). These en- hanced backscattering presumably due to the presence of wavelength scale surface roughness of a large compact par- ticle (e.g., Mukai et al. 1982). Thus, phase function of this object is consistent with our implication. 7 CONCLUSION We have studied how radius and structure of dust aggre- gates affect observational quantities of protoplanetary discs in near-infrared wavelengths. We have performed radiative transfer calculations of protoplanetary discs taking fluffy dust aggregates into account, where we firstly treated their optical properties in a proper manner. Furthermore, based on an approximate theory for optical properties of fractal dust aggregates (MMF) (Paper I; Paper II), we have argued scattering properties of fractal dust aggregates. Our primary findings are summarised as follows. • The ratio of a aggregate radius and wavelength can be assessed by the presence of brightness asymmetry in total intensity images. Both porous and compact dust aggregates can produce brightness asymmetry in total intensity when their radii exceed the observing wavelength (Sections 2.3 and 3).• Polarisation fraction of a disc is useful to probe struc- ture of dust aggregates as long as the aggregate radius ex- ceeds observing wavelength (Sections 2.3 and 3). Higher porosity produces higher polarisation fraction. We have pro- vided expected integrated polarisation fraction for various dust models and disk inclinations (Table 1). • Porous dust aggregates (BCCAs and BPCAs) large compared to near-infrared wavelengths show marginally grey or slightly blue in total or polarised intensity. Meanwhile, large compact dust aggregates show reddish scattered light colour in total intensity (Section 4.1). • For sufficiently large BCCA, the aggregate radius is less sensitive to disc scattered light because of saturation of scat- tering property (Figures 6 and 7). Even if the radius is in- creased up to millimetre-size, BCCA containing 0.1 µm-sized monomers is expected to show marginally grey or slightly blue colours (Section 4.2). • Within a parameter range we studied, the disc geome- try, inclination angle, and composition seems not to be sensi- MNRAS 000, 1 -- 17 (2019) Scattered-light images of fluffy dust discs 11 tive to the disc scattered light colours in near-infrared wave- lengths, although these can affect magnitude of fractional luminosity (Section 4.3). • The effective albedo for aggregate with various frac- tal dimensions and monomer radius are computed. As frac- tal dimension or monomer radius increases, wavelength de- pendence of the effective albedo varies from blue to red at near-infrared wavelengths. BCCA can show reddish colours in near-infrared wavelengths when the monomer is about micron-sized. On the other hand, smaller monomers makes BCCA a bluer scatterer in near-infrared wavelength (Section 5). ACKNOWLEDGEMENTS We sincerely thank the referee for a thorough and careful reading of the manuscript. R.T. would like to thank Daniel Mackowski and Yasuhiko Okada for the availability of the T-Matrix code with the QMC method. R.T. also thanks Cornelis P. Dullemond for making the RADMC-3D code pub- lic. R.T. thanks Robert Botet for fruitful discussion. R.T. was supported by a Research Fellowship for Young Scien- tists from the Japan Society for the Promotion of Science (JSPS) (17J02411). REFERENCES Augereau, J. C., Lagrange, A. M., Mouillet, D., Papaloizou, J. C. B., & Grorod, P. A. 1999, A&A, 348, 557 Avenhaus, H., Quanz, S. P., Schmid, H. M., et al. 2014, ApJ, 781, 87 Avenhaus, H., Quanz, S. P., Garufi, A., et al. 2018, ApJ, 863, 44 Berry, M. V., & Percival, I. C. 1986, Optica Acta, 33, 577 Bentley, M. S., Schmied, R., Mannel, T., et al. 2016, Nature, 537, 73 Blum, J., Schrapler, R., Davidsson, B. J. R., & Trigo-Rodr´ıguez, J. M. 2006, ApJ, 652, 1768 Bohren, C. F., & Huffman, D. R. 1983, New York: Wiley, 1983, Botet, R., Rannou, P., & Cabane, M. 1997, Appl. Opt., 36, 8791 Brisset, J., Heisselmann, D., Kothe, S., Weidling, R., & Blum, J. 2017, A&A, 603, A66 Bruggeman, D. A. G. 1935, Annalen der Physik, 416, 636 Canovas, H., M´enard, F., Hales, A., et al. 2013, A&A, 556, A123 Draine, B. T., & Lee, H. M. 1984, ApJ, 285, 89 Duchene, G., McCabe, C., Ghez, A. M., & Macintosh, B. A. 2004, ApJ, 606, 969 Dullemond, C. P., & Natta, A. 2003, A&A, 408, 161 Dullemond, C. P., Juhasz, A., Pohl, A., et al. 2012, Astrophysics Source Code Library, ascl:1202.015 Fukagawa, M., Tamura, M., Itoh, Y., et al. 2010, PASJ, 62, 347 Fulle, M., Della Corte, V., Rotundi, A., et al. 2015, ApJ, 802, L12 Fulle, M., Della Corte, V., Rotundi, A., et al. 2016, MNRAS, 462, S132 Halder, P., Deb Roy, P., & Das, H. S. 2018, Icarus, 312, 45 Henning, T., & Stognienko, R. 1996, A&A, 311, 291 Itoh, Y., Oasa, Y., Kudo, T., et al. 2014, Research in Astronomy and Astrophysics, 14, 1438-1446 Kempf, S., Pfalzner, S., & Henning, T. K. 1999, Icarus, 141, 388 Mackowski, D. W., & Mishchenko, M. I. 1996, Journal of the Optical Society of America A, 13, 2266 Mannel, T., Bentley, M. S., Schmied, R., et al. 2016, MNRAS, 462, S304 Min, M., Hovenier, J. W., & de Koter, A. 2005, A&A, 432, 909 12 R. Tazaki et al. Min, M., Kama, M., Dominik, C., & Waters, L. B. F. M. 2010, A&A, 509, L6 Min, M., Dullemond, C. P., Kama, M., & Dominik, C. 2011, Icarus, 212, 416 Min, M., Canovas, H., Mulders, G. D., & Keller, C. U. 2012, A&A, 537, A75 Min, M., Rab, C., Woitke, P., Dominik, C., & M´enard, F. 2016, A&A, 585, A13 Mukai, S., Mukai, T., Giese, R. H., Weiss, K., & Zerull, R. H. 1982, Moon and Planets, 26, 197 Mukai, T., Ishimoto, H., Kozasa, T., Blum, J., & Greenberg, J. M. 1992, A&A, 262, 315 Mulders, G. D., Min, M., Dominik, C., Debes, J. H., & Schneider, G. 2013, A&A, 549, A112 Munoz, O., Moreno, F., Vargas-Mart´ın, F., et al. 2017, ApJ, 846, 85 Kataoka, A., Tanaka, H., Okuzumi, S., & Wada, K. 2013, A&A, 557, L4 Kataoka, A., Okuzumi, S., Tanaka, H., & Nomura, H. 2014, A&A, 568, A42 Kenyon, S. J., & Hartmann, L. 1987, ApJ, 323, 714 Kimura, H., Kolokolova, L., & Mann, I. 2006, A&A, 449, 1243 Kirchschlager, F., & Wolf, S. 2014, A&A, 568, A103 Kolokolova, L., Kimura, H., Kiselev, N., & Rosenbush, V. 2007, A&A, 463, 1189 Kothe, S., Blum, J., Weidling, R., & Guttler, C. 2013, Icarus, 225, 75 Laor, A., & Draine, B. T. 1993, ApJ, 402, 441 Long, Z. C., Fernandes, R. B., Sitko, M., et al. 2017, ApJ, 838, 62 Okada, Y. 2008, J. Quant. Spec. Radiat. Transf., 109, 17 19 Okuzumi, S., Tanaka, H., Kobayashi, H., & Wada, K. 2012, ApJ, 752, 106 Ormel, C. W., Spaans, M., & Tielens, A. G. G. M. 2007, A&A, 461, 215 Perrin, M. D., Schneider, G., Duchene, G., et al. 2009, ApJ, 707, L132 Pinte, C., Fouchet, L., M´enard, F., Gonzalez, J.-F., & Duchene, G. 2007, A&A, 469, 963 Quanz, S. P., Schmid, H. M., Geissler, K., et al. 2011, ApJ, 738, 23 Silber, J., Gledhill, T., Duchene, G., & M´enard, F. 2000, ApJ, 536, L89 Shen, Y., Draine, B. T., & Johnson, E. T. 2008, ApJ, 689, 260-275 Shen, Y., Draine, B. T., & Johnson, E. T. 2009, ApJ, 696, 2126 Stolker, T., Dominik, C., Min, M., et al. 2016, A&A, 596, A70 Suyama, T., Wada, K., & Tanaka, H. 2008, ApJ, 684, 1310 Tanii, R., Itoh, Y., Kudo, T., et al. 2012, PASJ, 64, 124 Tazaki, R., Tanaka, H., Okuzumi, S., Kataoka, A., & Nomura, H. 2016, ApJ, 823, 70 (Paper I) Tazaki, R., & Tanaka, H. 2018, ApJ, 860, 79 (Paper II) Poteet, C. A., Chen, C. H., Hines, D. C., et al. 2018, ApJ, 860, 115 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2008, ApJ, 677, 1296 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2011, ApJ, 737, 36 Ysard, N., Jones, A. P., Demyk, K., Bout´eraon, T., & Koehler, M. 2018, A&A, 617, A124 Zubko, V. G., Mennella, V., Colangeli, L., & Bussoletti, E. 1996, MNRAS, 282, 1321 Table A1. Disc Scattered Light Colour for BCCA with Different Methods Method TMM MFT EMT DHS ηI −0.51 −0.67 0.53 0.030 ηPI −0.41 −0.67 0.65 0.099 APPENDIX A: COMPARISON BETWEEN APPROXIMATE METHODS We perform radiative transfer simulation of the BCCA model with approximate methods; modified mean field the- ory (MMF; Paper II), effective medium theory with Maxwell Garnett mixing rule (EMT; Mukai et al. 1992; Kataoka et al. 2014), and DHS, and then the results are compared to those obtained by TMM. Figure A1 shows scattered-light images of the disc at the wavelength of λ = 1.6 µm and the inclination angle of i = 60◦ where the dust model is the BCCA model with N = 1024 and R0 = 0.1 µm. As shown in Paper II, for this BCCA model, MMF coincide with the mean field theory (MFT) results at λ > 0.85 µm. Hence, we use MFT instead of MMF. Figure A1 clearly shows that MFT reproduces the TMM results, whereas EMT and DHS fail. First of all, we study the difference of each method in total intensity of the discs (middle panels in Figure A1). All methods show the front-back asymmetry. Both MFT and DHS show similar results to the TMM result; however, the EMT result significantly deviates from the TMM result. This is because EMT produces extremely faint backward scattering (Figure A2, see also Figure 5 of Paper I) due to destructive interference of scattered waves. In other words, extremely strong forward scattering predicted by the EMT model means that most of the photons coming from the star are just passing through the particle, and they are not likely to be scattered toward the observer. As a result, when EMT is applied to obtain optical properties of BCCA, the disc scattered light becomes too faint. Secondly, we study polarisation fraction (bottom panels in Figure A1). Figure A1 shows that TMM and MFT show almost the same polarisation fraction; however, EMT and DHS overestimates and underestimates polarisation frac- tion, respectively (Figure A3). The low polarisation fraction of DHS is due to scattering property of the particle (Fig- ure A2). The degree of polarisation of the BCCA model at wavelength 1.6 µm shows has a bell shaped angular profile (almost symmetric with respect to θ = 90◦), and maximum degree of polarisation is 96% at θ = 90◦ (Figure A2, see also Figure 15 of Paper I). However, the DHS model predicts the degree of polarisation is 40% at θ = 90◦, which is signifi- cantly lower than the TMM result. As a consequence, polar- isation fraction of DHS becomes lower than that of TMM. The overestimation of polarisation fraction of EMT is due to a radiative transfer effect. As shown in Paper I, both EMT and MFT show a similar polarisation profile to TMM 9. Nevertheless, EMT significantly overestimates polarisation 9 The degree of polarisation obtained by MFT is almost the same as that of the RGD theory because phase shift is less than unity MNRAS 000, 1 -- 17 (2019) Scattered-light images of fluffy dust discs 13 Figure A1. Polarised intensity (top), total intensity (middle), and polarisation fraction (bottom). From left to right, TMM, MFT, EMT, and DHS are used. Wavelength is λ = 1.6 µm and the inclination angle is i = 60◦. The dust model adopted here is BCCA with N = 1024 and R0 = 0.1 µm, corresponding to the characteristic radius Rc = 4.8 µm. fraction. As we have mentioned, in the EMT model, large angle scattering is suppressed by destructive interference, and hence, multiple scattering at the disc surface layer is not likely to occur. Since MFT can reproduce phase func- tion obtained by the TMM correctly (see Figure 5 in Paper I), MFT can mimic the depolarisation effect in radiative transfer, whereas EMT fails. The DHS model show some slits in the image of po- larised intensity. This is because oscillatory behaviour in a angular profile of the degree of polarisation originated from resonance of smooth spherical surface. Since the degree of polarisation becomes positive and negative in the oscillatory behaviour, polarisation vectors in DHS are very different from the TMM results. In the DHS model, optical proper- ties are averaged over the size distribution; nevertheless, the oscillatory behaviour remains. This is mainly because astro- nomical silicate is less absorbing at infrared wavelengths. Figure A4 show scattered-light colour in total inten- for the dust model and wavelength now we concern (see also Paper II). MNRAS 000, 1 -- 17 (2019) sity and in polarised intensity calculated at inclination an- gle i = 60◦ obtained by various approximate methods. Table A1 shows colours of the BCCA model for various solution methods for their optical properties. The MFT model is able to reproduce the TMM results compared to other approxi- mate methods. DHS shows red colour in total intensity. It is also found that once the EMT is used to obtain optical properties of large BCCA, the disc becomes faint and red- dish; however, it should be strengthen that this is due to the artefact arising from the use of the Mie theory. Finally, we show some obtained radiative transfer im- ages for large BCCA studied in Section 4.2. By applying MFT to BCCA with Rc = 1 µm to 1 mm, we obtain Figure A5. Because of saturation of scattering properties of BCCA (Figures 6 and 7), scattered light images of BCCA with var- ious radius look similar (see also Section 4.2). 14 R. Tazaki et al. Figure A2. Phase function Z11 (left), and the degree of linear polarisation −Z12/Z11 (right). The dust model is the BCCA model. Black line is the result obtained by TMM method (rigorous numerical method), while others (MFT, EMT, DHS) are those obtained by using the approximate method for optical property calculations. MFT can reproduce the TMM results for both intensity and polarisation fraction, whereas EMT and DHS fail. Hatched region indicates a range of scattering angle to be observed for a disc with the flaring index β = 1.25 and the inclination angle i = 60◦. when the aggregate radius exceeds 1.0 µm, the disc total scattered light becomes faint with increasing the aggregate radius. This is because scattering phase function of large compact dust aggregate is sharply peaked at forward scat- tering angles, and then this results in reducing effective sin- gle scattering albedo (Dullemond & Natta 2003; Min et al. 2010; Mulders et al. 2013). Second, we argue polarisation fraction. When the radius of compact dust aggregate is 0.1 µm, polarisation fraction becomes more than 90% because Rayleigh scattering hap- pens. As the aggregate radius increases, polarisation frac- tion decreases. However, once the aggregate radius exceeds 1.0 µm, polarisation fraction increases again as the aggre- gate radius increases. This is due to the effect of Brewster scattering, which finally brings the degree of polarisation to 100% at the Brewster angle (Min et al. 2016) in the limit of sufficiently large grain radius. Based on light scattering simulations of compact dust aggregates, Min et al. (2016) showed that compact dust aggregates do not show high de- gree of polarisation at the Brewster scattering, unlike the prediction of the DHS method. The perfect polarisation at the Brewster angle is though to be arising from the igno- rance of the surface roughness of compact dust aggregates in the DHS method (Min et al. 2016), and hence we expect that the realistic compact dust aggregates may not show the increase of polarisation fraction at the Brewster angle. B2 Scattered-light colours Figure B4 shows disc scattered-light colour in total intensity and in polarised intensity for the compact dust aggregate models. Table B1 shows colours of compact aggregates for various aggregate radii. When the radius of the dust aggregate is smaller than the wavelength, the disc is faint and very blue in total in- tensity. As the radius increases, disc luminosity increases and disc colour becomes grey when the aggregate radius is comparable to the wavelength. Once the aggregate radius MNRAS 000, 1 -- 17 (2019) Figure A3. Polarisation fraction as a function of scattering an- gles at R = 50 au measured from simulated images in Figure A1. APPENDIX B: COMPACT DUST AGGREGATES Scattered-light images and colours of the compact dust ag- gregate models with various radii are presented. B1 H-band Images Figure B1 shows scattered-light images at λ = 1.6 µm for the compact dust aggregate models and the single monomer model for comparison. By using the scattered-light mapping, Figure B3 shows scattered-light intensity and polarisation fraction as a function of scattering angle at R =50 au. First of all, we discuss total intensity. When the ag- gregate radius is smaller than λ/2π, the disc is faint and shows weak front-back brightness asymmetry. Once the ag- gregate radius exceeds λ/2π, the disc begins to show front- back brightness asymmetry in total intensity. In addition, 10−310−210−1100101102103104105020406080100120140160180Z11[(cm2/g)/str]Scatteringangle[◦]TMMMFTEMTDHS10−310−210−1100101102103104105020406080100120140160180−1−0.500.51020406080100120140160180−Z12/Z11Scatteringangle[◦]TMMMFTEMTDHS−1−0.500.5102040608010012014016018000.10.20.30.40.50.60.70.80.91020406080100120140160180i=60◦PolarizationfractionScatteringangle[deg]TMMMFTEMTDHS Scattered-light images of fluffy dust discs 15 Figure A4. Fractional luminosity of an inclined disc (i = 60◦) for total intensity (left) and polarised intensity (right). Figure A5. Polarised intensity (top), total intensity (middle), and polarisation fraction (bottom). From left to right, scattered-light images of BCCA with Rc = 1 µm, 10 µm, 100 µm, and 1 mm are shown, respectively, where optical properties of BCCA are computed by using MFT. MNRAS 000, 1 -- 17 (2019) 10−210−111.21.41.61.822.2i=60◦LIdisk,λ/Lstar,λλ(µm)TMMMFTEMTDHS10−410−310−210−111.21.41.61.822.2i=60◦LPIdisk,λ/Lstar,λλ(µm)TMMMFTEMTDHS 16 R. Tazaki et al. Figure B1. Polarised intensity (top), total intensity (middle), and polarisation fraction (bottom). Scattered-light images of the single monomer model (leftmost) and the compact dust aggregate models (right three columns). Figure B2. Phase function Z11 (left), and the degree of linear polarisation −Z12/Z11 (right). Optical properties of compact dust aggregates are obtained by DHS. Hatched region indicates a range of scattering angle to be observed for a disc with the flaring index β = 1.25 and the inclination angle i = 60◦. MNRAS 000, 1 -- 17 (2019) 100101102103104105020406080100120140160180Z11[(cm2/g)/str]Scatteringangle[◦]Compact(0.1µm)Compact(0.3µm)Compact(1.0µm)Compact(3.0µm)Compact(5.0µm)100101102103104105020406080100120140160180−1−0.500.51020406080100120140160180−Z12/Z11Scatteringangle[◦]Compact(0.1µm)Compact(0.3µm)Compact(1.0µm)Compact(3.0µm)Compact(5.0µm)−1−0.500.51020406080100120140160180 Scattered-light images of fluffy dust discs 17 where the magnitude of scattering vector q = 2k sin(θ/2) and S(q) is the structure factor (Paper I). When qRg (cid:29) 1 and df = 2, we can approximately decompose the scatter- ing phase function of fluffy dust aggregates by the sum of coherent and incoherent contribution: 11,agg + Sincoherent , S11,agg(µ) = Scoherent Scoherent 11,agg Sincoherent 11,agg (cid:39) N2S11,mono(µ)δ(µ − 1), (cid:39) NS11,mono(µ)[(qR0)−2 + 1], 11,agg (C3) (C4) (C5) where we have used Equation (29) of Paper I. Using Equa- tions (C1 and C3) and qR0 (cid:28) 1, we obtain g (cid:39) µ(qR0)−2S11,mono(µ)dµ, 2πN (C6) k2Csca,agg Þ 1 −1 Since scattering cross section is approximately proportional to k2 (Berry & Percival 1986) and S11,mono ∝ k6 (Rayleigh scattering) (Bohren & Huffman 1983), the asymmetry pa- rameter g becomes wavelength independent. Therefore, suf- ficiently large dust aggregates (qRg (cid:29) 1) with df = 2 con- taining small monomers (qR0 (cid:28) 1) yield wavelength inde- pendent asymmetry parameter. This paper has been typeset from a TEX/LATEX file prepared by the author. Figure B3. Polarisation fraction as a function of scattering an- gles at R = 50 au measured from simulated images in Figure B1. Table B1. Disc Scattered Light Colour for Compact Aggregates with Various Radii Dust model 0.1 µm 0.3 µm 1.0 µm 3.0 µm 5.0 µm ηI −2.4 −0.93 0.030 0.84 0.70 ηPI −2.3 2.6 0.099 0.60 0.11 exceeds the wavelength, the disc becomes reddish and faint as expected by Mulders et al. (2013). Next, we discuss polarised intensity. Similar to the case of total intensity, small dust aggregates show blue colours in polarised intensity. When the aggregate radius is comparable to the wavelength, polarised intensity is reddish. For the large aggregate radius, polarised intensity show grey colours; however, it is partially due to Brewster scattering. Hence, if we consider the effect of surface roughness of compact dust aggregates, polarised intensity colours might be more reddish, although more detail computations are necessary. APPENDIX C: THE ASYMMETRY PARAMETER OF SUFFICIENTLY LARGE BCCA The asymmetry parameter can be expressed as Þ 1 2π µS11,agg(µ)dµ, (C1) g = k2Csca,agg −1 where Csca,agg is the scattering cross section of the dust ag- gregate, k is the wave number, S11,agg is a (1,1) element of scattering matrix of dust aggregates, and µ = cos θ; where θ is the scattering angle. Using the single scattering assump- tion, a scattering matrix element of dust aggregates can be written by S11,agg(µ) = N2S11,mono(µ)S(q), (C2) MNRAS 000, 1 -- 17 (2019) 00.20.40.60.81020406080100120140160180i=60◦CompactPolarizationfractionScatteringangle[deg]0.1µm0.3µm1.0µm3.0µm5.0µm 18 R. Tazaki et al. Figure B4. Scattered-light colours in total intensity (left) and in polarised intensity (right) of the compact dust aggregate models. The disc inclination is assumed to be i = 60◦. Once the size parameter of the grain exceeds unity, the disc scattered light becomes faint and red as the grain radius increases. MNRAS 000, 1 -- 17 (2019) 10−210−110011.21.41.61.822.2i=60◦LIdisk,λ/Lstar,λλ(µm)Compact(0.1µm)Compact(0.3µm)Compact(1.0µm)Compact(3.0µm)Compact(5.0µm)10−310−210−111.21.41.61.822.2i=60◦LPIdisk,λ/Lstar,λλ(µm)Compact(0.1µm)Compact(0.3µm)Compact(1.0µm)Compact(3.0µm)Compact(5.0µm)
1102.5330
2
1102
2011-03-14T21:00:01
VLT/NACO astrometry of the HR8799 planetary system. L'-band observations of the three outer planets
[ "astro-ph.EP" ]
HR8799 is so far the only directly imaged multiple exoplanet system. The orbital configuration would, if better known, provide valuable insight into the formation and dynamical evolution of wide-orbit planetary systems. We present L'-band observations of the HR8799 system obtained with NACO at VLT, adding to the astrometric monitoring of the planets HR8799b, c and d. We investigate how well the two simple cases of (i) a circular orbit and (ii) a face-on orbit fit the astrometric data for HR8799d over a total time baseline of ~2 years. The results indicate that the orbit of HR8799d is inclined with respect to our line of sight, and suggest that the orbit is slightly eccentric or non-coplanar with the outer planets and debris disk.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. bergforsHR8799 July 31, 2021 c(cid:13) ESO 2021 VLT/NACO astrometry of the HR 8799 planetary system (cid:63) L(cid:48)-band observations of the three outer planets C. Bergfors1(cid:63)(cid:63), W. Brandner1, M. Janson2, R. Kohler1,3, and T. Henning1 1 Max-Planck-Institut fur Astronomie, Konigstuhl 17, 69117 Heidelberg, Germany e-mail: [email protected] 2 University of Toronto, Dept. of Astronomy, 50 St George Street, Toronto, ON, M5S 3H8, Canada 3 Landessternwarte, Zentrum fur Astronomie der Universitat Heidelberg, Konigstuhl, 69117 Heidelberg, Germany 1 1 0 2 r a M 4 1 . ] P E h p - o r t s a [ 2 v 0 3 3 5 . 2 0 1 1 : v i X r a Recieved 11 January 2011/ Accepted 16 February 2011 ABSTRACT Context. HR 8799 is so far the only directly imaged multiple exoplanet system. The orbital configuration would, if better known, provide valuable insight into the formation and dynamical evolution of wide-orbit planetary systems. Aims. We present data which add to the astrometric monitoring of the planets HR 8799 b, c and d. We investigate how well the two simple cases of (i) a circular orbit and (ii) a face-on orbit fit the astrometric data for HR 8799 d over a total time baseline of ∼ 2 years. Methods. The HR 8799 planetary system was observed in L(cid:48)-band with NACO at VLT. Results. The results indicate that the orbit of HR 8799 d is inclined with respect to our line of sight, and suggest that the orbit is slightly eccentric or non-coplanar with the outer planets and debris disk. Key words. planetary systems -- Stars: individual (HR 8799) 1. Introduction As the first and so far only directly imaged multiple exoplanet system, the HR 8799 system carries the promise of providing valuable insight into the structure and characteristics of plane- tary systems. While more than 500 extrasolar planets have now been discovered, most have been found by radial velocity and transit searches; the sample of known exoplanets is thus heav- ily biased towards short-period planets. Directly imaged giant extrasolar planets provide a necessary complement to these indi- rect detection techniques for a full picture of the characteristics of planets, and are crucial for theories of planet formation. Challenging as it may be to directly image planets, whose relatively faint light is easily lost in the bright stellar glare, several confirmed companions are known. As of November 2010, 7 planetary mass objects belonging to stars, including the quadruple-planet HR 8799 system, have been discovered with direct imaging (Fomalhaut b, Kalas et al. (2008); β Pic b, Lagrange et al. (2009, 2010); 1RXS J160929.1-210524 b, Lafreni`ere et al. (2008, 2010), and HR 8799 bcde, Marois et al. (2008, 2010)). The HR 8799 system is especially interesting since its multiple planet configuration allows for comparison of the characteristics of planets within the same environment of formation and evolution. The star is a young (30-160 Myr, Marois et al. 2008) A5 V star at at distance of 39.4 pc from the Sun (van Leeuwen 2007), surrounded by a debris disk (Rhee et al. 2007; Su et al. 2009). Three of the planets in the system, HR 8799 b, c and d, were discovered in 2008 and an additional planet, HR 8799 e, in 2010 (Marois et al. 2008, 2010), adding (cid:63) Based on observations collected at the European Southern Observatory, Chile, under programme ID 084.C-0072 (cid:63)(cid:63) Member of the International Max Planck Research School for Astronomy and Cosmic Physics at the University of Heidelberg up to at least four giant planets of masses 7-10 MJ at projected separations 14.5, 24, 38 and 68 AU from the central star. The astrometric analysis of the three outermost planets at the time of their discovery provided evidence that the planets are co-moving with the star, and suggested that their orbits are almost circular and seen close to face-on. However, dynamical modelling of the HR 8799 system has since shown that this ini- tially presumed configuration of orbits is unlikely for reasons of orbital stability of the system (Fabrycky & Murray-Clay 2010; Reidemeister et al. 2009; Go´zdziewski & Migaszewski 2009). Fabrycky & Murray-Clay (2010) found that for the masses de- rived by Marois et al. (2008) and circular, face-on orbits, the system would become unstable at an age of only ∼ 105 years, i.e. significantly younger than its assumed present age. Stable mean motion resonance configurations were found for the three outer planets known at the time by Fabrycky & Murray-Clay (2010); Go´zdziewski & Migaszewski (2009) and Reidemeister et al. (2009), and Marois et al. (2010) found stable resonant con- figurations including also the fourth planet. In this paper, we present astrometric measurements of the three outermost planets in the HR 8799 system. The observations were obtained with NACO at VLT in September 2009 -- one year after the discovery of planets b, c and d. We investigate how well two simple models with (i) a circular orbit and (ii) a face-on orbit fit the astrometric data when the new observations are included. 2. Observations and data reduction Acquisition images of HR 8799 in L(cid:48)-band were obtained with NACO/VLT (Lenzen et al. 2003; Rousset et al. 2003) on the nights of October 5 and 6, 2009, as part of the observation programme 084.C-0072, in which a spectrum of the planet HR 8799 c was obtained (Janson et al. 2010). The observations 1 C. Bergfors et al.: VLT/NACO astrometry of the HR 8799 planetary system est values at each pixel in order to remove the stellar flux. The skyframes were subtracted from each frame in the cubes, and bad pixels were replaced by the mean value of neighbouring pix- els. The frames were aligned by fitting a 2-D Gaussian to the star and measuring the centroid position, and then co-added to remove residual tip-tilt between individual frames of each data cube and produce one image per rotation angle for each night. Frames of poor quality or with the target too close to the detector edge, hence cutting out the planets, were rejected, resulting in a combination of 4-7 data cubes for each final image (4 images in total, one for each rotation angle on each night). Unsharp mask- ing was used on the co-added frames by smoothing one version of the image using a boxcar average with 15 pixels and subtract- ing the smoothed image from the original. The positions of the planets and central star were determined using the IRAF imex- amine task. While the brightest planet, HR 8799 c, was clearly detectable in all four images, the position of the fainter b-planet could only be determined from 3 measurements. Speckle con- tamination obscured planet d in the 33◦ rotated images and the position could thus only be determined from 2 frames. The re- ported position of planet e coincides with the third diffraction ring and is not detected with significant counts. Fig. 1 shows ro- tated and subtracted images from both nights, When imaging sources with vastly different spectral energy distributions through a broadband filter at an airmass > 1.0, the effect of differential atmospheric refraction on the relative as- trometry has to be considered (see, e.g., Hełminiak 2009). An effective wavelength λeff = 3.777µm for our NACO L(cid:48) imaging observation of the star HR 8799 was computed by convolving a spectrum of a star of similar spectral type from the IRTF SpeX spectral library (Cushing et al. 2005; Rayner et al. 2009) with the L(cid:48) transmission curve from the NACO Usermanual. For the three exoplanets, an effective wavelength of 3.905 ± 0.010 µm was computed using model spectra (Burrows et al. 2006, and priv. comm.) for an effective temperature of 1100 K and log g in the range 3.0 to 5.0. As a consistency check, we also computed the effective wavelength for L(cid:48) imaging observations of Jupiter based on the IRTF SpeX library, which resulted in λeff = 3.907 µm. Hence, once effective temperatures in the atmospheres of substellar ob- jects are low enough to allow for the pronounced presence of wa- ter, methane and CO molecular absorption bands, the λeff values for L(cid:48) observations seem to show little dependence on effective temperature and surface gravity. Next, we used the model fits for the refractive index of humid air in the infrared as computed by Mathar (2007) to estimate the amplitude of differential atmospheric refraction of HR 8799 and its exoplanets. We found that this effect caused a shift of ≈ 5− 6 mas of the planet positions along the parallactic angle. The positions of the planets relative to the star are presented in Table 1, together with all previously published position measurements. The errorbars were estimated by introducing artificial structures in the form of Gaussians with the same approximate peak flux and the same separations from the star as the real planets but at different angles. The deviations from the known positions were measured for a set of 3 different angles per planet and image. The standard deviations of all measurements for each "fake planet" are the estimated errors, mainly due to residual speckles for the two innermost planets. Fig. 1. NACO L(cid:48)-band images of the HR 8799 system. Rotated and subtracted 3(cid:48)(cid:48) x 3(cid:48)(cid:48) images from observations on October 5 (left) and October 6 (right) 2009. North is up and east is to the left. Scaling is linear. Fig. 2. Observations of HR 8799 d (see Table 1). A distance to the star of 39.4 pc is assumed, and the nominal circular, face-on orbit at 24 AU is overplotted. The star is marked with an asterisk at (0,0). The triangle marks our NACO L(cid:48) observation. were acquired in cube-mode and consisted on each night of 2 sets of 10 data cubes, one set taken at the default orientation with north up and one rotated by 33◦. The frames were obtained with the purpose of checking the alignment for slit orientation. Each cube contained 749 usable frames on October 5 and 1499 frames on October 6 with individual integration time 20.2 ms, yielding a total integration time of 15 s and 30 s respectively per data cube. Images of the astrometric binary HD 211742 obtained in September 2009 were retrieved from the ESO (VLT) archive and used for calibration of plate scale and true north orientation of the detector. We derive a field rotation of -0.6◦ ± 0.2◦ and plate scale of 27.1 mas/px, consistent with the 27.2 mas/px of the L27 camera described in the NACO Usermanual, assuming a sys- tematical error of 0.2◦ and 0.3 mas/px (1% of the pixel scale, see Kohler 2008). The data reduction was performed using IRAF and IDL. Skyframes were constructed by averaging the sum frame of 10 data cubes obtained at the approximate same time and airmass with the star in different dither positions, rejecting the 2-3 high- 2 C. Bergfors et al.: VLT/NACO astrometry of the HR 8799 planetary system 3. Results and discussion 3.1. Astrometric measurements of HR 8799 b, c and d The projected orbital motions of the two outermost planets b and c are slow (the periods are Pb ∼ 460 and Pc ∼ 190 years respec- tively, Marois et al. 2008), and the nominal mostly circular, face- on orbits are still consistent with observations when our data are added to the previously published astrometry. However, obser- vations in 2008 and 2009 of planet d by Currie et al. (2011) and Hinz et al. (2010)1 together with our observations suggest that the orbit of this planet might be eccentric and/or inclined. All published astrometric measurements of HR 8799 d are plotted in Fig. 2. 3.2. Testing the cases of i = 0 and e = 0 for HR 8799 d We now want to test if the observed changes in position angle and separation are in agreement with different models. Fabrycky & Murray-Clay (2010) investigated the stability of different pos- sible orbital families for the HR 8799 system. Many of these are for circular (e = 0) or face-on (i = 0) orbits. We modelled these two simple cases for HR 8799 d from the astrometric observa- tions over a ∼ 2 year time baseline. The nominal system mass of 1.5 M(cid:12) (Marois et al. 2008) was used to compute orbital period from semi-major axis. We did not attempt to fit model orbits that are both inclined and eccentric, since this would add two free parameters and reduce the statistical significance of the result. For the following orbital fits we assumed errors as given in the literature and listed in Table 1. We note, however, that while the literature measurements are probably subject to systematic er- rors of the same order as the ones present in the VLT/NACO set, systematic errors (plate scale, detector orientation) between the different telescopes and instruments used are not necessarily included in these error bars. 3.2.1. Face-on orbit model for HR 8799 d Assuming a face-on orbit (i = 0) for HR 8799 d, the χ2 was computed for a grid of 100 semi-major axes and 100 eccen- tricities and is shown in Fig. 3. The cross marks the best fit and the contour lines surround the 68.3% and 99.73% confi- dence regions. We find that a face-on orbit must have eccentric- ity e >∼ 0.4 to fit the observations with 99.73% confidence. The system should have become dynamically stable at the assumed age of 60 Myr, and one of the stability criteria of Fabrycky & Murray-Clay (2010) is that the orbits of the planets do not cross: ad(1+ed) < 0.85ac(1−ec). Our derived 99.73% confidence mini- mum eccentricity of e ≈ 0.4 corresponds to an apastron distance of rap(d) ≈ 34 AU for the nominal semi-major axis of planet d at ad = 24 AU, thereby violating the mentioned stability criterium for the nominal orbit of planet c (ac = 38 AU). The stability cri- terium cannot be fulfilled with respect to both the outer planet c and the inner planet e even if the orbits of both c and e are perfectly circular. A face-on orbit is, with the astrometric data points taken at face value, unstable because of the high eccen- tricity, and not likely to represent the true orbit of the planet. 1 The MMT/Clio data by Hinz et al. (2010) were rereduced by Currie et al. (2011). We have adopted the Currie et al. (2011) astrometry for this analysis. Fig. 3. χ2-fit of eccentricity and semi-major axis for the orbit of HR 8799 d assuming i = 0 (face-on orbit). The best fit is marked by a cross and the contour lines show the 68.3% and 99.73% confidence regions. Fig. 4. χ2-fit of inclination and semi-major axis for the orbit of HR 8799 d, assuming e = 0 (circular orbit). The best fit is marked by a cross, and the contour lines show the 68.3% and 99.73% confidence regions. 3.2.2. Circular orbit model for HR 8799 d The case of zero eccentricity (e = 0) was considered by varying the orbital inclination and semimajor axis in the same way as de- scribed above. Figure 4 shows the χ2 as a function of inclination and semi-major axis for a circular orbit. We find that the incli- nation is greater than 43◦ within the 99.73% confidence limits, with a best fit of i = 63◦ and a = 36 AU. This is consistent with 3 C. Bergfors et al.: VLT/NACO astrometry of the HR 8799 planetary system Table 1. Relative positions of the HR 8799 planets. Epoch 1998.83 2002.54 2004.53 2007.58 2007.81 2008.52 2008.61 2008.71 2008.89 2009.58 2009.58 2009.62 2009.70 2009.76 2009.77 2009.83 2010.53 2010.55 2010.83 HR 8799b ∆α, ∆δ (arcsec) 1.411±0.009, 0.986±0.009 1.481±0.023, 0.919±0.017 1.471±0.005, 0.884±0.005 1.522±0.003, 0.815±0.003 1.512±0.005, 0.805±0.005 1.527±0.004, 0.799±0.004 1.527±0.002, 0.801±0.002 1.528±0.003, 0.798±0.003 1.532±0.02, 0.796±0.02 ... ... 1.536±0.01, 0.785±0.01 1.538±0.03, 0.777±0.03 1.535±0.02,0.816±0.02 1.532±0.007, 0.783±0.007 ... ... ... ... HR 8799c ∆α, ∆δ (arcsec) ... ... -0.739±0.005, 0.612±0.005 -0.672±0.005, 0.674±0.005 -0.674±0.005, 0.681±0.005 -0.658±0.004, 0.701±0.004 -0.657±0.002, 0.706±0.002 -0.657±0.003, 0.706±0.003 -0.654±0.02, 0.700±0.02 ... ... ... -0.634±0.03, 0.697±0.03 -0.636±0.04, 0.692±0.04 -0.627±0.007, 0.716±0.007 ... ... ... ... HR8799d ∆α, ∆δ (arcsec) ... ... ... -0.170±0.008, -0.589±0.008 ... -0.208±0.004, -0.582±0.004 -0.216±0.002, -0.582±0.002 -0.216±0.003, -0.582±0.003 -0.217±0.02, -0.608±0.02 ... ... ... ... -0.270±0.07, -0.600±0.07 -0.241±0.007, -0.586±0.007 ... ... ... ... HR8799e ∆α, ∆δ (arcsec) ... ... ... ... ... ... ... ... ... -0.299±0.019, -0.217±0.019 -0.303±0.013, -0.209±0.013 ... ... ... -0.306±0.007, -0.217±0.007 -0.304±0.010, -0.196±0.010 -0.325±0.008, -0.173±0.008 -0.324±0.011, -0.175±0.011 -0.334±0.010, -0.162±0.010 Ref. 1 2 3 4 3 3 3 3 5 6 6 5 5 7 5 6 6 6 6 References. (1) Lafreni`ere et al. (2009); (2) Fukagawa et al. (2009); (3) Marois et al. (2008); (4) Metchev et al. (2009); (5) Currie et al. (2011); (6) Marois et al. (2010); (7) This work. Only statistical errors have been considered in the table. the asteroseismic constraints on the stellar rotational inclination of i >∼ 40◦, with a best fit of i = 65◦ derived by Wright et al. (2011). However, this inclination is higher than what has been derived for the orbit of planet b from a 10 year baseline of obser- vations (i ∼ 13− 23◦ for a circular orbit, Lafreni`ere et al. 2009), and also for the debris disk (3σ upper constraint, idisk < 40◦, Moro-Mart´ın et al. 2010). The observed configuration with four visible planets around the star at very different position angles supports a low inclination, if the orbits are coplanar. With the astrometric data points taken at face value, the high inclination from the fit with e = 0 thus indicates that the orbital eccentricity of HR 8799 d is non-zero, or that the planetary orbits are non- coplanar. 4. Conclusions The initial astrometric analysis performed by Marois et al. (2008) suggested mostly face-on and circular orbits for the three planets of HR 8799 known at the time. The orbital periods of the outermost planets b and c are of the order of hundreds of years, and with our additional astrometric measurements of these plan- ets the orbits are still consistent with the nominal orbits. Purely circular, face-on and coplanar orbits have been shown to be an unlikely configuration for reasons of dynamical stability (Fabrycky & Murray-Clay 2010; Go´zdziewski & Migaszewski 2009; Reidemeister et al. 2009). Our analysis of the orbit of HR 8799 d implies that such a configuration is also inconsistent with the astrometric observations when recent observations by Hinz et al. (2010), Currie et al. (2011) and our NACO data are included. For a purely face-on orbit, the eccentricity of HR 8799 d is e >∼ 0.4 within 99.7% confidence. The system is not sta- ble for such high e and the orbit is hence likely to be inclined with respect to our line of sight. In the case of a purely circular orbit we find that the inclination is i > 43◦ within the 99.7% confidence limits. While the astrometric data is still limited, this result agrees well with recent constraints on the the stellar rota- tional inclination (Wright et al. 2011), but not with other mea- surements of the orbital inclination of planet b and the debris 4 disk. The current astrometric data still allow for different orbital planes for the individual planets and the debris disk (and hence for non-coplanar orbits). Continued astrometric monitoring over a longer time baseline is required in order to put stronger con- straints on the orbits of the HR 8799 planets. Acknowledgements. We would like to thank Adam Burrows for providing model spectra, which enabled us to quantify the effect of differential atmospheric refrac- tion. References Burrows, A., Sudarsky, D., & Hubeny, I. 2006, ApJ, 640, 1063 Currie, T., Burrows, A. S., Itoh, Y., et al. 2011, ArXiv e-prints Cushing, M. C., Rayner, J. T., & Vacca, W. D. 2005, ApJ, 623, 1115 Fabrycky, D. C. & Murray-Clay, R. A. 2010, ApJ, 710, 1408 Fukagawa, M., Itoh, Y., Tamura, M., et al. 2009, ApJ, 696, L1 Go´zdziewski, K. & Migaszewski, C. 2009, MNRAS, 397, L16 Hełminiak, K. G. 2009, New A, 14, 521 Hinz, P. M., Rodigas, T. J., Kenworthy, M. A., et al. 2010, ApJ, 716, 417 Janson, M., Bergfors, C., Goto, M., Brandner, W., & Lafreni`ere, D. 2010, ApJ, 710, L35 Kalas, P., Graham, J. R., Chiang, E., et al. 2008, Science, 322, 1345 Kohler, R. 2008, Journal of Physics Conference Series, 131, 012028 Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk, M. H. 2008, ApJ, 689, L153 Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk, M. H. 2010, ApJ, 719, 497 Lafreni`ere, D., Marois, C., Doyon, R., & Barman, T. 2009, ApJ, 694, L148 Lagrange, A., Bonnefoy, M., Chauvin, G., et al. 2010, Science, 329, 57 Lagrange, A., Gratadour, D., Chauvin, G., et al. 2009, A&A, 493, L21 Lenzen, R., Hartung, M., Brandner, W., et al. 2003, the Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol. 4841, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. M. Iye & A. F. M. Moorwood, 944 -- 952 in Presented at Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322, 1348 Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., & Barman, T. 2010, Nature, 468, 1080 Mathar, R. J. 2007, Journal of Optics A: Pure and Applied Optics, 9, 470 Metchev, S., Marois, C., & Zuckerman, B. 2009, ApJ, 705, L204 Moro-Mart´ın, A., Rieke, G. H., & Su, K. Y. L. 2010, ApJ, 721, L199 Rayner, J. T., Cushing, M. C., & Vacca, W. D. 2009, ApJS, 185, 289 Reidemeister, M., Krivov, A. V., Schmidt, T. O. B., et al. 2009, A&A, 503, 247 Rhee, J. H., Song, I., Zuckerman, B., & McElwain, M. 2007, ApJ, 660, 1556 Rousset, G., Lacombe, F., Puget, P., et al. 2003, in Presented at the Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol. 4839, C. Bergfors et al.: VLT/NACO astrometry of the HR 8799 planetary system Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. P. L. Wizinowich & D. Bonaccini, 140 -- 149 Su, K. Y. L., Rieke, G. H., Stapelfeldt, K. R., et al. 2009, ApJ, 705, 314 van Leeuwen, F. 2007, A&A, 474, 653 Wright, D. J., Chen´e, A., De Cat, P., et al. 2011, ApJ, 728, L20+ 5
1802.06856
2
1802
2018-02-21T19:19:05
The Habitable Zone of Kepler-16: Impact of Binarity and Climate Models
[ "astro-ph.EP", "astro-ph.SR" ]
We continue to investigate the binary system Kepler-16, consisting of a K-type main-sequence star, a red dwarf, and a circumbinary Saturnian planet. As part of our study, we describe the system's habitable zone based on different climate models. We also report on stability investigations for possible Earth-mass Trojans while expanding a previous study by B. L. Quarles and collaborators given in 2012. For the climate models we carefully consider the relevance of the system's parameters. Furthermore, we pursue new stability simulations for the Earth-mass objects starting along the orbit of Kepler-16b. The eccentricity distribution as obtained prefers values close to circular, whereas the inclination distribution remains flat. The stable solutions are distributed near the co-orbital Lagrangian points, thus enhancing the plausibility that Earth-mass Trojans might be able to exist in the Kepler-16(AB) system.
astro-ph.EP
astro-ph
Version: February 23, 2018 The Habitable Zone of Kepler-16: Impact of Binarity and Climate Models S. Y. Moorman1, B. L. Quarles2, Zh. Wang1, and M. Cuntz1 1Department of Physics, Box 19059 University of Texas at Arlington, Arlington, TX 76019; [email protected]; [email protected]; [email protected] 2Homer L. Dodge Department of Physics and Astronomy University of Oklahoma, Norman, OK 73019; [email protected] ABSTRACT We continue to investigate the binary system Kepler-16, consisting of a K-type main-sequence star, a red dwarf, and a circumbinary Saturnian planet. As part of our study, we describe the system's habitable zone based on different climate models. We also report on stability investigations for possible Earth-mass Trojans while expanding a previous study by B. L. Quarles and collaborators given in 2012. For the climate models we carefully consider the relevance of the system's parameters. Furthermore, we pursue new stability simulations for the Earth- mass objects starting along the orbit of Kepler-16b. The eccentricity distribution as obtained prefers values close to circular, whereas the inclination distribution remains flat. The stable solutions are distributed near the co-orbital Lagrangian points, thus enhancing the plausibility that Earth-mass Trojans might be able to exist in the Kepler-16(AB) system. Subject headings: astrobiology - binaries: general - celestial mechanics - planetary systems - stars: individual (Kepler-16) – 2 – 1. Introduction Kepler-16 is a well-documented example of a closely separated binary system with a Saturnian planet in a P-type orbit (Slawson et al. 2011; Doyle et al. 2011). P-type orbit means that the planet encircles both stars instead of only one star with the other star acting as a perturber (Dvorak 1982). Previous results on the existence and orbital properties of planets in binary systems have been given by, e.g., Raghavan et al. (2006, 2010) and Roell et al. (2012), among others. Detailed information on the abundance of circumstellar planets has been given by Wang et al. (2014) and Armstrong et al. (2014). So far, eleven circumbinary planets have been discovered by Kepler with Kepler-453b and Kepler-1647b constituting number 10 and 11, as reported by Welsh et al. (2015) and Kostov et al. (2016), respectively. The main purpose of the Kepler mission is to identify exoplanets via the transit method near or within the host star's habitable zone (HZ). The lion's share of stars-of-study encom- pass main-sequence stars of spectral types G, K, and M, with latter ones also referred to as red dwarfs. Recent catalogs of stars studied by Kepler have been given by Kirk et al. (2016) and Thompson et al. (2017). Here Thompson et al. (2017) offer the latest results for the general catalog from Kepler, as it contains all observed objects, including circumbinary planets, potentially habitable planets, and (most likely) non-habitable planets. On the other hand, the catalog by Kirk et al. (2016) is mostly focused on eclipsing binary systems. Previous theoretical work about circumbinary planets in binary systems has been given by, e.g., Kane & Hinkel (2013), Eggl et al. (2013), Haghighipour & Kaltenegger (2013), Cuntz (2014, 2015), Zuluaga et al. (2016), Popp & Eggl (2017), Shevchenko (2017), and Wang & Cuntz (2017), and references therein. These types of studies focus on the formation, orbital sta- bility, secular evolution, and/or environmental forcings pertaining to those systems. For example, recently, Wang & Cuntz (2017) presented fitting formulae for the quick determi- nation of the existence of P-type HZs in binary systems. Objects hosted by P-type systems which might be potentially habitable could include exoplanets, exomoons, and exo-Trojans. For Kepler-16, the latter two kinds of objects have been discussed by Quarles et al. (2012), hereafter QMC12. Kepler-16(AB) is a pivotal example of a planet-hosting binary; it is 61 parsecs (199 light years) from Earth (see Table 1); for more detailed information see Doyle et al. (2011), and references therein. The system consists of the primary star, Kepler-16A, a K-dwarf of about 0.69 M⊙, and the secondary star, Kepler-16B, a red dwarf star. The circumbinary planet of that system is similar to Saturn in mass and density. Kepler-16b has a nearly circular orbit with an eccentricity of approximately 0.007 and a small deviation in orbital inclination to that of its host stars indicating that it may have formed within the same circumbinary disk – 3 – as the two stars. Although Kepler-16b proves to be an interesting exoplanet, it is considered to be cold, gaseous, and ultimately uninhabitable. However, previous work by QMC12 has focused on the possibility of both Earth-mass exomoons and Trojans, which if existing may be potentially habitable. Among other considerations, we intend to expand the work by QMC12 in this article. The structure of our paper is as follows. In Section 2, we report the stellar parameters. A special effort is made to determine the effective temperature of Kepler-16B. Section 3 discusses the HZ of the Kepler-16(AB) binary system in consideration of different types of climate models available in the literature. For tutorial reasons, we also discuss the HZ of Kepler-16A, with Kepler-16B assumed absent. In Section 4, we consider the previous results by QMC12 for Earth-mass moons and Trojans in relationship to Kepler-16's HZ. Furthermore, additional stability simulations based on a modified version of the mercury6 integration package are pursued to explore the possible parameter space of stable objects in the Kepler-16(AB) system. Our summary and conclusions are given in Section 5. 2. Stellar Parameters Regarding our study, stellar parameters are of pivotal importance for the calculation of stellar HZs as well as for orbital stability simulations of possible exomoons and Trojan ob- jects. Most relevant parameters of the Kepler-16(AB) system have been previously reported by Doyle et al. (2011), as they announce a transiting circumbinary planet observed by the Kepler spacecraft. Kepler-16A was identified as a K-type main-sequence star with effective temperature, radius, and mass given as (see Table 1) 4450± 150 K, 0.6489± 0.0013 R⊙, and 0.6897 ± 0.0035 M⊙, respectively. Here the relative uncertainty bar is largest for the stellar effective temperature (see Table 2). However, less information has been conveyed for Kepler-16B, which based on its mass of about 0.20255 M⊙ (Doyle et al. 2011) is identified as a red dwarf. But Kepler-16B's effective temperature needs to be determined as well to compute the HZ for the Kepler-16 binary system. Thus, to determine Kepler-16B's stellar effective temperature, we utilize the mass – effective temperature relationship by Mann et al. (2013). They have analyzed moderate resolution spectra for a set of nearby K and M dwarfs with well-known parallaxes and interferometrically determined radii to define their effective temperatures, among other quantities. They also adopt state-of-the-art PHOENIX atmosphere models, as described. Thus, we conclude that the effective temperature of Kepler-16B is 3308± 110 K (see Fig. 1). Here the uncertainty bar has been estimated based on the results of similar objects included in the sample. From other work as, e.g., that by Kirkpatrick et al. (1991) and Baraffe et al. – 4 – (1998) the spectral type of Kepler-16B has been deduced as ∼M3 V. Both the effective temperature and radius of Kepler-16B are important for determining the different types of HZs of the Kepler-16(AB) system (see Sect. 4). 3. The Kepler-16 Habitable Zone A crucial aspect of this study is the evaluation of Kepler-16's HZ. The HZ is a region around a star or a system of stars in which terrestrial planets could potentially have surface temperatures at which liquid water could exist, given a sufficiently dense atmosphere (e.g., Kasting et al. 1993; Jones et al. 2001; Underwood et al. 2003). When determining the HZ, both inner limits and outer limits are calculated, in response to different types of criteria, thus defining the HZ. The determination of the location of the HZ is significant in the context of theoretical studies as well as for the purpose of planet search missions (e.g., Lammer et al. 2009; Kasting et al. 2014; Kaltenegger 2017, and references therein). Inner limits previously used for stellar HZs include those set by the recent Venus (RV), the runaway greenhouse effect, and the onset of water loss. Furthermore, outer limit of the stellar HZ has been set by the first CO2 condensation, the maximum greenhouse effect for a cloud-free CO2 atmosphere and the early Mars (EM) setting. For example, Kasting et al. (1993) describe the runaway greenhouse effect such that the greenhouse phenomenon is enhanced by water vapor, thus promoting surface warming. The latter further increases the atmospheric vapor content, thus resulting in an additional rise of the planet's surface temperature. Consequently, this will lead to the rapid evaporation of all surface water. On the other hand (see, e.g., Underwood et al. 2003), the water loss criterion means that an atmosphere is warm enough to have a wet stratosphere, from where water is gradually lost by atmospheric chemical processes to space. Table 3 shows the HZ limits for Kepler-16A, treated as a single star, for tutorial reasons. Here GHZ denotes the general habitable zone, bracketed by the runaway greenhouse and maximum greenhouse criteria, whereas RVEM denotes the kind of HZ, defined by the settings of recent Venus and early Mars; this latter type of HZ is also sometimes referred to as (most) optimistic HZ; see, e.g., Kaltenegger (2017) and references therein. Figure 2 and Tables 3 and 4 convey the results for the various HZ limits as well as for the GHZ and RVEM. The most recent results based on Kopparapu et al. (2013, 2014) have been included as well, which indicate updated HZ limits. For the inner and outer HZ limits, they assumed H2O and CO2 dominated atmospheres, respectively, while scaling the background N2 atmospheric pressure with the radius of the planet. Moreover, from said climate model, several equations were generated, which correspond to select inner and outer HZ limit criteria. – 5 – Surely, most of our study focuses on Kepler-16 as a binary thus taking into account both Kepler-16A (an orange dwarf) and Kepler-16B (a red dwarf); see Table 1 for data. The computation of the GHZ and RVEM of Kepler-16(AB) follows the work by Cuntz (2014, 2015) and Wang & Cuntz (2017). Information is given in Fig. 3; here RHZ refers to the so-called radiative habitable zone (applicable to both the GHZ and RVEM), which is based on the planetary climate enforcements set by both stellar components, while deliberately ignoring the orbital stability criterion regarding a possible system planet. Figure 3 indicates the inner and outer RHZ limits with the inner HZ limit defined as the maximum radial distance of the inner RHZ (red lines) and the outer HZ limit defined as the minimum radial distance of the outer RHZ (blue lines). This approach conveys the HZ region for GHZ (darkest green) and RVEM (medium green) criteria (see also Table 5). As expected the RVEM criteria produces a more generous HZ region. We also indicate the orbital stability limit (black dashed line) based on Holman & Wiegert (1999), referred to as aorb. In fact it is found that the widths of the GHZ and RVEM for Kepler-16(AB) are significantly less than for Kepler-16A (single star approach), owing to the additional criterion of orbital stability for possible system planets. Previous work by Mischna et al. (2000) argues that the HZ about a main-sequence star might be further extended if CO2 cloud coverage is assumed. In the case of the Sun, this assumption would amount to an outer limit of 2.40 AU for the hereupon defined ex- tended habitable zone (EHZ)1. Von Bloh et al. (2007) have explored the habitability around Gliese 581 with focus on the possible planet GJ 581d. They argue that the RHZ could be fur- ther extended if the atmospheric structure is determined by particularly high base pressures. Thus, the outer limit for the EHZ is not very well constrained, but could be parameterized as ǫ√L with ǫ in the likely range between 2.0 and 3.0 and L defined as stellar luminosity (in units of solar luminosity). Hence, ǫ = 2.4 corresponds to the value of Mischna et al. (2000). Results for the EHZ of Kepler-16(AB) are given in Figure 4 and Table 6. Another aspect of study is that concerning the GHZ and RVEM, we also have explored the impact of the observational uncertainties of the stellar luminosities on inner and outer 1The previous work by Mischna et al. (2000) has been superseded by more recent studies, including those given by Halevy et al. (2009), Pierrehumbert & Gaidos (2011), Wordsworth et al. (2013), and Kitzmann (2016); see also summary by Seager (2013). For example, Kitzmann (2016) argued that the heating assumed by Mischna et al. (2000) has been overestimated, thus putting the extension of the outer HZ in question. However, in the following, we will parameterize the outer limit of Mischna et al. (2000), and the significance of our results will not rely on the full extent of the HZ introduced by them. Moreover, Pierrehumbert & Gaidos (2011) argued that planetary HZs could extend to up to 10 AU for single G-type stars (or, say, about 3 AU for single K-type stars, as indicated by their Fig. 1), which is well beyond the outer limit advocated by Mischna et al. (2000). – 6 – limits of the RHZs (see Figs. 5 and 6). It is found that the uncertainty in the stellar luminosity ∆L moves the inner and outer limits of both the GHZ and the RVEM by about ±6%. Our results are summarized in Table 7. Here we also see that the inner limits of both the GHZ and RVEM are set by the additional criterion of orbital stability regarding possible circumbinary planets, referred to by Cuntz (2014) as PT habitability. Additionally, it is found that the HZ around Kepler-16A (if treated as a single star) would be significantly more extended than the HZ of Kepler-16(AB). Thus, Kepler-16B notably reduces the prospect of habitability in that system. 4. Stability Investigations for Earth-Mass Exomoons and Trojans Previously, QMC12 have exemplary case studies for the orbital stability of Earth-mass objects (i.e., Trojan exoplanet or exomoon) in the Kepler-16(AB) system. Their numerical methods were based on the Wisdom-Holman mapping technique and the Gragg-Burlisch- Stoer algorithm (Grazier et al. 1996). The resulting equations of motion were integrated forward in time for 1 million years using a fixed/initial (WH/GBS) time step. QMC12 showed that, in principle, both Trojan exoplanets and exomoons are able to exist in the Kepler-16(AB) system. Figures 7 and 8 show the results by QMC12 together with the updated system's HZs, i.e., the GHZ, RVEM, and EHZ. It is found that the orbital settings of those objects are consistent with an EHZ (with ǫ ∼< 2.2) or with the RVEM if upper limits of the stellar luminosities, consistent with the observational uncertainties, are considered. In order to better understand the dynamical domain of possible exo-Trojans, we perform additional 5,000 stability simulations using a modified version of the mercury6 integration package that is optimized for circumbinary systems (Chambers et al. 2002). In these simu- lations, we adopt the orbital parameters from Doyle et al. (2011) for the binary components and the Saturnian planet. We also consider Earth-mass objects with different initial con- ditions. Table 8 conveys the initial conditions for exomoon sample cases, which are: the semimajor axis a, eccentricity e, inclination i, argument of periastron ω, and mean anomaly M for each body. A simulation is terminated when the Earth-mass body either crosses the binary orbit or has a radial distance from the center of mass greater than 100 AU; this will be viewed as an ejection. The orbital evolution of the four bodies are evaluated on a 10 Myr timescale. The initial orbital elements are chosen using uniform distributions. The initial semimajor axis of the Earth-mass object is selected from values ranging from 0.6875 AU to 0.7221 AU (i.e., ± 0.5 Hill radii); furthermore, eccentricities are limited to 0.1 and inclinations are limited to 1◦. The initial argument of periastron and mean anomalies are selected randomly between 0◦ – 7 – and 360◦. The statistical distributions of the surviving population are shown in Figure 9 to illustrate possible correlations between parameters. Overall ∼10% of the simulations (496) are identified as stable (i.e., survived for 10 Myr) as depicted in Figure 10. By delineating the stable (cyan) and unstable (gray) points, it is seen that the stable initial conditions correspond to Trojans and are separated in relative phase from Kepler-16b by ∼60◦ to 90◦. This also appears in Figure 9 through the distribution for λ∗, the relative mean longitude. The inclinations of the orbitally stable Earth-mass objects in Fig. 9 remain uniformly distributed and thus are unlikely to affect the overall stability. Figure 11 illustrates the orbital evolution in a rotated-reference frame of two initial conditions taken from Figure 10. The panels of Fig. 11 show the first ∼1,000 years of orbital evolution, where the run in the top panel would continue in a Trojan orbit for the 10 Myr simulation time and the other run (bottom panel) evolves in a horseshoe orbit, which quickly becomes unstable. We also found that the eccentricity distribution as obtained prefers values close to circular, whereas the relative mean longitude distribution reflects, by a factor of two, more trailing orbits than preceding orbits. 5. Summary and Conclusions The purpose of our study is to continue investigating the habitable zone as well as the general possibility of Earth-mass exomoons and Trojans in Kepler-16. The binary system Kepler-16(AB) consists of a low-mass main-sequence star, a red dwarf and a circumbinary Saturnian planet. The temperatures of the two stellar components are given as 4450± 150 K and 3308± 110 K, respectively. Previously, QMC12 pursued an exploratory study about this system, indicating that based on orbital stability considerations both Earth-mass exomoons and Earth-mass Trojan planets might be possible. The aim of the present study is to offer a more thorough analysis of this system. We found the following results: (1) As previously said by QMC12, Kepler-16 possesses a circumbinary HZ; its width de- pends on the adopted climate model. Customarily, these HZs are referred to as GHZ and RVEM; the latter is also sometimes referred to as optimistic HZ (e.g., Kopparapu et al. 2013; Kaltenegger 2017). For objects of thick CO2 atmospheres including clouds, the HZ is assumed to be further extended, thus giving rise to the EHZ as proposed by Mischna et al. (2000). (2) Our work confirms earlier simulations by QMC12 that both Earth-mass exomoons and Earth-mass Trojan planets could stably orbit in that system. However, in this study, we – 8 – adopted longer timescales and also explored the distributions of eccentricity and inclinations of the Earth-mass test objects considered in our study. (3) Exomoons and Trojans, associated with the Saturnian planet, are found to be situated in the lower portion of the EHZ (i.e., ǫ ∼< 2.2). A more detailed analysis also implies that the distances of those objects may be consistent with the RVEM (i.e., optimistic HZ) if a relatively high luminosity for the stellar components is assumed (but still consistent with the uncertainty bars) or if the objects are allowed to temporarily leave the RVEM-HZ without losing habitability. The latter property is maintained if habitability is provided by a relatively thick atmosphere Williams & Pollard (2002). (4) For tutorial reasons, we also compared the HZ of the system's primary to that of the binary system. We found that the latter is reduced by 42% (GHZ) and 48% (RVEM) despite the system's increase in total luminosity given by the M-dwarf. The reason is that for the binary, the RHZ is unbalanced and it is further reduced by the additional requirement of orbital stability as pointed out previously (e.g., Eggl et al. 2013; Cuntz 2014). (5) Moreover, we pursued new stability simulations for Earth-mass objects while taking into account more general initial conditions. The attained eccentricity distribution prefers values close to circular, whereas the inclination distribution is relatively flat. The distribution in the initial relative phase indicates that the stable solutions are distributed near the co-orbital Lagrangian points, thus increasing the plausibility for the existence of those objects. Our study shows that the binary system Kepler-16(AB) has a HZ of notable extent, though smaller than implied by the single-star approach, with its extent critically depend- ing on the assumed climate model for the possible Earth-mass Trojan planet or exomoon. Thus, Kepler-16 should be considered a valuable target for future planetary search mis- sions. Moreover, it is understood that comprehensive studies of habitability should take into account additional forcings by planet host stars, such as stellar activity and strong winds expected to impact planetary conditions as indicated through analyses by, e.g., Lammer (2007), Tarter et al. (2007), Lammer et al. (2009), Kasting et al. (2014), and Kaltenegger (2017). Recent articles about the impact on stellar activity on prebiotic environmental con- ditions have been given by, e.g., Cuntz & Guinan (2016) and Airapetian et al. (2017). This work has been supported by the Department of Physics, University of Texas at Arlington (UTA). The simulations presented here were performed using the OU Supercom- puting Center for Education & Research (OSCER) at the University of Oklahoma (OU). – 9 – Furthermore, we would like to thank the anonymous referee for useful suggestions, allowing us to improve the manuscript. – 10 – REFERENCES Airapetian VS, Glocer A, Khazanov GV, Loyd ROP, France K, Sojka J, Danchi WC and Liemohn MW (2017) How hospitable are space weather affected habitable zones? the role of ion escape. Astrophysical Journal Letters 836, L3. Armstrong DJ, Osborn HP, Brown DJA, Faedi F, G´omez Maqueo Chew Y, Martin DV, Pollacco D and Udry S (2014) On the abundance of circumbinary planets. Monthly Notices of the Royal Astronomical Society 444, 1873. Baraffe I, Chabrier G, Allard F and Hauschildt PH (1998) Evolutionary models for solar metallicity low-mass stars: mass-magnitude relationships and color-magnitude diagrams. Astronomy and Astrophysics 337, 403. Chambers JE, Quintana EV, Duncan MJ and Lissauer JJ (2002) Symplectic Inte- grator Algorithms for Modeling Planetary Accretion in Binary Star Systems. Astro- nomical Journal 123, 2884. Cuntz M (2014) S-type and P-type habitability in stellar binary systems: a comprehensive approach. I. Method and applications. Astrophysical Journal 780, 14. Cuntz M (2015) S-type and P-type habitability in stellar binary systems: a comprehensive approach. II. Elliptical orbits. Astrophysical Journal 798, 101. Cuntz M and Guinan EF (2016) About exobiology: the case for dwarf K stars. Astro- physical Journal 827, 79. Doyle LR, Carter JA, Fabrycky DC, Slawson RW, Howell SB, Winn JN, Orosz JA, Prsa A, Welsh WF, Quinn SN, Latham D, Torres G, Buchhave LA, Marcy GW, Fortney JJ, Shporer A, Ford EB, Lissauer JJ, Ragozzine D, Rucker M, Batalha N, Jenkins JM, Borucki WJ, Koch D, Middour CK, Hall JR, McCauliff S, Fanelli MN, Quintana EV, Holman MJ, Caldwell DA, Still M, Stefanik RP, Brown WR, Esquerdo GA, Tang S, Furesz G, Geary JC, Berlind P, Calkins ML, Short DR, Steffen JH, Sasselov D, Dunham EW, Cochran WD, Boss Alan, Haas MR, Buzasi D and Fischer D (2011) Kepler-16: A Transiting Circumbinary Planet. Science 333, 1602. Dvorak R (1982) Planetary orbits in double star systems. Osterreichische Akademie der Wissenschaften Mathematisch-Naturwissenschaftliche Sitzungsberichte 191, 423. – 11 – Eggl S., Haghighipour N and Pilat-Lohinger E (2013) Detectability of Earth-like plan- ets in circumstellar habitable zones of binary star systems with Sun-like components. Astrophysical Journal 764, 130. Grazier KR, Newman WI, Varadi F, Goldstein DJ and Kaula WM (1996) Inte- grators for long-term solar system dynamical simulations. Bulletin of the American Astronomical Society 28, 1181. Haghighipour N and Kaltenegger L (2013) Calculating the habitable zone of binary star systems. II. P-type binaries. Astrophysical Journal 777, 166. Halevy I, Pierrehumbert RT and Schrag DP (2009) Radiative transfer in CO2-rich paleoatmospheres. Journal of Geophysical Research 114, D18112. Holman MJ and Wiegert PA (1999) Long-term stability of planets in binary systems. Astronomical Journal 117, 621. Jones BW, Sleep PN and Chambers JE (2001) The stability of the orbits of terres- trial planets in the habitable zones of known exoplanetary systems. Astronomy and Astrophysics 366, 254. Kaltenegger L (2017) How to characterize habitable worlds and signs of life. Annual Review of Astronomy and Astrophysics 55, 433. Kane SR and Hinkel NR (2013) On the habitable zones of circumbinary planetary sys- tems. Astrophysical Journal 762, 7. Kasting JF, Whitmore DP and Reynolds RT (1993) Habitable zones around main sequence stars. Icarus 101, 108. Kasting JF, Kopparapu R, Ramirez RM and Harman CE (2014) Remote life- detection criteria, habitable zone boundaries, and the frequency of Earth-like planets around M and late K stars. Proceedings of the National Academy of Sciences 111, 12641. Kirk B, Conroy K, Prsa A, Abdul-Masih M, Kochoska A, Matijevic G, Hamble- ton K, Barclay T, Bloemen S, Boyajian T, Doyle LR, Fulton BJ, Hoekstra AJ, Jek K, Kane SR, Kostov V, Latham D, Mazeh T, Orosz JA, Pepper J, Quarles B, Ragozzine D, Shporer A, Southworth J, Stassun K, Thomp- son SE, Welsh WF, Agol E, Derekas A, Devor J, Fischer D, Green G, Gropp J, Jacobs T, Johnston C, LaCourse DM, Saetre K, Schwengeler H, Toczyski J, Werner G, Garrett M, Gore J, Martinez AO, Spitzer I, Stevick – 12 – J, Thomadis PC, Vrijmoet EH, Yenawine M, Batalha N and Borucki W (2016) Kepler eclipsing binary stars. VII. The catalog of eclipsing binaries found in the entire Kepler data set. Astronomical Journal 151, 68. Kirkpatrick JD, Henry TJ and McCarthy DW Jr (1991) A standard stellar spec- tral sequence in the red/near-infrared - classes K5 to M9. Astrophysical Journal Supplement Series 77, 417. Kitzmann D (2016) Revisiting the scattering greenhouse effect of CO2 ice clouds. Astro- physical Journal Letters 817, L18. Kopparapu RK, Ramirez R, Kasting JF, Eymet V, Robinson TD, Mahadevan S, Terrien RC, Domagal-Goldman S, Meadows V and Deshpande R (2013) Habitable zones around main-sequence stars: new estimates. Astrophysical Journal 765, 131; Erratum 770, 82. Kopparapu RK, Ramirez RM, SchottelKotte J., Kasting JF, Domagal-Goldman S and Eymet V (2014) Habitable zones around main-sequence stars: dependence on planetary mass Astrophysical Journal 787, L29. Kostov VB, Orosz JA, Welsh WF, Doyle LR, Fabrycky DC, Haghighipour N, Quarles B, Short DR, Cochran WD, Endl M, Ford EB, Gregorio J, Hinse TC, Isaacson H, Jenkins JM, Jensen ELN, Kane S, Kull I, Latham DW, Lissauer JJ, Marcy GW, Mazeh T, Muller TWA, Pepper J, Quinn SN, Ragozzine D, Shporer A, Steffen JH, Torres G, Windmiller G and Borucki WJ (2016) Kepler-1647b: the largest and longest-period Kepler transiting circumbi- nary planet. Astrophysical Journal 827, 86. Lammer H (2007) M star planet habitability. Astrobiology 7, 27. Lammer H, Bredehoft JH, Coustenis A, Coustenis A, Khodachenko ML, Kaltenegger L, Grasset O, Prieur D, Raulin F, Ehrenfreund P, Yamauchi M, Wahlund J-E, Griessmeier J-M, Stangl G, Cockell CS, Kulikov YuN, Grenfell JL and Rauer H (2009) What makes a planet habitable? Astronomy and Astrophysics Reviews 17, 181. Mann AW, Gaidos E and Ansdell M (2013) Spectro-thermometry of M dwarfs and their candidate planets: too hot, too cool, or just right? Astrophysical Journal 779, 188. Mischna MA, Kasting JF, Pavlov A and Freedman R (2000) Influence of carbon dioxide clouds on early martian climate. Icarus 145, 546. – 13 – Pierrehumbert R and Gaidos E (2011) Hydrogen greenhouse planets beyond the habit- able zone. Astrophysical Journal Letters 734, L13. Popp M and Eggl S (2017) 3D climate simulations of an Earth-like circumbinary planet. 19th European Geosciences Union General Assembly EGU2017, 9885. Press WH, Flannery BP, Teukolsky SA and Vetterling WT (1986) Numerical Recipes: The Art of Scientific Computing (Cambridge: Cambridge Univ. Press). Quarles B, Musielak ZE and Cuntz M (2012) Habitability of Earth-mass planets and moons in the Kepler-16 system. Astrophysical Journal 750, 14 (QMC12). Raghavan D, Henry TJ, Mason BD, Subasavage JP, Jao W-C, Beaulieu TD and Hambly NC (2006) Two Suns in the sky: stellar multiplicity in exoplanet systems. Astrophysical Journal 646, 523. Raghavan D, McAlister HA, Henry TJ, Latham DW, Marcy GW, Mason BD, Gies DR, White RJ and ten Brummelaar TA (2010) A survey of stellar families: multiplicity of solar-type stars. Astrophysical Journal Supplement Series 190, 1. Roell T, Neuhauser R, Seifahrt A and Mugrauer M (2012) Extrasolar planets in stellar multiple systems. Astronomy and Astrophysics 542, A92. Seager S (2013) Exoplanet habitability. Science 340, 577. Selsis F, Kasting JF, Levrard B, Paillet J, Ribas I and Delfosse X (2007) Habitable planets around the star Gliese 581? Astronomy and Astrophysics 476, 1373. Shevchenko II (2017) Habitability properties of circumbinary planets. Astronomical Jour- nal 153, 273. Slawson RW, Prsa A, Welsh WF, Orosz JA, Rucker M, Batalha N, Doyle LR, Engle SG, Conroy K, Coughlin J, Gregg TA, Fetherolf T, Short DR, Wind- miller G, Fabrycky DC, Howell SB, Jenkins JM, Uddin K, Mullally F, Seader SE, Thompson SE, Sanderfer DT, Borucki W and Koch D (2011) Kepler eclipsing binary stars. II. 2165 eclipsing binaries in the second data release. Astronomical Journal 142, 160. Tarter JC, Backus PR, Mancinelli RL, Aurnou JM, Backman DE, Basri GS, Boss AP, Clarke A, Deming D, Doyle LR, Feigelson ED, Freund F, Grinspoon DH, Haberle RM, Hauck SA II, Heath MJ, Henry TJ, Hollingsworth JL, Joshi MM, Kilston S, Liu MC, Meikle E, Reid IN, Rothschild LJ, Scalo J, – 14 – Segura A, Tang CM, Tiedje JM, Turnbull MC, Walkowicz LM, Weber AL and Young RE (2007) A reappraisal of the habitability of planets around M dwarf stars. Astrobiology 7, 30. Thompson SE, Coughlin JL, Hoffman K, Mullally F, Christiansen JL, Burke CJ, Bryson S, Batalha N, Haas MR, Catanzarite J, Rowe JF, Barentsen G, Caldwell DA, Clarke BD, Jenkins JM, Li J, Latham DW, Lissauer JJ, Mathur S, Morris RL, Seader SE, Smith JC, Klaus TC, Twicken JD, Wohler B, Akeson R, Ciardi DR, Cochran WD, Barclay T, Campbell JR, Chaplin WJ, Charbonneau D, Henze CE, Howell SB, Huber D, Prsa A, Ramirez SV, Morton TD, Christensen-Dalsgaard J, Dotson JL, Doyle L, Dunham EW, Dupree AK, Ford EB, Geary JC, Girouard FR, Isaacson H, Kjeldsen H, Steffen JH, Quintana EV, Ragozzine D, Shporer A, Silva Aguirre V, Still M, Tenenbaum P, Welsh WF, Wolfgang A, Zamudio KA, Koch DG and Borucki WJ (2017) Planetary candidates observed by Kepler. VIII. A fully automated catalog with measured completeness and reliability based on data release 25. Astrophysical Journal Supplement Series submitted; arXiv:1710.06758. Underwood DR, Jones BW and Sleep PN (2003) The evolution of habitable zones during stellar lifetimes and its implications on the search for extraterrestrial life. International Journal of Astrobiology 2, 289. von Bloh W, Bounama C, Cuntz M and Franck S (2007) The habitability of super- Earths in Gliese 581. Astronomy and Astrophysics 476, 1365. Wang J, Fischer DA, Xie J-W and Ciardi DR (2014) Influence of stellar multiplic- ity on planet formation. II. Planets are less common in multiple-star systems with separations smaller than 1500 AU. Astrophysical Journal 791, 111. Wang, Zh and Cuntz M (2017) Fitting formulae and constraints for the existence of S-type and P-type habitable zones in binary systems. Astronomical Journal 154, 157. Welsh WF, Orosz JA, Short DR, Cochran WD, Endl M, Brugamyer E, Haghigh- ipour N, Buchhave LA, Doyle LR, Fabrycky DC, Hinse TC, Kane SR, Kos- tov V, Mazeh T, Mills SM, Muller TWA, Quarles B, Quinn SN, Ragozzine D, Shporer Avi, Steffen JH, Tal-Or L, Torres G, Windmiller G and Borucki WJ (2015) Kepler 453b - the 10th Kepler transiting circumbinary planet. Astro- physical Journal 809, 26. Williams DM and Pollard D (2002) Earth-like worlds on eccentric orbits: excursions beyond the habitable zone. International Journal of Astrobiology 1, 61. – 15 – Wordsworth R, Forget F, Millour E, Head JW, Madeleine J-B and Charnay B (2013) Global modelling of the early martian climate under a denser CO2 atmosphere: water cycle and ice evolution. Icarus 222, 1. Zuluaga JI, Mason PA and Cuartas-Restrepo PA (2016) Constraining the radiation and plasma environment of the Kepler circumbinary habitable-zone planets. Astro- physical Journal 818, 160. This preprint was prepared with the AAS LATEX macros v5.0. – 16 – 3600 3500 3400 3300 3200 3100 ) K ( e r u t a r e p m e T e v i t c e f f E 3000 0.4 0.35 Kepler-16B GJ 699 GJ 581 GJ 725A GJ 725Bd Empirical Formula 0.15 0.1 0.3 0.2 Stellar Mass (Solar Masses) 0.25 Fig. 1.-: Depiction of the effective temperature of Kepler-16B determined via an empirical formula given by Mann et al. (2013) that relates the mass to the effective temperature, and vice versa, for M dwarf stars. By knowing the mass of Kepler-16B, its effective temperature can be extracted, resulting in an effective temperature of 3308± 110 K. In addition, a subset of the sample of M dwarf stars is depicted, used to derive the adopted empirical formula. – 17 – Inner Boundaries: Statistical Uncertainties Moist Greenhouse Runaway Greenhouse Recent Venus - Kop13/14 - Kas93/Und03/Sel07 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 Distance from Star (AU) Outer Boundaries: Statistical Uncertainties Early Mars Maximum Greenhouse First CO2 Condensation - Kop13/14 - Kas93/Und03/Sel07 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Distance from Star (AU) Fig. 2.-: Inner and outer HZ limits for Kepler-16A (single star approach) while comparing two different determination methods. We also include information on the inherent statis- tical uncertainties of those based on Press et al. (1986) (see also Table 3). The blue data correspond to the inner and outer HZ boundaries as expected from utilizing the method of Kasting et al. (1993) with updates by Underwood et al. (2003) and Selsis et al. (2007). Con- versely, the red data correspond to the inner and outer HZ limits as expected from utilizing the method specified by Kopparapu et al. (2013, 2014). 120 240 120 – 18 – Kepler-16 GHZ RHZ 90 1 0.75 0.5 0.25 270 Kepler-16 RVEM RHZ 90 1 0.75 0.5 0.25 60 300 60 240 300 270 30 0 330 30 0 330 180 150 210 180 150 210 Fig. 3.-: Depiction of the RHZ for the GHZ and RVEM criteria based on methods given by Cuntz (2014, 2015) and Wang & Cuntz (2017). In both plots the red and blue lines cor- respond to the inner and outer RHZ limits with the inner HZ limit defined as the maximum radial distance of the inner RHZ (red lines) and the outer HZ limit defined as the minimum radial distance of the outer RHZ (blue lines). This approach produces the conventional HZ region for GHZ (darkest green) and RVEM (medium green) criteria. As expected the RVEM criteria produces a more generous HZ region as shown. Lastly, the black dashed line repre- sents the orbital stability limit, calculated using the formula provided by Holman & Wiegert (1999) for P-type orbits, in which bodies exterior to that line are orbitally stable while bodies interior to that line are orbitally unstable. – 19 – Kepler-16 Binary EHZ RHZ (variable Epsilon) 120 60 90 1.5 1.25 1 0.75 0.5 0.25 180 150 210 30 0 330 240 300 270 Fig. 4.-: Different outer boundaries (blue lines) of the EHZ resulting from the different epsilon values ranging from ǫ = 2.0 (innermost blue line) to ǫ = 3.0 (outermost blue line). A median value of ǫ = 2.5 has been chosen for our definition of the EHZ akin to Mischna et al. (2000), which is also adopted for our analysis in the subsequent Figs. 7, 8, and 10 and depicted as the lightest green regions. – 20 – Kepler-16 GHZ RHZ with Different Luminosities 120 60 90 1 0.75 0.5 0.25 180 150 210 30 0 330 240 300 270 Kepler-16 RVEM RHZ with Different Luminosities 180 150 210 120 60 90 1 0.75 0.5 0.25 240 300 270 30 0 330 Fig. 5.-: Depiction of the inner and outer boundaries of the GHZ and RVEM, while utilizing the upper and lower bounds of the stellar luminosities to illustrate how the uncertainty in the luminosity affects the determination of the HZs. In both plots the inner and outer HZ limits are shown in red and blue, respectively, with the inner sets of red and blue lines corresponding to the lower bound luminosity and the outer sets of red and blue lines corresponding to the upper bound luminosity. As expected, the upper bound luminosity shifts the GHZ and RVEM limits outward while the lower bound luminosity shifts those limits inward. – 21 – Binary RHZ of Kepler-16 120 60 90 1 0.75 0.5 0.25 150 180 210 30 0 330 GHZ min GHZ max RVEM min RVEM max Stab 240 300 270 Fig. 6.-: Similar to Fig. 5, as this figure combines the inner and outer boundaries for the GHZ and RVEM criteria while also incorporating the upper and lower luminosity bounds; its emphasis is to illustrate the extents of the achievable HZs based on luminosity and HZ criteria specification. Additionally, the black dashed line represents the orbital stability limit. The blue and red dotted lines correspond to the minimum possible inner limits (associated with the lower bound luminosity) for the GHZ and RVEM, respectively. The blue and red solid lines correspond to the maximum possible outer limits (associated with the upper bound luminosity) for GHZ and RVEM, respectively. – 22 – Kepler-16 with Captured S-Type Exomoon 150 180 210 150 180 210 120 60 90 1.25 1 0.75 0.5 0.25 240 300 270 Kepler-16 with Trojan Exomoon 120 60 90 1.25 1 0.75 0.5 0.25 240 300 270 30 0 330 30 0 330 Kepler-16A Kepler-16B Kepler-16b test body Kepler-16A Kepler-16B Kepler-16b test body Fig. 7.-: Illustration of previous results by QMC12 with updated HZ regions. (a) Depiction of an S-type captured Earth-mass exomoon (black in QMC12); the primary and secondary stars (orange and red in QMC12, respectively) and the Saturnian planet Kepler-16b are also given (magenta in QMC12). (b) Depiction of a possible Trojan exomoon in a rotating reference frame (black in QMC12). The darkest green region represents the GHZ, the medium green region represents the RVEM, and the lightest green region represents the EHZ. The dashed yellow line represents the outer limit of the GHZ if the stellar luminosities are assumed at their upper limits as informed by the observational uncertainties. – 23 – Kepler-16 with Stable Co-formed Earth-mass Exomoon 120 60 90 1.25 1 0.75 0.5 0.25 150 180 210 30 0 330 240 300 270 Kepler-16 with Unstable Co-formed Earth-mass Exomoon 150 180 210 120 60 90 1.5 1.25 1 0.75 0.5 0.25 240 300 270 30 0 330 Kepler-16A Kepler-16B Kepler-16b test body Kepler-16A Kepler-16B Kepler-16b test body Fig. 8.-: Illustration of the previous results by QMC12 with updated HZ regions. (a) De- piction of a stable S-type coformed Earth-mass exomoon (black in QMC12); the primary and secondary stars (orange and red in QMC12, respectively) and the Saturnian planet Kepler- 16b are also given (magenta in QMC12). (b) Depiction of an unstable S-type coformed Earth-mass exomoon (black in QMC12). See Fig. 7 for information on the color coding of the HZs. – 24 – e ) . g e d ( i ) . g e d ( * λ 0.08 0.06 0.04 0.02 1.0 0.8 0.6 0.4 0.2 100 50 0 −50 −100 0.696 0.712 0.704 a (AU) 0.720 0.02 0.06 0.08 0.04 e 0.2 0.4 0.6 0.8 1.0 −100 i (deg.) 100 0 −50 50 λ * (deg.) Fig. 9.-: Distributions of the initial semimajor axis a, eccentricity e, inclination i, and relative mean longitude λ∗, given as λ∗ = λ⊕ − λ16B that produces a stable Earth-mass co- orbital planet in Kepler-16. These initial conditions are chosen relative to the center-of-mass of the stellar components. – 25 – Fig. 10.-: Illustration of the starting locations of stable (cyan) and unstable (gray) initial conditions out of 5,000 trials. These simulations differ from QMC12 as the relative phase between the binary and planetary orbit is now taken into account, where the positive x-axis is taken to be the line-of-sight. See Fig. 7 for information on the color coding of the HZs. – 26 – Kepler-16 with Stable Earth-mass Trojan 120 60 90 1.25 1 0.75 0.5 0.25 30 0 330 180 150 210 240 300 270 Kepler-16 with Unstable Earth-mass Trojan 120 60 90 1.25 1 0.75 0.5 0.25 180 150 210 30 0 330 240 300 270 Fig. 11.-: Examples of orbital evolution (magenta) of a stable (top) and unstable (bottom) Earth-mass planet co-orbiting with Kepler-16b. These orbits are shown in a rotated-reference frame depicting the relative motions with Kepler-16b to illustrate Trojan (top) and horseshoe (bottom) configurations. See Fig. 7 for information on the color coding of the HZs. – 27 – Table 1. Stellar and Planetary Parameters of Kepler-16 Parameter Valuea Distance (pc) FB/FA M1 (M⊙) M2 (M⊙) Teff,1 (K) Teff,2 (K) R1 (R⊙) R2 (R⊙) Pb (d) ab (AU) eb Mp (MJ) ap (AU) ep ∼ 61 0.01555 ± 0.0001 0.6897 ± 0.0035 0.20255 ± 0.00066 4450 ± 150 3308 ± 110 0.6489 ± 0.0013 0.22623 ± 0.00059 41.079220 ± 0.000078 0.22431 ± 0.00035 0.15944 ± 0.00061 0.333 ± 0.016 0.7048 ± 0.0011 0.0069 ± 0.001 Note. - aData as provided by Doyle et al. (2011) and reported by QMC12, except for Teff,2 and R2, which have been determined in this study. All parameters have their usual meaning. – 28 – Table 2. Percentage Uncertainty of Kepler-16 Parameters Parameter % Uncertainty FB/FA M1 (M⊙) M2 (M⊙) Teff,1 (K) Teff,2 (K) R1 (R⊙) R2 (R⊙) 0.64 0.51 0.33 3.37 3.33 0.20 0.26 – 29 – Table 3. Single Star Habitable Zone Limits Habitable Zone Limit Kas93/Und03 Kop1314 HZ Type Recent Venus Runaway Greenhouse Water Loss First CO2 Condensation Maximum Greenhouse Early Mars 0.299 0.334 0.376 0.592 0.708 0.746 0.308 0.390 0.402 ... 0.723 0.766 RVEM (in) GHZ (in) ... ... GHZ (out) RVEM (out) Note. - Kas93: Kasting et al. (1993), Und03: Underwood et al. (2003), Kop1314: Kopparapu et al. (2013, 2014) – 30 – Table 4. Statistical Uncertainties Habitable Zone Limit ... Kas93/Und03 Kop1314 Min-Max Statis Min-Max Statis Recent Venus Runaway Greenhouse Water Loss First CO2 Condensation Maximum Greenhouse Early Mars 1.97 % 1.93 % 1.98 % 1.93 % 2.21 % 2.15 % 2.51 % 2.41 % 2.51 % 2.44 % 2.59 % 2.51 % 3.34 % 3.22 % 4.14 % 3.98 % 4.15 % 5.67 % 4.40 % 4.26 % 4.39 % 5.97 % ... ... Note. - For references, see comments of Table 3. Min-Max means that the minimum/maximum values for the luminosities and effective temperatures are applied. Statis means adequately applied statistical uncertainty propagation. – 31 – Table 5. GHZ and RVEM RHZs of Binary System Reference Distance ... GHZ RVEM Relevance (AU) (AU) ... Inner RHZ Limit, innermost Inner RHZ Limit, outermost Outer RHZ Limit, innermost Outer RHZ Limit, outermost Orbital Stability Limit 0.368 0.444 0.704 0.783 0.510 0.285 0.361 0.747 0.827 0.510 No Yes Yes No Yes Note. - The RHZ bounds have previously been referred to as RHLs (Cuntz 2014). Here the innermost and outermost points of these limits are reported, which are of different rel- evance for setting the respective RHZ. – 32 – Table 6. EHZ of Kepler-16(AB) ǫ ... 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 EHZ (AU) 0.765 0.801 0.837 0.873 0.910 0.946 0.982 1.018 1.055 1.091 1.127 Table 7. Comparison of Habitable Zone Limits Type Single Star Binary System Approach ... ... GHZ RVEM GHZ GHZ (L−) GHZ (L+) RVEM RVEM (L−) RVEM (L+) (AU) (AU) (AU) (AU) (AU) (AU) (AU) (AU) RHZin RHZout aorb ∆HZ Type 0.390 0.723 ... 0.308 0.763 ... 0.333 0.455 ... ... 0.444 0.704 0.510 0.194 PT 0.419 0.662 0.510 0.152 PT 0.470 0.746 0.510 0.236 PT 0.361 0.747 0.510 0.237 PT 0.341 0.702 0.510 0.192 PT 0.381 0.792 0.510 0.282 PT – 3 3 – Note. - L+ and L− indicate L± ∆L, respectively, with variations in Teff and R simultaneously applied to both stellar components (see Table 1). ∆HZ indicates the width of the HZ with con- sideration of the orbital stability limit, if applicable. PT conveys that the P-type HZ is truncated due to the additional requirement of orbital stability. – 34 – Table 8. Initial Conditions for Exomoon Sample Cases Publication Type QMC12 This Work Kepler-16(AB) Kepler-16(AB)b Stable Retrograde Stable Trojan Stable Prograde Unstable Prograde Kepler-16(AB) Kepler-16(AB)b Stable Trojan Unstable Trojan a (AU) 0.22431 0.7048 0.619 0.7048 0.715 0.721 0.22431 0.7048 0.7096 0.6902 e 0.15944 0.0069 0.13 0.0069 0 0 0.15944 0.0069 0.0088 0.0651 i (◦) 0 0 180 0 0 0 0 0.3079 0.8175 0.0795 ω (◦) 0 180 180 180 180 180 M (◦) 180 180 180 240 180 180 318 263.464 −171.114 −211.49 35.272 154.849 37.499 124.235 Note. - Initial conditions in terms of orbital elements for the binary (Kepler-16(AB)), the Saturnian planet (Kepler-16b), and the possible Earth-mass exomoon. These orbital elements can be used to reproduce our new results (Fig. 11) and the previous results of QMC12 (Figs. 7 and 8).
1909.00706
1
1909
2019-09-02T13:19:53
The Observability Of Vortex-Driven Spiral Arms In Protoplanetary Disk: Basic Spiral Properties
[ "astro-ph.EP" ]
Some circumstellar disks are observed to show prominent spiral arms in infrared scattered light or (sub-)millimeter dust continuum. The spirals might be formed from self-gravity, shadows, or planet-disk interactions. Recently, it was hypothesized that massive vortices can drive spiral arms in protoplanetary disks in a way analogous to planets. In this paper, we study the basic properties of vortex-driven spirals by the Rossby Wave Instability in 2D hydrodynamics simulations. We study how the surface density contrast, the number, and the shape of vortex-driven spirals depend on the properties of the vortex. We also compare vortex-driven spirals with those induced by planets. The surface density contrast of vortex-driven spirals in our simulations are comparable to those driven by a sub-thermal mass planet, typically a few to a few tens of Earth masses. In addition, different from the latter, the former is not sensitive to the mass of the vortex. Vortex-driven spiral arms are not expected to be detectable in current scattered light observations, and the prominent spirals observed in scattered light in a few protoplanetary disks, such as SAO 206462 (HD 135344B), MWC 758, and LkHa 330, are unlikely to be induced by the candidate vortices in them.
astro-ph.EP
astro-ph
Draft version September 4, 2019 Preprint typeset using LATEX style emulateapj v. 12/16/11 9 1 0 2 p e S 2 . ] P E h p - o r t s a [ 1 v 6 0 7 0 0 . 9 0 9 1 : v i X r a THE OBSERVABILITY OF VORTEX-DRIVEN SPIRAL ARMS IN PROTOPLANETARY DISK: BASIC SPIRAL PROPERTIES Pinghui Huang CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China University of Chinese Academy of Sciences, Beijing 100049, China and Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA Department of Physics & Astronomy, University of Victoria, Victoria, BC, Canada Ruobing Dong Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA Hui Li, Shengtai Li CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China and CAS Center for Excellence in Comparative Planetology, Hefei 230026, China Jianghui Ji Draft version September 4, 2019 ABSTRACT Some circumstellar disks are observed to show prominent spiral arms in infrared scattered light or (sub-)millimeter dust continuum. The spirals might be formed from self-gravity, shadows, or planet-disk interactions. Recently, it was hypothesized that massive vortices can drive spiral arms in protoplanetary disks in a way analogous to planets. In this paper, we study the basic properties of vortex-driven spirals by the Rossby Wave Instability in 2D hydrodynamics simulations. We study how the surface density contrast, the number, and the shape of vortex-driven spirals depend on the properties of the vortex. We also compare vortex-driven spirals with those induced by planets. The surface density contrast of vortex-driven spirals in our simulations are comparable to those driven by a sub-thermal mass planet, typically a few to a few tens of Earth masses. In addition, different from the latter, the former is not sensitive to the mass of the vortex. Vortex-driven spiral arms are not expected to be detectable in current scattered light observations, and the prominent spirals observed in scattered light in a few protoplanetary disks, such as SAO 206462 (HD 135344B), MWC 758, and LkHα 330, are unlikely to be induced by the candidate vortices in them. Subject headings: instabilities -- hydrodynamics -- protoplanetary disks -- submillimeter: planetary systems 1. INTRODUCTION In recent years, many circumstellar disks were resolved by high angular resolution infrared and (sub-)millimeter observations (Brogan et al. 2015; Long et al. 2018; An- drews et al. 2018; Avenhaus et al. 2018; Liu et al. 2016; Huang et al. 2018a,b). These disks present a large diver- sity in morphology, showing rings, cavities, spiral arms, and dust crescents, likely produced by planets (Ou et al. 2007; Fung et al. 2014; Zhu & Stone 2014; Jin et al. 2016; Liu et al. 2018; van der Marel et al. 2018; Jin et al. 2019). Specifically, several disks present both dust crescents in continuum emission and spirals in scattered light, such as MWC 758 (Isella et al. 2010; Boehler et al. 2018; Dong et al. 2018b; Grady et al. 2012; Benisty et al. 2015), SAO 206462 (HD 135344B) (van der Marel et al. 2016; P´erez et al. 2014; Muto et al. 2012; Garufi et al. 2013), LkHα 330 (Isella et al. 2013; Akiyama et al. 2016; Uyama et al. 2018), V1247 Ori (Ohta et al. 2016; Kraus et al. 2017), HD 142527 (Casassus et al. 2013; Canovas et al. 2013; Avenhaus et al. 2014), and AB Aur (Hashimoto et al. 2011; Tang et al. 2017). 1 Crescents in continuum emission have been proposed to be dust trapping in vortices generated by Rossby Wave Instability (Lovelace et al. 1999; Li et al. 2000, 2001, RWI). The RWI can be triggered by local Rossby wave trapped in a steep density bump, either at the edges of planet-opened gaps or at a dead zone edge (Lin 2014; Miranda et al. 2017). Spiral features have been pro- posed to be the density waves excited by planets (Dong et al. 2015b) or by the gravitational instability (Kratter & Lodato 2016; Dong et al. 2015a). They may also be produced by shadows (Benisty et al. 2017, 2018; Mon- tesinos & Cuello 2018). van der Marel et al. (2016) and Cazzoletti et al. (2018) highlighted the connection be- tween the two-arm spirals and the dust crescent seen in the SAO 206462 disk. They proposed that the dust cres- cent is a massive vortex, and it is exciting the observed spirals in a way similar to a planet with a similar mass. In this paper, we study the properties of spirals induced by a large vortex generated by the RWI. §2 shows the numerical setup of our hydrodynamic models. §3 and §4 present hydrodynamic results and the properties of vortices induced spiral arms. We will summarize our conclusions in §5. 2 2. SIMULATIONS 0/GM(cid:63) (cid:1)1/2 K,0 =(cid:0)R3 We carry out 2D global hydrodynamical simula- tions in polar coordinate using the LA-COMPASS code (Los Alamos COMPutional Astrophysics Simulation Suite) (Li et al. 2005, 2008; Fu et al. 2014) to follow the evolution of gas in a protoplanetary disk. The code units of mass, length, and time are the star mass M(cid:63) = 1 M(cid:12), R0 = 50 au, and the dynamical timescale at R0, i.e. τdyn,0 = Ω−1 , respectively. The equa- tion of the state is locally isothermal Pg = c2 s Σg, where Pg is the vertically integrated pressure, and cs and Σg are the sound speed and the surface density of gas, respec- tively. The behaviours of density waves in linear regime are insensitive to the EOS (Dong et al. 2011b; Miranda & Rafikov 2019b), while the spiral shock location and wave amplitude depend on the EOS in non-linear regime (Goodman & Rafikov 2001; Dong et al. 2011a). The −1/4 distribution of cs and Σg are cs (R) = cs,0 (R/R0) −1, with Σ0 normalized by the total and Σg = Σ0 (R/R0) disk mass Mdisk. The scale height profile is chosen as H/R = 0.05 (R/R0)1/4. The inner and outer boundaries of our simulations are Rmin = 0.2R0 = 10 au and Rmax = 4.5R0 = 225 au. We used the outflow and inflow boundary conditions for the inner and outer boundaries, respectively. The simu- lations are carried on a linear-grid. The resolutions are 2048 × 3072 along the R and Φ (azimuthal) directions. The scale height at R0 is resolved by 25 cells along the radial direction. To avoid the disturbance of planet-induced density waves, we choose the dead zone edge model (no planet in simulations) to generate vortices and spirals (Miranda et al. 2017). The dead zone is a region with low vis- cosity and consequently weak accretion. Viscosity in PPDs come from turbulence, which are likely produced by instabilities, such as the Magnetorotational Instabil- ity (Balbus & Hawley 1998). A sufficiently ionized disk threaded by magnetic fields in the ideal MHD regime can be MRI active. However, at tens of AU the midplane of PPDs are subject to non-ideal MHD effects such as the Ohmic resistivity, Hall effect, and in particular Am- bipolar diffusion. As a result the MRI is not active, and PPDs have a "dead zone" at the midplane with weak tur- bulence, thus low viscosity (Gammie 1996; Bai & Stone 2011; Bai 2014, 2015). The outer disk beyond the dead zone may be MRI active due to cosmic ray ionization (Ar- mitage 2011). Here we adopt the toy model of the dead zone edge as described by Reg´aly et al. (2011): 2 1 − tanh α (R) = α0 − α0 − αDZ where α0 = 10−3 and αDZ = 10−5 are the Shakura & Sunyaev viscosity parameters outside and inside the dead zone. The viscosity transition is located at RDZ = 1.5R0 and the width of the transition region is ∆DZ. ∆DZ (1) Due to the low viscosity within the dead zone, gas grad- ually piles at the dead zone edge to form a density bump. If ∆DZ < 2H, the RWI will be triggered to generate large-scale anticyclonic vortices (Reg´aly et al. 2011). In this paper, we set up five models to investigate how the viscosity transition region width (SD vs H-SH vs Q- (cid:20) (cid:18) R − RDZ (cid:19)(cid:21) , SH) and the disk self-gravity (SD vs HM-10MJ vs HM- 30MJ) affect vortex-driven spiral arms. The setup of the models is in Table 1. We run 3000 orbits for each model in our hydrodynamic simulations. TABLE 1 List of Models Name SD H-SH Q-SH HM-10MJ HM-30MJ ∆DZ Self-Gravity Disk Mass (MJ) H H/2 H/4 H H No No No Yes Yes 3 3 3 10 30 SD: Standard H-SH: Half-Scale Height Q-SH: Quarter-Scale Height HM-10MJ: High Disk Mass with Mdisk = 10 MJ HM-30MJ: High Disk Mass with Mdisk = 30 MJ 3. THE VORTICES IN THE MODELS 3.1. The Standard Model (SD) R3 d dR (cid:0)R4Ω2(cid:1) < 0, where κ is the epicyclic Figure 1 shows the evolution of the gas surface density in Model SD at various epochs. Due to the low viscosity in the dead zone region, the gas piles up at the dead zone edge (∼ 1.3R0). Prior to the violation of the Rayleigh cri- terion (κ2 ≡ 1 frequency), the RWI condition (Lovelace et al. 1999) is violated first. The RWI is then triggered, and gener- ates large scale vortices (Li et al. 2000, 2001; Ono et al. 2016, 2018). The density bump in Model SD becomes Rossby wave unstable at 300 orbits (upper left panel), then breaks into three mode-3 vortices. They then merge into a single vortex at about 600 orbits. There are two pairs of spirals connected with the edges of the vortex (e.g., at 1000 orbits). The spirals protect the vortex from being smeared out by the background Keplerian flow (Li et al. 2001). Before a vortex is smeared out, the vortex grows stronger by accreting gas (1600 orbits). A vortex is part of the disk, and it generates spiral density waves by perturbing the velocity in the back- ground flow (Bodo et al. 2005; Heinemann & Papaloizou 2009). Vortices generate density waves by perturbing the background flow and producing sound waves, which are sheared by the Keplerian rotation. There is no planet and no disk self-gravity in Model SD, so the wave excita- tion mechanism is unrelated to gravitational interaction between vortices and the disk -- the vortices are mass- less. This is different from how planets excite density waves, in which case a planet is a point mass and launches density waves at Lindblad resonances through gravita- tional disk-planet interactions (Goldreich & Tremaine 1979). van der Marel et al. (2016) showed the ALMA Band 7 continuum observations on SAO 206462. They found that a previously observed asymmetric ring in SAO 206462 is consisted of an inner ring and an outer bump, and proposed the possibility that the latter is a vortex driving the observed spiral arms in scattered light (Muto et al. 2012). Cazzoletti et al. (2018) pre- sented multi-wavelengths ALMA observations of the dust continuum from the SAO 206462 circumstellar disk. They resolved the disk into an axisymmetric ring and 3 Fig. 1. -- The surface density profile normalized by Σinit at various epochs for Model SD. Σinit is the initial surface density for each model, i.e. the surface density for each model at 0 orbit. The horizontal axis is the radius, and the vertical axis is the azimuth (the center of the first generation vortex is at Φc = 0). We define the vortex center where the velocity in the vortex equals to the local Keplerian velocity. The centers of vortices are very close to the their density maxima. a crescent. They claimed that the crescent is a massive vortex with several MJ and it launches the observed spi- rals arms in a way similar to the interactions between a disk and a massive planet. This interpretation is incon- sistent with spirals driven by vortices. In addition to the origin, the contrast of spirals induced by vortices is too low to be detectable in scattered light. We will explain it in Section 4.3. The density waves exert torques onto the vortex. At 2000 orbits, the negative torque by the outer den- sity waves surpasses the positive torque by the in- ner density waves due to the radial asymmetry in the disk (Paardekooper et al. 2010). As a result, the vortex migrates inward. The vortex in Model SD forms at about 1.3R0, launching spiral waves and migrating to ∼ 1.0R0 at the end of our simulation (3000 orbits). The migra- tion of a vortex is similar to the type I migration of a planet (Paardekooper et al. 2010). After that, gas keep piling up at the dead zone edge, triggering the RWI to generate a secondary generation of vortices (the outer vortex seen at 2000 orbits and later). However, we do not see significant spirals excited by the second genera- tion of vortices due to their small velocity perturbation on the background and a larger aspect ratio (Surville & Barge 2015). At 3000 orbits, there are two shallow gaps at R ∼ 0.7 and 0.9 opened by the spirals driven by the first genera- tion vortex in the dead zone. Dong et al. (2017, 2018a) and Bae et al. (2017) showed that a Super Earth opens multiple gaps in low viscosity disks with α (cid:46) 10−4. It is possible that the vortex in Model SD opens multiple gaps in a similar way. Based on Meheut et al. (2013), the RWI is triggered by the local Rossby waves trapped in the disk. The Rossby wave is a potential vorticity (a.k.a. vortensity) wave. The vortensity is conserved along the streamlines in an inviscid and barotropic fluid. In the RWI, the enthalpy perturbation ψ ≡ δP/Σ is similar with the Schrodinger equation (Lovelace & Romanova 2014). When Rossby waves escape the potential well, they transfer into spiral density waves propagating inward and outward. This phenomenon is similar with the "tunnel through" in quantum mechanics (Li et al. 2001). The spiral density waves cause the dissipation and shrinking of the vortex. The primary vortex at 3000 orbits in Figure 1 is weaker than that at 2500 orbits due to this effect. 3.2. The Effects of Viscosity Transition Width Figure 2 show the vortices and spirals in Model SD, H-SH and Q-SH. We pick the orbital frames prior to 4 Fig. 2. -- The surface density contrast defined by Equation 3 of the models with various viscosity transition widths (SD, H-SH and Q-SH). The two rows are the same expect for the mark ups and symbols. The black solid points mark the centers of the vortices. In the top row, the spirals are marked using letters (primary spirals A-H; secondary spirals X-Y). In the bottom row, the cross symbols mark the expected wake of the primary spirals in the linear density wave theory described by Equation 2 , and the vertical dashed lines mark the radial locations 5 scale heights away from the vortex centers. The azimuthal cuts at these locations will be shown in Figure 4. the production of the secondary generation vortices. By comparing models with different radial widths of the vis- cosity transition region, we find that the Rossby wave unstable density bump is replenished faster when the re- gion shrinks. The growth rate of the RWI becomes higher while the lifetime of the resulting vortices is shorter. The primary vortex in Model SD survives for at least 3000 or- bits. The vortices in Model H-SH and Q-SH only survive for 2000 and 1000 orbits, respectively. After that, the vortices are smeared out by the background Keplerian flow into axisymmetric density bumps. Different from Model SD, there is no secondary vortex in H-SH and Q- SH. Based on Surville & Barge (2013), more elongated vortices migrate more slowly, and smaller vortices are more compressible. The asymmetries between the inner and outer spirals in smaller vortices lead faster migrated rate. 3.3. The Effects of Self-Gravity and Disk Mass Lin & Papaloizou (2011) found that disk self-gravity can stabilize high mode number vortices. This is con- sistent with our high mass self-gravitating models -- at 1600 orbits, there are two mode-2 vortices in Model HM- 10MJ, four mode-4 vortices in Model HM-30MJ, but only one vortex (mode-1) in Model SD (Figure 3). These vor- tices in HM-10MJ and HM-30MJ at 1600 orbits survive to the end of the simulations (3000 orbits). Comparing Model SD, HM-10MJ and HM-30MJ, the effect of disk self-gravity on the vortices becomes larger when the disk becomes more massive. The Toomre Q ≡ κcs πGΣ (Toomre 1964) reaches its maximum at the location of the vortex in all models. Q is about 20 and 10 in Models HM-10MJ and HM-30MJ at 1600 orbits, respectively, while Q > 50 in the other three viscosity transition models by the end of the simulations (3000 orbits). Based on Lovelace & Hohlfeld (2012), disk self-gravity is important for the RWI only when Q < R/H. In our models, the disk self-gravity becomes significant when Q < R/H ∼20 at the locations of the vortices. 4. PROPERTIES OF SPIRALS 4.1. Number of Spirals Figures 2 and 3 show the general morphology and con- trast of the vortices and spirals at various epochs for different models. There are 4, 6, and 8 spirals connected with the vortex in Model SD (A-D), H-SH (A-F), and Q-SH (A-H), respectively. Li et al. (2001) showed that in the case of four spirals, two of them (A and C) are shocks, while the other two (B and D) are rarefaction waves. The vortices of Model H-SH and Q-SH are more elongated than the vortex in Model SD. When a vortex is too long (aspect ratio (cid:38) 9 in our simulations), it can not communicate along the Φ direction. An elongated vortex will either lead to multiple density peaks or pro- duction of multiple spirals. Multiple density peaks can be the outcome of incomplete vortex merger in the ini- 5 Fig. 3. -- The surface density contrast defined by Equation 3 of the models with self-gravity (HM-10MJ and HM-30MJ) comparing with Model SD (no self-gravity). The black solid points mark the centers of the vortices. The vertical dashed lines mark the radial locations 5 scale heights away from the vortex centers; the azimuthal cuts at these locations will be shown in Figure 4. tial formation stage. The spirals E and F in H-SH are produced from spirals D and B. The vortex in Q-SH has two density peaks (i.e. two cores), so that it excites eight spirals. In Model HM-10MJ and HM-30MJ, spiral arms induced by different vortices start to interfere with each other (Figure 3), making identifying each unique arm difficult. In the past, a sub-thermal mass planet was thought to generate a pair of spiral arms (one on each side of the planet's orbit; Ogilvie & Lubow 2002). However, based on the simulations of Bae & Zhu (2018a) and Miranda & Rafikov (2019a) (see also Zhu et al. 2015; Bae et al. 2017; Bae & Zhu 2018b; Fung & Dong 2015; Lee 2016), the angular momentum flux of the primary density waves induced by a sub-thermal mass planet can be transferred into higher-order density waves (secondary, tertiary, etc) during their propagation. In our simulations, the sec- ondary spirals (spirals X and Y) are not originated at the vortices, and they are likely the higher order spirals generated by the primary spirals (spirals A and C) as a result of the constructive interference among different azimuthal modes (Bae & Zhu 2018a; Miranda & Rafikov 2019a). 4.2. Shape of Spirals He 1 1−α+β Re (cid:17)(cid:27) (cid:16) R Re (cid:17)−α(cid:21) (cid:26)(cid:16) R (cid:17)1+β(cid:20) Based on the linear spiral density wave theory (Goldre- ich & Tremaine 1979; Rafikov 2002; Muto et al. 2012), the shape of spirals induced by a planet in the linear regime is: Φ (R) = Φe − sgn(R−Re) 1+β − 1 × 1−α+β (2) where the orbital angular frequency Ω ∝ R−α and the sound speed cs ∝ R−β (α = 3 4 in our simula- tions). Re and Φe are the radial and azimuthal locations of the planet. When the mass of a planet is larger than the thermal mass, the linear theory breaks down and there is no analytic theory for the shape of the spirals. −(cid:16) 1 1+β − 1 2 and β = 1 Equation 2 is marked using cross symbols in Figure 2 for Model SD, H-SH, and Q-SH. The shapes of the spi- rals A-D in these models are consistent with Equation 2. However, as the multiple spiral arms are likely interact- ing with each others, there are deviations in the shapes of the spirals E-H. As mentioned before, spirals induced by a vortex are similar to those induced by a sub-thermal mass planet. After the excitation of the density waves, the specific excitation mechanism no longer affects the waves as they propagate. The spirals X and Y are not connected with the vortices; it is harder to compare their shapes with the linear theory due to the uncertainty in the origin of the spirals. 4.3. Contrast of Spirals The contrast of density waves in gas surface density at a given radius R is defined as: δΣg (R) (cid:104)Σg (R)(cid:105)Φ ≡ Σg,peak (cid:104)Σg (R)(cid:105)Φ − 1 (3) where (cid:104)Σg (R)(cid:105)Φ is the 2π azimuthally averaged surface density at R. Figure 4 shows the azimuthal cuts of the scaled surface density at 5 scale heights from the vortex centers, and highlights how the contrast depends on the width of the viscosity transition (panels b and e) and self- gravity and disk mass (panels c and f ). We note that the outer spirals are stronger than the inner ones. To compare the spirals induced by vortices and plan- ets, we show the same azimuthal cuts of the spiral waves driven by a 4 M⊕ planet in an H/R = 0.05 and α = 5×10−5 disk (Dong et al. 2017) in panels (a) and (d). The difference in the contrast between the inner and outer spi- rals is smaller in the case of planet-induced spirals than vortex-induced spirals. This leads to higher migration rates of vortices than planets (Paardekooper et al. 2010). In planet-disk interaction, we can define the thermal mass as Mth = (H/R)3 M(cid:63) (Goodman & Rafikov 2001), which is about 50 M⊕ at 1.3R0 in our simulations. For planet-induced spiral arms, the peak of the scaled sur- 2 ≈ Mplanet/Mth face density contrast is (cid:16) δΣg (cid:17)(cid:16) HX (cid:17) 1 (cid:104)Σg(cid:105)Φ 6 Fig. 4. -- Azimuthal density profiles across the spirals at a radial distance 5 scale heights away from the vortex center (the vertical dashed lines in Figures 2 and 3). The horizontal axis is the azimuth. The vertical axis is the surface density contrast normalized by (X/H)1/2, where X ≡ R − Rc, and Rc is the radial location of the vortex center. The upper and lower rows are for the inner (at R = Rc − 5H) and outer spirals (at R = Rc + 5H), respectively. The black lines in panels (a) and (d) are the azimuthal cuts of the spiral waves generated by a 4 M⊕ planet in the H/R = 0.05 planet-disk interaction simulation in Dong et al. (2017). in the linear regime (Mplanet (cid:28) Mth), where X ≡ R− Rc and Rc is the radius of the vortex peak (Goodman & Rafikov 2001; Dong et al. 2011b; Duffell & MacFadyen 2012). The surface density contrasts of the spirals pro- duced by the vortices in our simulations are similar to those induced by a sub-thermal mass planet. As an ex- ample, the contrast of the spirals wakes in Model SD are similar with that of the spirals induced by a 4 M⊕ planet (∼ 0.1 Mth). The density bumps in H-SH and Q-SH are more Rossby wave unstable than that in SD due to the steeper vis- cosity transition compared with SD. The vortensity of vortices in H-SH and Q-SH are about 1.5 and 2.0 times the vortensity of the primary vortex in SD. Because the vortensity represents the rotation of vortices with respect to the background flow, the velocity perturbations and contrast of the vortices in H-SH and Q-SH are larger and stronger than those in SD. arms is smaller than order unity (corresponding to spirals driven by sub-Mth planets), it is difficult to detect the spirals at tens of AU in systems at 140 pc under the angular resolution currently achievable in NIR scattered light imaging. 1. A massless vortex can generate spiral arms as its velocity field compresses the background gas to produce density waves. This is different from how planets generate spiral arms through gravitational planet-disk interactions. 5. CONCLUSIONS We have run hydrodynamic simulations to investigate the properties of spiral arms induced by a vortex. Here are the main conclusions of this paper: Model HM-10MJ and HM-30MJ each produces multi- ple long-lasting vortices. Specifically, the vortex located at 0◦ has a shorter rotation period and a lower aspect ratio, the contrasts of its spirals are higher than those induced by the other more extended vortices (Bodo et al. 2005; Surville & Barge 2015). While the disk self-gravity compress the gas, the contrast of vortex-induced spirals is insensitive to the disk self-gravity and the mass of the vortex. The vortex in Model HM-10MJ and Model HM- 30MJ weights ∼ 0.8 MJ and ∼ 1.2 MJ, respectively (5.0 and 7.5 times the thermal mass); however, the peak con- trasts of their density waves are within a factor of 2 from those in Model SD, which has a massless vortex (see also §4.3). The contrast of the spirals in our models is in between 0.1 and 0.3. Juh´asz et al. (2015) and Dong & Fung (2017) showed that when the surface density contrast of spiral 2. The surface density contrast of vortex-driven spi- rals is from 0.1 to 0.3 in our simulations. This is similar to the spirals produced by a planet with a mass on the order of 0.1 thermal mass, equivalent to a few to a few tens of Earth masses at tens of AU under typical conditions. Such arms are too weak and should not be expected to be detectable in current direct imaging observations. The promi- nent spirals observed in protoplanetary disks such as MWC 758, SAO 206462, LkHα 330 are unlikely to be driven by the candidate vortex seen in them. 3. The disk self-gravity becomes important to the de- velopment of the Rossby Wave Instability when Q (cid:46) R/H, in which case it stabilizes the high mode number vortices produced by the RWI. The surface density contrast of vortex-driven spirals is insensi- tive to the disk self-gravity. Specifically, the con- trasts of the spiral arms driven by the vortices in Model HM-30MJ, which weight about 1.2 MJ (7.5 times the disk thermal mass), are still comparable to those driven by sub-thermal mass planets. 4. A vortex generates at least 4 spiral arms (two on each side) that are directly connected with itself, and the shape of these spirals are consistent with the linear density wave theory. More elongated vor- tex can generate more than 4 spirals. A vortex can produce secondary spirals in the disk. ACKNOWLEDGMENTS 7 We thank the referee for the detailed comments that improved the presentation of this paper significantly. We are grateful to Jaehan Bae, Tomohiro Ono, Min-Kai Lin and Chong Yu for useful discussions. This work is sup- ported by the National Natural Science Foundation of China (grant Nos. 11773081, 11661161013, 11633009 and 11873097), the CAS Interdisciplinary Innovation Team, the Strategic Priority Research Program on Space Science, the Chinese Academy of Sciences, Grant No. XDA15020302 and the Foundation of Minor Planets of Purple Mountain Observatory. We also acknowledge the support by a LANL/CSES project. This work was par- tially performed at the Aspen Center for Physics, which is supported by National Science Foundation grant PHY- 1607611. REFERENCES Akiyama, E., Hashimoto, J., baobabu Liu, H., et al. 2016, The Goldreich, P., & Tremaine, S. 1979, Astrophysical Journal, 233, Astronomical Journal, 152, 222 Andrews, S. M., Huang, J., P´erez, L. M., et al. 2018, The Astrophysical Journal Letters, 869, L41 Armitage, P. J. 2011, Annual Review of Astronomy and Astrophysics, 49 857 552, 793 Goodman, J., & Rafikov, R. 2001, The Astrophysical Journal, Grady, C., Muto, T., Hashimoto, J., et al. 2012, The Astrophysical Journal, 762, 48 Avenhaus, H., Quanz, S. P., Schmid, H. M., et al. 2014, The Hashimoto, J., Tamura, M., Muto, T., et al. 2011, The Astrophysical Journal, 781, 87 Astrophysical Journal Letters, 729, L17 Avenhaus, H., Quanz, S. P., Garufi, A., et al. 2018, The Heinemann, T., & Papaloizou, J. 2009, Monthly Notices of the Astrophysical Journal, 863, 44 Bae, J., & Zhu, Z. 2018a, The Astrophysical Journal, 859 -- . 2018b, The Astrophysical Journal, 859, 119 Bae, J., Zhu, Z., & Hartmann, L. 2017, The Astrophysical Journal, 850 Royal Astronomical Society, 397, 52 Huang, J., Andrews, S. M., Dullemond, C. P., et al. 2018a, The Astrophysical Journal Letters, 869, L42 Huang, J., Andrews, S. M., P´erez, L. M., et al. 2018b, The Astrophysical Journal Letters, 869, L43 Bai, X.-N. 2014, The Astrophysical Journal, 791, 137 -- . 2015, The Astrophysical Journal, 798, 84 Bai, X.-N., & Stone, J. M. 2011, The Astrophysical Journal, 736, Isella, A., Natta, A., Wilner, D., Carpenter, J. M., & Testi, L. 2010, The Astrophysical Journal, 725, 1735 Isella, A., P´erez, L. M., Carpenter, J. M., et al. 2013, The Balbus, S. A., & Hawley, J. F. 1998, Reviews of modern physics, Jin, S., Isella, A., Huang, P., et al. 2019, The Astrophysical Benisty, M., Juhasz, A., Boccaletti, A., et al. 2015, Astronomy & Jin, S., Li, S., Isella, A., Li, H., & Ji, J. 2016, The Astrophysical Benisty, M., Stolker, T., Pohl, A., et al. 2017, Astronomy & Juh´asz, A., Benisty, M., Pohl, A., et al. 2015, Monthly Notices of Benisty, M., Juh´asz, A., Facchini, S., et al. 2018, Astronomy & Kratter, K., & Lodato, G. 2016, Annual Review of Astronomy Bodo, G., Chagelishvili, G., Murante, G., et al. 2005, Astronomy Kraus, S., Kreplin, A., Fukugawa, M., et al. 2017, The Astrophysical Journal, 775, 30 Journal, 881, 108 Journal, 818, 76 144 70, 1 Astrophysics, 578, L6 Astrophysics, 597, A42 Astrophysics, 619, A171 & Astrophysics, 437, 9 Journal, 853, 162 Journal Letters, 808, L3 Astrophysics, 556, A123 191 Boehler, Y., Ricci, L., Weaver, E., et al. 2018, The Astrophysical Brogan, C., P´erez, L., Hunter, T., et al. 2015, The Astrophysical Astrophysical Journal, 551, 874 Canovas, H., M´enard, F., Hales, A., et al. 2013, Astronomy & Astrophysical Journal, 533, 1023 Casassus, S., van der Plas, G., Perez, S., et al. 2013, Nature, 493, 624, 1003 Cazzoletti, P., van Dishoeck, E., Pinilla, P., et al. 2018, Astronomy & Astrophysics, 619, A161 Dong, R., & Fung, J. 2017, The Astrophysical Journal, 835, 38 Dong, R., Hall, C., Rice, K., & Chiang, E. 2015a, The Astrophysical Journal Letters, 812, L32 the Royal Astronomical Society, 451, 1147 and Astrophysics, 54, 271 Astrophysical Journal Letters, 848, L11 Lee, W.-K. 2016, The Astrophysical Journal, 832, 166 Li, H., Colgate, S., Wendroff, B., & Liska, R. 2001, The Li, H., Finn, J., Lovelace, R., & Colgate, S. 2000, The Li, H., Li, S., Koller, J., et al. 2005, The Astrophysical Journal, Li, H., Lubow, S., Li, S., & Lin, D. N. 2008, The Astrophysical Lin, M.-K. 2014, Monthly Notices of the Royal Astronomical Journal Letters, 690, L52 Society, 437, 575 Lin, M.-K., & Papaloizou, J. C. 2011, Monthly Notices of the Royal Astronomical Society, 415, 1426 Dong, R., Li, S., Chiang, E., & Li, H. 2017, The Astrophysical Liu, H. B., Takami, M., Kudo, T., et al. 2016, Science Advances, Journal, 843, 127 -- . 2018a, The Astrophysical Journal, 866, 110 Dong, R., Rafikov, R. R., & Stone, J. M. 2011a, The Astrophysical Journal, 741, 57 2, e1500875 Liu, S.-F., Jin, S., Li, S., Isella, A., & Li, H. 2018, The Astrophysical Journal, 857, 87 Long, F., Pinilla, P., Herczeg, G. J., et al. 2018, The Dong, R., Rafikov, R. R., Stone, J. M., & Petrovich, C. 2011b, Astrophysical Journal, 869, 17 The Astrophysical Journal, 741, 56 Lovelace, R., & Hohlfeld, R. 2012, Monthly Notices of the Royal Dong, R., Zhu, Z., & Whitney, B. 2015b, The Astrophysical Astronomical Society, 429, 529 Dong, R., Liu, S.-y., Eisner, J., et al. 2018b, The Astrophysical Astrophysical Journal, 513, 805 Duffell, P. C., & MacFadyen, A. I. 2012, The Astrophysical 46, 041401 Fu, W., Li, H., Lubow, S., Li, S., & Liang, E. 2014, The Astrophysical Journal Letters, 795, L39 Fung, J., & Dong, R. 2015, The Astrophysical Journal Letters, Journal, 835, 118 Fung, J., Shi, J.-M., & Chiang, E. 2014, The Astrophysical 875, 37 Gammie, C. F. 1996, The Astrophysical Journal, 457, 355 Garufi, A., Quanz, S. P., Avenhaus, H., et al. 2013, Astronomy & Astrophysics, 560, A105 Lovelace, R., Li, H., Colgate, S., & Nelson, A. 1999, The Lovelace, R., & Romanova, M. 2014, Fluid Dynamics Research, Meheut, H., Lovelace, R., & Lai, D. 2013, Monthly Notices of the Royal Astronomical Society, 430, 1988 Miranda, R., Li, H., Li, S., & Jin, S. 2017, The Astrophysical Miranda, R., & Rafikov, R. R. 2019a, The Astrophysical Journal, -- . 2019b, The Astrophysical Journal Letters, 878, L9 Montesinos, M., & Cuello, N. 2018, Monthly Notices of the Royal Astronomical Society: Letters, 475, L35 Journal, 809, 93 Journal, 860, 124 Journal, 755, 7 815, L21 Journal, 782, 88 8 Muto, T., Grady, C., Hashimoto, J., et al. 2012, The Astrophysical Journal Letters, 748, L22 Surville, C., & Barge, P. 2013, in EPJ Web of Conferences, Vol. 46, EDP Sciences, 05002 Ogilvie, G., & Lubow, S. 2002, Monthly Notices of the Royal Surville, C., & Barge, P. 2015, Astronomy & Astrophysics, 579, Astronomical Society, 330, 950 A100 Ohta, Y., Fukagawa, M., Sitko, M. L., et al. 2016, Publications of Tang, Y.-W., Guilloteau, S., Dutrey, A., et al. 2017, The the Astronomical Society of Japan, 68 Ono, T., Muto, T., Takeuchi, T., & Nomura, H. 2016, The Astrophysical Journal, 823, 84 Ono, T., Muto, T., Tomida, K., & Zhu, Z. 2018, The Astrophysical Journal, 864, 70 Astrophysical Journal, 840, 32 Toomre, A. 1964, The Astrophysical Journal, 139, 1217 Uyama, T., Hashimoto, J., Muto, T., et al. 2018, The Astronomical Journal, 156, 63 van der Marel, N., Cazzoletti, P., Pinilla, P., & Garufi, A. 2016, Ou, S., Ji, J., Liu, L., & Peng, X. 2007, The Astrophysical The Astrophysical Journal, 832, 178 Journal, 667, 1220 van der Marel, N., Williams, J. P., & Bruderer, S. 2018, The Paardekooper, S.-J., Lesur, G., & Papaloizou, J. C. 2010, The Astrophysical Journal Letters, 867, L14 Astrophysical Journal, 725, 146 Zhu, Z., Dong, R., Stone, J. M., & Rafikov, R. R. 2015, The P´erez, L. M., Isella, A., Carpenter, J. M., & Chandler, C. J. 2014, Astrophysical Journal, 813, 88 Zhu, Z., & Stone, J. M. 2014, The Astrophysical Journal, 795, 53 The Astrophysical Journal Letters, 783, L13 Rafikov, R. R. 2002, The Astrophysical Journal, 572, 566 Reg´aly, Z., Juh´asz, A., S´andor, Z., & Dullemond, C. 2011, Monthly Notices of the Royal Astronomical Society, 419, 1701
1210.4836
3
1210
2012-11-17T16:23:54
Re-Evaluating WASP-12b: Strong Emission at 2.315 micron, Deeper Occultations, and an Isothermal Atmosphere
[ "astro-ph.EP", "astro-ph.SR" ]
We revisit the atmospheric properties of the extremely hot Jupiter WASP-12b in light of several new developments. First, new narrowband (2.315 micron) secondary eclipse photometry that we present here, which exhibits a planet/star flux ratio of 0.45% +/- 0.06 %, corresponding to a brightness temperature of 3640 K +/- 230 K; second, recent Spitzer/IRAC and Hubble/WFC3 observations; and third, a recently observed star only 1" from WASP-12, which has diluted previous observations and which we further characterize here. We correct past WASP-12b eclipse measurements for the presence of this object, and we revisit the interpretation of WASP-12b's dilution-corrected emission spectrum. The resulting planetary emission spectrum is well-approximated by a blackbody, and consequently our primary conclusion is that the planet's infrared photosphere is nearly isothermal. Thus secondary eclipse spectroscopy is relatively ill-suited to constrain WASP-12b's atmospheric abundances, and transmission spectroscopy may be necessary to achieve this goal.
astro-ph.EP
astro-ph
Accepted to ApJ: 2012 Oct 17 Preprint typeset using LATEX style emulateapj v. 5/2/11 2 1 0 2 v o N 7 1 . ] P E h p - o r t s a [ 3 v 6 3 8 4 . 0 1 2 1 : v i X r a RE-EVALUATING WASP-12b: STRONG EMISSION AT 2.315 µm, DEEPER OCCULTATIONS, AND AN ISOTHERMAL ATMOSPHERE Ian J. M. Crossfield1,2, Travis Barman3, Brad M. S. Hansen2, Ichi Tanaka4, Tadayuki Kodama4 Accepted to ApJ: 2012 Oct 17 ABSTRACT We revisit the atmospheric properties of the extremely hot Jupiter WASP-12b in light of several new developments. First, new narrowband (2.315 µm) secondary eclipse photometry that we present here, which exhibits a planet/star flux ratio of 0.45 %± 0.06 %, corresponding to a brightness temper- ature of 3640 K ± 230 K; second, recent Spitzer/IRAC and Hubble/WFC3 observations; and third, a recently observed star only 1" from WASP-12, which has diluted previous observations and which we further characterize here. We correct past WASP-12b eclipse measurements for the presence of this object, and we revisit the interpretation of WASP-12b's dilution-corrected emission spectrum. The resulting planetary emission spectrum is well-approximated by a blackbody, and consequently our primary conclusion is that the planet's infrared photosphere is nearly isothermal. Thus secondary eclipse spectroscopy is relatively ill-suited to constrain WASP-12b's atmospheric abundances, and transmission spectroscopy may be necessary to achieve this goal. Subject headings: infrared: stars -- planetary systems -- stars: individual (WASP-12, Bergfors-6) -- stars: multiple -- techniques: photometric -- techniques: spectroscopic -- eclipses 1. INTRODUCTION Transiting extrasolar planets allow the exciting possi- bility of studying the intrinsic physical properties of these planets. The latest new frontier to emerge is the detailed study of molecular chemistry in the atmospheres of these planets, many of which exist in intensely irradiated en- vironments. Recent years have seen rapid strides in this direction, with measurements of precise masses and radii, detection of secondary eclipses and phase curves and the start of ground-based spectroscopy (Redfield et al. 2008; Swain et al. 2010; Bean et al. 2010). Based on observed day/night temperature contrasts (e.g., Cowan & Agol 2011), atmospheric circulation patterns (Knut- son et al. 2009), and atmospheric chemistry (Stevenson et al. 2010; Madhusudhan et al. 2011) these planets' at- mospheres are likely to be quite different from anything previously known. 1.1. Introducing the WASP-12 System A prime example is the transiting Hot Jupiter WASP- 12b, which is one of the largest and hottest transiting planets known (Hebb et al. 2009; Chan et al. 2011; Ma- ciejewski et al. 2011). The planet is significantly over- inflated compared to standard interior models (Fort- ney et al. 2007), though its radius and age can be ex- plained by an appropriate dynamical history involving an initially eccentric orbit and subsequent interior dis- sipation of tidal torques (Ibgui et al. 2011). Radial ve- locity measurements associated with the initial transit discovery and the first occultation observation both sug- gested WASP-12b had a nonzero eccentricity (Hebb et al. 1 Max-Planck Institut fur Astronomie, Konigstuhl 17, D- 69117, Heidelberg, Germany; [email protected] 2 Department of Physics & Astronomy, University of Califor- nia Los Angeles, Los Angeles, CA 90095, USA 3 Lowell Observatory, 1400 West Mars Hill Road, Flagstaff, AZ 86001, USA 4 Subaru Telescope, National Astronomical Observatory of Japan, 650 North A'ohoku Place, Hilo, HI 96720, USA 2009; L´opez-Morales et al. 2010). However, subsequent orbital characterization via timing of secondary eclipses (Croll et al. 2011; Campo et al. 2011; Cowan et al. 2012) and further radial velocity measurements (Husnoo et al. 2011) set an upper limit on the eccentricity of ∼ 0.03(1σ). The 2.315 µm narrow band eclipse we present here is also consistent with a circular orbit. Due to its close proximity to its host star the planet is thought to be significantly distorted and may even be un- dergoing Roche lobe overflow (Li et al. 2010). Such over- flow, if verified, would be the first evidence of the tidal inflation instability (Gu et al. 2003). Possible evidence for the overflow scenario has come (1) from HST/COS UV spectra taken during transit (Fossati et al. 2010), which show tentative evidence of a deeper transit with earlier ingress than observed in the optical (Hebb et al. 2009), (2) from a tentative detection of an extended Ks band secondary eclipse duration (Croll et al. 2011), which could be interpreted as an opaque accretion stream or disk, and (3) from Spitzer/IRAC phase curve observa- tions of WASP-12b, which detect ellipsoidal variations from the planet at 4.5 µm at a level consistent with a planet filling (or overfilling) its Roche lobe (Cowan et al. 2012). However: (1) there is no evidence for an ex- tended occultation duration in Spitzer/IRAC observa- tions (Campo et al. 2011) or in the 2.315 µm narrowband eclipse we present here; (2) degeneracies between ellip- soidal variations, thermal phase variations, and instru- mental systematics prevent an unambiguous determina- tion of WASP-12b's geometry from the Spitzer observa- tions (Cowan et al. 2012); and (3) recent HST/WFC3 secondary eclipse spectroscopy suggest that WASP-12b is not substantially distorted (Swain et al. 2012). WASP-12b is intensely irradiated by its host star, mak- ing the planet one of the hottest known and giving it a favorable ((cid:38) 10−3) NIR planet/star flux contrast ra- tio; its atmosphere has quickly become one of the best- studied outside the Solar System. The planet's large size, 2 Crossfield et al. low density, and high temperature motivated a flurry of optical, (L´opez-Morales et al. 2010), NIR (Croll et al. 2011), and mid-infrared (Campo et al. 2011) secondary eclipse photometry has been interpreted to reflect an at- mosphere with an unusual carbon to oxygen (C/O) ratio greater than one (Madhusudhan et al. 2011). Subsequent ground-based observations (Zhao et al. 2012; Crossfield et al. 2012) and the recent WFC3 1.1 -- 1.7 µm spectrum (Swain et al. 2012) are consistent with these earlier mea- surements and the C/O> 1 model, but the 2.315 µm eclipse we present here is inconsistent (at > 3σ) with such models. In addition, under the so-called "null hy- pothesis" (i.e., a spherical planet) of Cowan et al. (2012) the IRAC 4.5 µm secondary eclipse is significantly deeper than the previous measurement (Campo et al. 2011), sug- gesting less absorption by CO and weakening the case for a high C/O ratio. As yet transmission spectroscopy (which determines atmospheric opacity at a planet's limb via multi- wavelength transit measurements Seager & Sasselov 2000) has so far been limited for this system. Opti- cal transit measurements show some disagreement (Hebb et al. 2009; Chan et al. 2011; Maciejewski et al. 2011), which makes interpretation difficult. Spitzer/IRAC tran- sit observations suggest that WASP-12b's radius may be greater at 3.6 µm than at 4.5 µm (Cowan et al. 2012), but only if the planet is much more prolate (Rlong/Rp = 1.8) than suggested by WFC3 observations (3σ upper limit of 1.7; Swain et al. 2012). Under the null hypothesis of Cowan et al. (2012), the transit radius is larger at 4.5 µm (as expected from atmospheric models). WASP-12b's low density and high temperature ensure that this planet will continue to be a target for future efforts in this direc- tion; if (as we suggest) the planet's atmosphere is in fact nearly isothermal at the pressures probed in secondary eclipse, transmission spectroscopy may be the only hope for constraining WASP-12b's atmospheric composition. Thus significant uncertainties remain in the interpre- tation of the current ensemble of atmospheric measure- ments. At the moment this situation is typical even for the best-characterized systems (Madhusudhan & Seager 2010) because (a) broadband photometry averages over features caused by separate opacity sources and (b) at- mospheric models have many more free parameters than there are observational constraints. When properly cali- brated, spectrally resolved measurements can break some of these degeneracies. Such results can test the interpre- tation of photometric observations at higher resolution, and can more precisely refine estimates of atmospheric abundances, constrain planetary temperature structures, and provide deeper insight into high-temperature exo- planetary atmospheres. These goals provided the mo- tivation for our earlier ground-based spectroscopy of WASP-12b (Crossfield et al. 2011) and serve as the im- petus for the analysis presented here. 1.2. Paper Outline This paper presents new secondary eclipse observations of WASP-12b's emission in a narrow band centered at 2.315 µm, our detection and characterization of a cool star (which we call Bergfors-6) with high surface gravity near WASP-12, a correction of past eclipse measurements for the dilution caused by Bergfors-6, and our interpre- tation of WASP-12b's atmospheric emission. We describe our secondary eclipse observations and ini- tial data reduction in Sec. 2. As described in Sec. 3 we fit numerous model light curves to the data, select the sta- tistically optimal combination of parameters to use, and present the results of this eclipse. In Sec. 4 we describe our analysis of Bergfors-6's properties, and in Sec. 5 we use the results of this analysis to correct past transits and occultations of WASP-12b. In Sec. 6 we discuss our analysis of WASP-12b's corrected emission spectrum and provide updated constraints on the planet's bolometric luminosity. Finally, we conclude and suggest relevant possibilities for followup in Sec. 7. 2. SUBARU/MOIRCS NARROWBAND TIME-SERIES PHOTOMETRY 2.1. Summary of Observations We described recently the first tentative detection of emission from WASP-12b via spectroscopy at the 3 m NASA Infrared Telescope Facility (IRTF) (Crossfield et al. 2012). However, our precision was strongly lim- ited by chromatic and time-dependent slit losses result- ing from the use of a single, narrow (3") slit. We sub- sequently obtained time on the Multi-Object InfraRed Camera and Spectrograph (MOIRCS; Ichikawa et al. 2006; Suzuki et al. 2008) at Subaru Observatory to con- duct multi-object occultation spectroscopy of WASP- 12b. A coolant leak at Subaru caused damage that pre- vented us from obtaining spectroscopy, so we operated the instrument in imaging mode using a custom nar- rowband filter. This filter (NB2315) is centered at ap- proximately 2.315 µm with a width at half maximum of 27 nm5, and so is very well suited to probe the strong absorption feature predicted to lie at this wavelength by models used to infer a high C/O ratio (Madhusudhan et al. 2011, their Figure 1). We observed one secondary eclipse of WASP-12b on 14 Dec 2011 (UT). The start of observations was de- layed by instrument problems, but we managed to begin about half an hour before ingress and observed contin- uously thereafter. We observed at a position angle of 330◦ and read out frames in correlated double sampling (CDS) mode with a constant integration time of 21 s per frame, using a readout speed of 8 and two dummy reads (to suppress a known, variable-bias effect; Katsuno et al. 2003). These readout parameters result in substantial overhead penalties, and we averaged only one frame per 61 s over our 6.5 hr of observations (which cover an air- mass range of 1.6−1.02−1.3). We recorded 388 frames in total. Conditions were nearly photometric, with stellar flux variations of 1-2 % apparent. Following standard practices for high-precision pho- tometry (e.g., de Mooij & Snellen 2009; Rogers et al. 2009) we defocussed the telescope to spread the starlight over more pixels, thereby increasing observing efficiency and reducing the effect of residual flat fielding errors. The instrumental seeing improved throughout the night, and to avoid any substantially nonlinear detector re- sponse we added additional defocus to the telescope sev- eral times. Because the Subaru autoguider was inoper- ative we had to periodically apply manual offsets to the telescope tracking. The tracking was rather poor and de- 5 A transmission profile of the NB2315 filter is available upon request from T.K. Re-evaluating WASP-12b 3 et al. 2004)6. Using a set of flat frames taken with typi- cal counts ranging from 2,000 -- 22,000 ADU, we compute the median linearity correction coefficients (a1, . . . , a4) of Vacca et al. (2004)'s Eq. 20: Cnl =(cid:0)1 + a1x + a2x2 + a3x3 + a4x4(cid:1)−1 (1) (where Cnl is the ratio of an ideally linear signal to the measured signal) to be a1, a2, a3, a4 = (0.00347574,−0.00436064,−0.00111471, 0.00048908) for Chip 1, and a1, a2, a3, a4 = (0.0063929,−0.0139614, 0.00548463,−0.00081818) for Chip 2. We define x as the measured ADU counts divided by 104 to avoid very small coefficients. We then iteratively apply the correction algorithm in Vacca et al.'s Eqs. 21-26, while further requiring that Cnl is always ≥ 1. Convergence typically occurs within 4-5 iterations. The nonlinearity correction is critical in our analysis: it changes our relative photometry by as much as 0.5% in some frames, and it slightly reduces our final, residual photometric RMS from 0.234% to 0.232%. At this point in the analysis substantial scattered and/or background light remains: we remove this by scal- ing and subtracting a median-combined set of median- normalized, dithered, dark-subtracted sky images. We follow these procedures independently for data from both channels; requiring an identical level of sky subtraction in both channels does not significantly change our results. We extract photometry using our own aperture pho- tometry package7, which uses bilinear interpolation to account for partial pixels while conserving flux. In each frame we extract subregions around each star, perform 1D cross-correlations to measure relative stellar motions, and interpolate over hot pixels, stuck pixels, and any pixels more than 6σ discrepant from their mean value. We then recenter the photometric apertures and perform standard aperture photometry. In imaging mode MOIRCS offers a roughly 4'×7' field of view split equally over two 20482 HAWAII-2 detec- tors, which allows several comparison stars to be fit into the WASP-12 field of view. Using stars more than about 1.8 mag fainter than WASP-12 decreases our final precision. Our large photometric apertures also require us to avoid choosing comparison stars with nearby companions. This leaves five comparison stars: 2MASS stars 06302437+2937293 and 06303222+2937347 (on Chip 1) and 06302377+2939118, 06301801+2939204, and 06302280+2938338 (WASP-12 is on Chip 2). Our fi- nal results are consistent (though of lower precision) if we use fewer comparison stars or use comparison stars falling only on a single detector. We examine the results from photometric apertures of various sizes and ultimately use target and inner and outer sky apertures with diameters of 39, 47, and 72 pixels (4.6", 5.5", and 8.4"). This choice minimizes the root mean square (RMS) of the residuals to our model fits; the final RMS is 0.232%. 2.3. Instrumental Systematics 6 A Python implementation of our MOIRCS nonlinearity cor- rection algorithm is available from the primary author's website. 7 Available from the primary author's website. Fig. 1. -- Instrumental trends during our observations. From top to bottom: Chip 1 and 2 electronics box temperatures, rela- tive x and y motions, telescope focus encoder setting, WASP-12 raw flux, WASP-12 relative flux, ratio of median sky background (pre-calibration) in Chips 1 and 2, and airmass. The dotted lines indicate the four points of contact corresponding to a circular orbit with our best-fit secondary eclipse center. The vertical dashed line corresponds to the onset of the anomalous trend apparent in the sky background and relative photometry: we exclude all data after this in our final analysis. spite our corrections we observed image drifts as large as 1.2" (10 pixels); however, subsequent software develop- ment at Subaru has improved the tracking in the absence of the autoguider. The temperature of both detectors (as reported by the CHIPBOX FITS header keywords) in- creased from 76.2 K to a constant 77.0 K over the first 1.5 − 2 hr. All these instrumental trends are shown in Figure 1, but we ultimately find that they do not signif- icantly affect our photometry. 2.2. Initial Data Reduction MOIRCS splits its field of view across two detectors, and we reduce the data from each detector indepen- dently. We calibrate the raw frames following the stan- dard MOIRCS reduction prescription, which proceeds as follows. MOIRCS returns the UT date and time at the beginning and end of each exposure in its FITS header. We convert these to BJDT DB for our subsequent analysis (Eastman et al. 2010). We dark-subtract each frame and divide the result by the stack median of a set of dark- subtracted dome flats. Next, we correct our data for the intrinsic nonlinearity of infrared detector arrays (Vacca 76.276.677.077.4Chip 1 Box [K]76.276.677.077.4Chip 2 Box [K]100x [pix]1555y [pix]0.200.240.280.32Focus [mm]0.960.981.00Raw flux0.4760.4800.484Relative flux0.940.960.98Sky1 / Sky20.400.450.500.550.600.650.70Orbital Phase1.11.31.5Airmass 4 Crossfield et al. We plot several variable instrumental parameters, along with the absolute and relative photometry of WASP-12, in Figure 1. One variable dominates in terms of its impact on our photometry: the curious trend in the ratio of the median sky background background mea- sured in the two detectors, which begins an anomalous excursion as WASP-12 crosses the meridian (only 10- 15 min after egress) before later stabilizing at a new level. The relative photometry shows a qualitatively sim- ilar trend superimposed on a secondary eclipse (visible in the raw data). We also see this trend when dividing stellar photome- try from Chip 2 (excluding WASP-12) by photometry from Chip 1, and we even see it (at a lower ampli- tude) when comparing multiple reference stars on Chip 2 against each other; it is thus a field-dependent effect. Because the trend begins just as WASP-12 crosses the meridian, we hypothesize that some loose component in the telescope or instrument settled in response to the change in the direction of the gravity vector. One possible culprit in this scenario is our narrow- band filter, whose spectral transmission profile depends on the angle of incidence of incoming light. To first order, increasing the angle of incidence translates the transmission profile to shorter wavelengths. The filter profile intersects a particularly strong telluric absorption (CH4) bandhead; from the vendor-supplied characteriza- tion data for our filter we estimate that a shift in the filter's angle of incidence of ∼ 10◦ could induce a pho- tometric variation of the magnitude observed. However, no strong sky emission features are seen at these wave- lengths, so this scenario still has difficulty explaining the observed variation in the sky background. Regardless, subsequent MOIRCS multi-object spectroscopic data do not show this anomalous trend, a fact consistent with our hypothesis that the trend's presence is somehow related to the NB2315 filter. Whatever the cause of this anomalous trend, so long as we restrict our analysis to times before orbital phase 0.5675 (the vertical dashed bar in Figure 1) our results change by less than 1.5σ no matter which comparison stars we choose. This choice leaves pre- and post-eclipse baselines which are rather short. We explored ways to use our entire data set by using the sky background trend as a decorrelation parameter (see Section 3 below), but such analyses resulted in larger fit residuals with substan- tially higher correlations on long timescales. We there- fore proceed by excluding the later data, while acknowl- edging the existence of this poorly-understood system- atic effect in MOIRCS narrowband imaging data. 3. 2.315 µm NARROWBAND SECONDARY ECLIPSE We now present our analysis of the narrowband pho- tometry discussed in the preceding setion. In Section 3.1 we describe the process of selecting an optimal model for our data and of fitting this model to the data. In Section 3.2 we describe the primary result of the fitting process: a 2.315 µm eclipse depth significantly discrepant from previous predictions. Then in Section 3.3 we show that the occultation we detect has a duration and time of center consistent with that expected for WASP-12b on a circular orbit. 3.1. Fitting to the Data We fit our photometric time series with the following relation, representing a relative secondary eclipse light curve subjected to systematic effects: 1 + J(cid:88)  Fi = f0 (1 + d(cid:96)i) cjvij (2) j=1 The symbols are: Fi, the relative flux measured at timestep i; f0, the true relative flux; (cid:96)i, the flux in an occultation light curve scaled to equal zero out of eclipse and −1 inside eclipse; d, the normalized depth of sec- ondary eclipse; vij, the J state vectors (i.e., image mo- tions, sky background, airmass, orbital phase, or low- order polynomials of these quantities) exhibiting a lin- early perturbative effect on the instrumental sensitivity; and cj, the coefficients for each state vector. Experience shows that the choice of instrumental model is of crucial importance in extracting the most accurate system parameters from transit and occulta- tion observations (e.g., Campo et al. 2011). We therefore explore a large region of model parameter space by fit- ting our photometry using many different combinations of state vectors and a fixed secondary eclipse time and duration. We then use the Bayesian Information Crite- rion (BIC8) to choose which of these many models best represents our data. To do this we first assign uncertain- ties to each data point equal to the RMS of the residuals to an eclipse fit with no additional decorrelation vari- ables. We find that the BIC-minimizing model includes a transit light curve and a linear function of time, but no additional parameters. The model with the next-best BIC (3.2 units higher) also include a linear function of the x position of the stars on the detector; both this model and a model including no decorrelation parame- ters (∆BIC = 10.7) return secondary eclipse depths con- sistent with that of our optimal model. We then again fit our preferred instrumental model to the data, but now allowing the secondary eclipse cen- ter, duration, and depth to vary while holding fixed the scaled semimajor axis (a/R∗) and the orbital inclination at the values listed in Table 1. We assess the uncertain- ties on the best-fit parameters using both the Markov Chain Monte Carlo and prayer bead (as described in Winn et al. 2008, ; see also Jenkins et al. et al. 2002) approaches. The two sets of parameter distributions are quite consistent, which suggests correlated noise does not strongly affect our photometry. In both cases the re- sulting parameter distributions are unimodal, symmet- ric, approximately normal, and (excepting the standard correlation between occultation depth and baseline flux) uncorrelated. In the following we quote only the MCMC results, which provide substantially denser sampling of the posterior distributions than the prayer-bead results. 3.2. Initial Narrowband Eclipse Depth We plot the data, our best-fit (linear baseline) model, and the residuals in Figure 2. The best-fit secondary eclipse depth is 0.41 % ± 0.05 %; note that the true planet/star flux ratio is ∼ 10% greater than this, as we describe in Section 5 and Table 2. We find shal- lower or deeper best-fit eclipse depths (ranging from 8 Bayesian Information Criterion (BIC) = χ2 + k ln N , where k is the number of free parameters and N the number of data points. Re-evaluating WASP-12b 5 WASP-12b: 2.315 µm Narrowband Secondary Eclipse Parameters TABLE 1 Parameter P a/R∗ Tc,e Toffset T58 e cos ω e sin ω FP /F∗ (observed) FP /F∗ (corrected) TB,2.315 Units days -- BJDTDB s min -- -- -- -- K Value 1.091423 2455910.9090 ± 0.0013 3.14 7 ± 110 s +0.00006 ± 0.00091 179.6 ± 4.5 min 0.011 ± 0.013 0.41 % ± 0.05 % 0.45 % ± 0.06 % 3640 K ± 230 K Reference Hebb et al. (2009) Hebb et al. (2009) This work This work This work This work This work This work This work This work Fig. 3. -- RMS from binning the residuals shown in Figure 2 over increasing numbers of data points (solid black line). This curve is exceeds the N−1/2 expectation from white noise (dashed line) by increasing amounts on successively longer timescales. The timescale of ingress or egress is indicated by the vertical dotted line. The residuals to the best fit have a relative RMS of 0.232%. Photon (target + sky) noise considerations pre- dict a typical per-frame precision of 0.06%, so our per- formance is comparable to that obtained with other NIR secondary eclipse photometry (e.g., Croll et al. 2011). Figure 3 shows that our residuals bin down somewhat more slowly than N−1/2, indicating a moderate level of correlated residuals. The largest residuals occur at a "bump" in the light curve at orbital phase 0.47 -- 0.48. One possible explana- tion is that this excursion is caused by telluric variations that are not entirely common mode across the MOIRCS field of view. An alternate explanation would be a short- term flare from Bergfors-6. We apply Difference Image Analysis (Bramich 2008)9 to two images generated by co-adding individual frames during and immediately af- ter the bump. The difference image shows no flux excess at the location of Bergfors-6 during this bump, indicating that the residual feature in the time series in not associ- ated with Bergfors-6. More exotic explanations, such as an attribution of this bump to accretion onto WASP-12, should be treated with skepticism at present. 9 Our Python implementation of this algorithm is available at the primary author's website. Fig. 2. -- Top: relative 2.315 µm narrowband photometry of WASP-12 (points, binned by a factor of ten for plotting purposes; errorbars are the standard deviation on the mean of each set of ten points) and our best-fit model (solid line). The 1σ range of our model is also indicated by the dashed curves. Solid points are used in our analysis, while open points are excluded. Middle: Residuals to the fit. Bottom: differential sky background measured in the two MOIRCS detectors. The vertical dashed line indicates the onset of the photometric ramp apparent in sky and stellar photometry; we exclude this data from our analysis, though the occultation depth is unchanged if we use all data and include the sky trend as an additional decorrelation parameter. The dotted lines indicate the four points of contact corresponding to a circular orbit with our best-fit eclipse center. roughly 0.41% to 0.50%) when using, respectively, larger or smaller photometric apertures, which indicates the sensitivity of this analysis to our particular choice of pa- rameters. For this reason we quote the measurement uncertainty of 0.05 % above, which is roughly twice that predicted from our prayer-bead analysis. 1.0001.0051.0101.015Normalized flux0.0000.0040.008Residuals0.400.450.500.550.600.650.70Orbital phase0.930.950.970.99Sky1 / Sky2131030Number of Individual Points Binned10-410-3RMS of Binned Residuals 6 Crossfield et al. TABLE 2 Dilution Factors and Corrected WASP-12b Transit and Occultation Depths Filter NB2315 Ks (MKO) z J H Ks Reported Depth 0.00082 ± 0.00015 0.00131 ± 0.00028 0.00176 ± 0.00018 0.00309 ± 0.00013 0.00281 ± 0.00085 0.0041 ± 0.0005 0.00379 ± 0.00013 0.00382 ± 0.00019 0.00629 ± 0.00052 0.00636 ± 0.00067 0.0033 ± 0.0004 0.0039 ± 0.0003 0.0050 ± 0.0004 0.01252 ± 0.00045 0.0126 ± 0.0004 IRAC CH2 0.0125 ± 0.0003 IRAC CH1 0.0112 ± 0.0004 IRAC CH2 Johnson R 0.01380 ± 0.00016 IRAC CH1 IRAC CH2 IRAC CH3 IRAC CH4 IRAC CH1 IRAC CH2 IRAC CH2 V/i'c B V i' -- -- -- 1.000 ± 0.000 1.000 ± 0.000 1.000 ± 0.000 1.000 ± 0.000 1.000 ± 0.000 1.000 ± 0.000 0.902 ± 0.018 0.911 ± 0.016 0.855 ± 0.036 0.788 ± 0.068 0.850 ± 0.038 0.833 ± 0.049 0.833 ± 0.049 0.230 ± 0.100 0.833 ± 0.049 0.850 ± 0.038 0.833 ± 0.049 1.000 ± 0.000 Aperture Fraction Dilution Fraction 0.0397 ± 0.0017 0.0606 ± 0.0029 0.0830 ± 0.0039 0.0981 ± 0.0047 0.0994 ± 0.0047 0.1002 ± 0.0045 0.1168 ± 0.0055 0.1204 ± 0.0060 0.1217 ± 0.0060 0.1307 ± 0.0063 0.1168 ± 0.0055 0.1204 ± 0.0060 0.1204 ± 0.0060 0.0186 ± 0.0023 0.1204 ± 0.0060 0.1168 ± 0.0055 0.1204 ± 0.0060 0.01571 ± 0.00096 0.0045 ± 0.0015 0.0090 ± 0.0030 0.0281 ± 0.0014 -- -- -- Referencea This work LM10 Cr11 Cr11 Cr11 Z12 Corrected Depth 0.00085 ± 0.00016 0.00139 ± 0.00030 0.00191 ± 0.00020 0.00339 ± 0.00014 0.00309 ± 0.00093 0.0045 ± 0.0006 0.00419 ± 0.00014 0.00424 ± 0.00021 0.00694 ± 0.00057 0.00701 ± 0.00074 0.00363 ± 0.00044 0.00429 ± 0.00033 0.00550 ± 0.00044 Co12 (null)b 0.01257 ± 0.00045 0.01386 ± 0.00044 Co12 (null)b 0.01374 ± 0.00033 0.01232 ± 0.00044 0.01402 ± 0.00016 Ca11 Ca11 Ca11 Ca11 Co12 Co12 Co12 Co12 M11 C11c -- -- -- -- -- -- a LM10: L´opez-Morales et al. (2010), Cr11: Croll et al. (2011), Z12: Zhao et al. (2012), Ca11: Campo et al. (2011), Co12: Cowan et al. (2012), M11: Maciejewski et al. (2011), C11: Chan et al. (2011) b Results from the "null hypothesis" of Cowan et al. (2012), which assumes zero ellipsoidal variation in their 4.5 µm observations. c These transit analyses average multiple photometric bands, so their correction factors may be less precise. See Sec. 5. We also attempted an alternative analysis in which we used all the data (including that affected by the anoma- lous background trend) and included the sky background trend as an additional decorrelation state vector. This analysis gives a marginally deeper secondary (0.53 % ± 0.05 %), but has a higher residual RMS (0.261%), ex- hibits substantially higher levels of correlated noise when averaging on long time scales, and shows a significantly nonzero e cos ω (inconsistent with previous analysis; cf. Campo et al. 2011; Croll et al. 2011; Cowan et al. 2012). These results suggest that our simpler model, which ex- cludes the latter part of our data set, gives the more reliable occultation measurement. 3.3. Occultation Duration and Timing: No Surprises We find no evidence for significant deviations in sec- ondary eclipse duration or in the eclipse's time of cen- ter as compared to expectations from transit observa- tions and a circular orbit. We find a best-fit eclipse du- ration of 179.6 ± 4.5 min, and the eclipse occurs later than predicted by 7 ± 110 s (after accounting for the 23 s light travel time across the system). Again, the pa- rameter distributions are unimodal and approximately normal. An analysis of previous Ks band observations reported a marginally longer secondary eclipse dura- tion (195 ± 7 min; Croll et al. 2011), while a weighted mean of Spitzer/IRAC occultations give a duration of 177.7±2.1 min (Campo et al. 2011, ; Cowan et al. do not report the durations of their transits and eclipses) and the z' occultation showed a duration of 169 min (with no uncertainty reported; L´opez-Morales et al. 2010). Our narrowband measurement is consistent with these last two values (and with the duration expected from a cir- cular orbit) and is within 3σ of the Ks band broadband results. Our data provide no evidence for an offset or longer-duration eclipse. Together, the secondary eclipse timing and duration tightly constrain the orbital eccentricity and longitude of periastron (Winn 2010; Seager 2011). We deter- mine e cos ω and e sin ω to be +0.00006 ± 0.00091 and 0.011 ± 0.013, respectively, which we interpret as being consistent with a circular orbit and with previous results based on eclipse and radial velocity observations (Campo et al. 2011; Croll et al. 2011; Husnoo et al. 2011). The time of eclipse also constrains the planetary velocity off- set expected at transit center, which can mimic wind- induced velocity offsets measured with high-resolution spectroscopy (Snellen et al. 2010; Fortney et al. 2010; Montalto et al. 2011; Miller-Ricci Kempton & Rauscher 2011). Using Eq. 3 of Montalto et al. (2011) we set a 3σ upper limit on any such orbit-induced velocity offset of +0.59 km s −1. 4. Bergfors-6: AN OBJECT VERY CLOSE TO WASP-12 4.1. Introducing Bergfors-6 Before undertaking an analysis of WASP-12b's at- mospheric properties, we first pause to describe our improved characterization of a recently detected point source ∼ 3 mag fainter than WASP-12 and only 1" away (Bergfors et al. 2011, 2012). This object requires us to revise upward past measurements of the planet's transits and occultations. During our Subaru observations one of us (I.T.) noticed a slight elongation in our (defocussed) images. This moti- vated us to refocus the system at the end of the night, and we recorded the image shown in Figure 4. It clearly shows a point source roughly 1" from WASP-12. A subsequent literature search revealed that this object was recently Re-evaluating WASP-12b 7 TABLE 3 WASP12/Bergfors-6 Astrometry Parameter Units Subaru/MOIRCS IRTF/SpeX Filter Flux ratio Separation Position Angle Date -- -- arcsec deg UT NB2315 0.108 ± 0.007 1.055(cid:48)(cid:48) ± 0.026(cid:48)(cid:48) 250◦ ± 1◦ 2011 Dec 14 KM KO 0.1048 ± 0.0059 1.078(cid:48)(cid:48) ± 0.033(cid:48)(cid:48) 249.4◦ ± 1.1◦ 2012 Feb 25 fit to the two-dimensional PSFs of both stars. We use multiple elliptical Gaussian functions, holding the rota- tion and dispersion parameters fixed in each of the model PSFs and allowing only a single central location for each star. Thus for n Gaussian functions we have (5+5n) free parameters. We set the pixel uncertainties equal to the expectation from photon and read noise. We find that three elliptical Gaussians minimize the BIC, so we adopt this model and use Markov Chain Monte-Carlo (MCMC) techniques to explore the range of valid parameter space. We find the subsequent pa- rameter distributions to be unimodal and approximately Gaussian. To conservatively account for the uncertain- ties inherent in estimating accurate photometry and as- trometry from a single frame, we inflate the parameter uncertainties estimated from our MCMC by a factor of two. Our final determination of the flux ratio, separa- tion, and system position angle from the MOIRCS data are listed in Table 3. The astrometry is consistent with the initial discovery values (Bergfors et al. 2011, 2012). We confirm this 2.315 µm flux ratio by comparing aper- ture photometry of WASP-12 and (after subtraction of the best-fit WASP-12 PSF model) of Bergfors-6: this ap- proach gives a consistent result. The measurements pre- sented here are consistent with, but more precise than, estimates derived from several of our more poorly focused MOIRCS frames. 4.2.2. IRTF/SpeX K Band Lucky Imaging The well-focused Subaru image described above mo- tivated us to acquire additional images of Bergfors-6. On 2012 Feb 25 (UT) we imaged the WASP-12 system with the IRTF/SpeX guide camera (Rayner et al. 2003), which uses a 5122 Aladdin 2 Insb array with a plate scale of 0.1185" pix−1. Observing through sometimes patchy clouds, we acquired 1,200 0.4 s KM KO (Tokunaga et al. 2002) frames (from airmass 1.2 − 1.5) and 900 0.21 s JM KO frames (from airmass 1.5−2.2). In all observations we held the SpeX instrument rotator at a position angle of 90◦. The J band data were not sufficient to reliably detect Bergfors-6 (presumably because of the increased noise penalties resulting from the use of very short expo- sures and the smaller total integration time); hereafter we discuss only the K band data. We calibrate the SpeX images using a median stack of internal (thermal) flat fields and subsequently per- form binlinear interpolation over a few noticeably bad pixels. We select the top 10% of all frames on the ba- sis of the peak pixel flux near the location of WASP-12, then use the "shift and add" algorithm to align and stack these frames (our S/N is too low for more advanced algo- rithms; Jefferies & Christou 1993; Schoedel et al. 2011). Changing the fraction of frames used in our analysis from Fig. 4. -- Images used for astrometric and relative flux mea- surements: at left, seeing-limited image from Subaru/MOIRCS (2.315 µm narrowband); at right, speckle image from IRTF/SpeX (KM KO). Astrometric parameters derived from these images are listed in Table 3. Both images are displayed at the same orienta- tion and scale; they have different (logarithmic) color stretches in order to highlight the fainter companion. discovered using i and z photometry and assigned a pre- liminary spectral type of K4-M1 V (Bergfors et al. 2011, 2012). We refer to this object as Bergfors-6, because it is the sixth object in Table 2 of Bergfors et al. (2012). The existence of a bound companion at this projected separation ((cid:46) 300 AU) would have potentially profound implications for the dynamical history of the system, and could provide a mechanism for Kozai-induced eccentric- ity and subsequent tidal heating to inflate the planet's radius to its present size (Fabrycky & Tremaine 2007; Nagasawa et al. 2008; Ibgui et al. 2011). Bergfors-6 has not been remarked upon in previous op- tical and infrared transit and occultation observations of the WASP-12 system (Hebb et al. 2009; L´opez-Morales et al. 2010; Chan et al. 2011; Maciejewski et al. 2011; Campo et al. 2011; Croll et al. 2011; Cowan et al. 2012; Crossfield et al. 2012; Zhao et al. 2012). This is likely because with seeing-limited or Spitzer/IRAC resolution the two objects are at best only marginally resolved. The case is worse for most high-precision ground-based pho- tometry because of the common practice of substantially defocusing the telescope, which will clearly preclude de- tection of objects such as Bergfors-6. This star falls within the photometric apertures used in most previous analyses and dilutes the transit and secondary eclipse signals that have been measured (e.g. Daemgen et al. 2009). In this section we confirm the pre- vious detection of Bergfors-6 and more tightly constrain its spectral type. In the following section we then correct previous transit and occultation measurements for the photometric contamination of WASP-12 by Bergfors-6. 4.2. Observations of Bergfors-6 4.2.1. Subaru/MOIRCS NB2315 Image As described above, we recorded the single well- focused MOIRCS image shown in Figure 4. We register the image's coordinate system using the 2MASS point source catalogue (Skrutskie et al. 2006) and confirm the MOIRCS plate scale to be 0.117± 0.001" pix−1 (as listed in the instrument documentation). We conservatively adopt an uncertainty of 1◦ in the instrumental position angle. Bergfors-6 sits in the wings of the WASP-12 point spread function (PSF), so we perform a simultaneous 6h30m32.70s32.75s32.80s32.85s32.90sRA (J2000)+29°40'19.0"19.5"20.0"20.5"21.0"21.5"Dec (J2000)1.00.50.00.51.01.5∆δ [arcsec] 8 Crossfield et al. IRTF/SpeX Astrometric Calibratorsa TABLE 4 WDS identifier 06295+3414 06051+3016 06508+2927 Separation [arcsec] 4.25 ± 0.14 11.74 ± 0.40 6.60 ± 0.22 Position Angle [degrees] 256.8 ± 0.6 177.3 ± 0.6 23.7 ± 0.6 a All observations were made in the MKO K band on UT 2012 Feb 25. 5% to 40% leaves our results unchanged within our esti- mated uncertainties. (However, our most selective anal- yses (which use only 1 − 2% of the data) show a hint of north-south elongation (cid:46) 0.3(cid:48)(cid:48). Though the resulting im- age is quite noisy, we note that the initial discovery image (Bergfors et al. 2011, 2012) also shows a similar elonga- tion. We recommend additional high-resolution imaging of Bergfors-6 to test this elongation.) The final image from our standard (10%) analysis is shown in Figure 4: in this image WASP-12 exhibits an axisymmetric PSF with a Strehl ratio of roughly 8% and a full width at half maximum of 0.33" (roughly a factor of 3 better than the seeing-limited resolution). For astrometric reference we observed three known multiple systems taken from Version 2012-02-12 of the Washington Visual Double Star Catalog (WDS 06295+3414, 06051+3016, and 06508+2927 Mason et al. 2001) moderately near WASP-12, with comparable mag- nitudes to WASP-12, and with separations of 4-12". We took twenty 0.5-1 s frames of each system in the same region of the detector as our WASP-12 images, and in each frame we compute the centroids of both components using standard IRAF tasks. From these measurements and their dispersion we derive a SpeX guider plate scale of 0.116(cid:48)(cid:48) ± 0.004(cid:48)(cid:48) and an intrinsic field rotation (i.e., true position angle minus measured position angle) of −0.5◦ ± 0.6◦. We adopt these values in our subsequent analysis and list our astrometry of the WDS stars in Ta- ble 4. We determine the flux ratio of the two stars using aper- ture photometry. Bergfors-6 is located in the PSF wings of WASP-12, so we must account for this contamination. Because our PSF is quite symmetric (though distinctly non-Gaussian) we compute an average radial profile for WASP-12 (after masking out the 90◦ wedge of sky di- rected toward Bergfors-6). We reinterpolate this one- dimensional profile into a two-dimensional model PSF. We estimate the uncertainty of the profile by taking the standard deviation on the mean in each annular bin, and propagate these uncertainties along with the combined photon and read noise. We then subtract the WASP- 12 model PSF from the image and compute partial-pixel aperture photometry at the locations of Bergfors-6 and WASP-12. Residuals are still apparent near the center of WASP-12, so we restrict our analysis to smaller aper- tures: an inner aperture radius of 2.5 pix provides the highest S/N (and least evidence for contamination) for Bergfors-6, so we use this aperture for both systems. Our final estimate of the KM KO flux ratio is listed in Table 3, and it is consistent with our narrowband MOIRCS mea- surement. We measure the relative astrometry of WASP-12 and Bergfors-6 by computing the centroid of WASP-12 in the speckle image, and of Bergfors-6 in the profile-subtracted image. We estimate the uncertainties in these measure- ments by bootstrap resampling (Press 2002), in which we repeat our analysis many times using synthetic data sets, constructed by sampling (with replacement) our original set of 1,200 images. We list the separation and position angle derived from the IRTF speckle data in Table 3. 4.2.3. Spitzer/IRAC Imaging We also examined Spitzer/IRAC subarray data (3.6 µm and 4.5 µm, from Cowan et al. 2012) to search for evidence of Bergfors-6. We performed a weighted least squares fit to each median stack of 64 subarray frames (using the pixel uncertainties provided by the IRAC cal- ibration pipeline, Version 18.18.0) by linearly interpo- lating the appropriate 5× oversampled point response functions10 (PRF) to account for subpixel motions. We do see evidence for an additional point source in the IRAC data, located approximately 1-2 pixels west- southwest of WASP-12. However, we are unable to mea- sure precise astrometry or relative photometry with these data for several reasons. First, the IRAC plate scale (1.09" pix−1) is comparable to the WASP-12/Bergfors-6 separation; second, the IRAC PSF is undersampled at these wavelengths. Consequently, we see clear evidence for oversubtraction in the PRF fitting at the location of Bergfors-6, so we cannot reliably determine the system flux ratio (de-weighting the pixels closest to Bergfors-6, but offset from the WASP-12 core, does not change this result). From these measurements we estimate a flux ra- tio of > 7% in the two IRAC channels, consistent with our ultimate interpretation of Bergfors-6 as a cool stellar object. 4.2.4. Keck/NIRSPEC Spectroscopy We searched online data archives for additional evidence of Bergfors-6, and found a set of high- resolution K band spectra taken with Keck/NIRSPEC, a high-resolution, cryogenic, echelle, NIR spectrograph (McLean et al. 1998), on UT 2010 Apr 22 (Keck Pro- gram ID C269NS, P.I. G. Blake). This data set con- sists of 16 four-minute integrations of WASP-12 taken using the 0.432"× 24" slit. The WASP-12 observations were taken at a position angle of ∼73◦(roughly aligned with WASP-12 and Bergfors-6), and the seeing was suffi- ciently good to distinctly resolve the two components in the spectra. We extract our spectra using our own set of Python tools to trace the spectra in the dark-corrected and flat- fielded NIRSPEC frames. In each echelle order of each frame, we compute a high S/N mean spectral profile by collapsing the trace along the dispersion direction and fit two Gaussian functions to this profile: this provides an estimate of the projected separation of WASP-12 and Bergfors-6 in each frame. We then fit two Gaussian func- tions to each resolution element while holding constant the positions of the two sources: the amplitude of each Gaussian represents the flux in that wavelength element. We then compute weighted means from the individual 10 Available at http://irsa.ipac.caltech.edu/data/SPITZER/ docs/irac/calibrationfiles/psfprf/ Re-evaluating WASP-12b extracted spectra and estimate uncertainties by measur- ing the variations in each pixel, after excluding points deviating by > 3σ. Using a high-resolution simulated telluric spectrum (generated using ATRAN; Lord 1992) we identify known telluric lines and compute a best-fit dispersion function in each echelle order. Estimating a line centroid preci- sion of 0.5 pix, we find a cubic or quartic polynomial minimizes the BIC of these fits. The residuals to our dis- persion solutions have RMS values (cid:46) 0.1 Aand maximum excursions of < 0.2 A. In the raw NIRSPEC frames the spatial axis of the slit is not aligned with the NIRSPEC detector columns, so spectra taken at the A and B nod positions are offset from each other. We spline-interpolate the spectrum in each echelle order and cross-correlate it at sub-pixel in- crements with a high signal to noise (S/N) template spec- trum (Deming et al. 2005). We construct our template by taking the temporal average, after removing outliers, of all our spectra. A parabolic fit to the peak of each spectrum's cross correlation provides the optimal offset value, and we then spline-interpolate all the spectra to a single, common reference frame. We then combine the re- sulting set of aligned spectra (excluding outlying points) and thereby provide a set of simultaneous high-resolution spectra of both WASP-12 and Bergfors-6. The spectra of WASP-12 and Bergfors-6 have median S/N values of 203 and 32 per pixel, respectively. Because the two spectra are obtained simultaneously and the objects are separated by only 1" we expect the telluric signature in both spectra to be indistinguishable. We therefore divide the spectrum of Bergfors-6 by that of WASP-12 to remove the effect of telluric absorption. Possible misalignment of the spectrograph slit prevent these data from usefully constraining the absolute flux ratio of these two objects, but the data constrain tightly Bergfors-6's spectral type from the relative strengths of individual spectral features. The NIRSPEC spectra do not cover the wavelengths of standard gravity indicators such as Na or Ca lines, so our subsequent analysis focuses on the two most prominent gravity-sensitive features cov- ered by these data: the 12CO (2, 0) and (4, 2) bandheads located at 2.294 µm and 2.353 µm (Kleinmann & Hall 1986). As we now describe, we find that Bergfors-6 is a hot M dwarf. 4.3. The Spectral Type and Nature of Bergfors-6 4.3.1. Photometric Constraints With our four (i, z, KM KO, K2315) relative photomet- ric measurements and the NIRSPEC spectrum we de- termine the spectral type of Bergfors-6, as described be- low. First, we apply the relationship between spectral type and absolute magnitude of Kraus & Hillenbrand (2007) using our photometry. This relationship is for main sequence stars, and WASP-12 is 25% larger than a zero-age main sequence star of the same mass (Hebb et al. 2009; Torres et al. 2010). Accounting for this, and assuming Tef f = 6300 ± 100 K, gives a distance modulus for WASP-12 of µW12 = 7.7 ± 0.2 mag (signif- icantly nearer than the previous estimate; Chan et al. 2011). For Bergfors-6, our estimate of the i − K color (after applying the color transformations of Carpenter 2001) implies a main-sequence spectral type of M0-M1 9 (Tef f = 3700 ± 100 K). Assuming systematic uncertain- ties of 0.2 mag gives µB6 = 7.1 ± 0.2 mag, rather closer than WASP-12 if Bergfors-6 is still on the main sequence. That Bergfors-6 is closer to Earth than is WASP-12 is the opposite of the trend noted by Daemgen et al. (2009), who uniformly estimated that their faint companions to planet host stars were more distant than the brighter component. Next, we fit only the relative (Bergfors-6/WASP-12) stellar photometry using low-resolution stellar atmo- sphere models (Castelli & Kurucz 2004). We interpolate to the effective temperature, [Fe/H], and surface gravity of WASP-12 and hold these values fixed in the model- ing. For Bergfors-6 we assume a metallicity equal to that of WASP-12 and allow three free parameters: sur- face gravity, effective temperature, and a geometric fac- tor (f = RB6 ) relating the relative sizes and helio- RW12 centric distances of the two stars. A standard Pythonic minimizer and an MCMC analysis using the emcee affine- invariant sampler (Foreman-Mackey et al. 2012) provide the desired physical parameters and their uncertainties. The derived parameters for Bergfors-6 are 3840 ± 70 K (which agrees well with our previous estimate of this ob- ject's effective temperature) and f =0.452 ± 0.015; the covariance between these parameters is -1.73 K. dW12 dB6 We also considered that Bergfors-6 might be an ex- tragalactic, rather than a stellar, contaminant (Luhman & Mamajek 2010). However, a comparison of its pho- tometric spectral energy distribution with low-resolution galactic spectral templates (Assef et al. 2010) suggests that this explanation is unlikely. 4.3.2. Spectroscopic Constraints Spectroscopy is particularly well-suited to constrain surface gravity. We now use both the NIRSPEC spec- trum and the relative photometry described above to constrain Bergfors-6's parameters, again using the emcee MCMC sampler (Foreman-Mackey et al. 2012). For this analysis we use the BT-Settl library11 computed using the PHOENIX atmosphere code(Allard et al. 2010). This library provides high-resolution model spectra across a wide range of parameter space. We use the so-called "hot" models using abundances from Asplund et al. (2009) with no alpha enhancement. As noted previously, we use only the two NIRSPEC echelle orders that cover CO bandhead features -- the wavelengths from 2.273- 2.308 µm and 2.344-2.380 µm. For a given set of input parameters, our spectral mod- eling algorithm begins by logarithmically interpolating between BT-Settl models at the nearest values of Tef f , log g, and [M/H]. The model then applies (a) a Doppler shift, (b) a quadratic continuum normalization (because the absolute slope and curvature of the spectrum is un- known owing to possible slit misalignments), and (c) a convolution with a Gaussian kernel of specified width. Finally, we bin (not interpolate) the model spectrum onto our NIRSPEC pixel grid and compute relative broadband photometry as described in the preceding section -- ex- cept that here we propagate the uncertainties in WASP- 12's parameters into the modeling by performing a ran- dom draw from normal distributions in Tef f , log g, and 11 Available online at http://phoenix.ens-lyon.fr/ 10 Crossfield et al. (FWHM = 8.1A) Brackett γ line. After correcting for the Earth's velocity along the line of sight (using the Python routine astrolib.baryvel) we estimate radial velocities for WASP-12 and Bergfors-6 of 16.5 ± 2.6 km s −1 and 19.7 ± 1.3 km s −1, respectively. These values are consis- tent with the radial velocity of 19.1 km s−1 derived from WASP-12b's initial radial velocity measurements (Hebb et al. 2009; Campo et al. 2011). This common veloc- ity is consistent with a scenario in which either WASP- 12 and Bergfors-6 are gravitationally bound and share a common three-dimensional space motion, or in which the consistency of the two stars' radial velocities is merely a coincidence. We also find no evidence for multiple spectral line pro- files, which would indicate Bergfors-6 is an unresolved binary. The cross-correlation profiles of our data and spectral template have full-widths at half-maximum of approximately 14 km s−1. All three echelle orders con- taining strong CO features (centered on 2.29, 2.36, and 2.44 µm) show unimodal cross-correlation peaks and no evidence of the secondary peaks that would suggest an additional cool companion. Thus Bergfors-6 shows no spectroscopic evidence of binarity. 4.5. Interpretation of Bergfors-6 We conclude that Bergfors-6 is a cool star showing ab- sorption features consistent with a high surface gravity. Our spectroscopy implies that Bergfors-6 is a main se- quence star, and our relative photometry suggests that Bergfors-6 lies 50% closer to Earth than does WASP- 12. In this scenario, the hint of elongation alluded to in Sec. 4.2.2 is spurious and the consistent radial velocities of WASP-12 and Bergfors-6 is coincidental. However, if Bergfors-6 were a binary M dwarf system observed near conjunction then the binary and WASP-12 could lie at the same distance from Earth; if the three components were bound this scenario would offer a natural explana- tion for the consistent systemic velocities. Because of the intriguing possibilities inherent in a gravitationally bound arrangement, we briefly discuss the implications of such a scenario below. At the distance of WASP-12, the projected separation of Bergfors-6 is roughly 400 ± 100 AU. Such an object would have an orbital period of several thousand years, and as such would be marginally compatible with the inferred upper limit of binary separation (∼ 300 AU) needed to substantially influence planetary migration and dynamics (Desidera & Barbieri 2007). Assuming Bergfors-6 is near apastron implies a Kozai oscillation timescale (Fabrycky & Tremaine 2007) of (very roughly) 20[(1 − eB6)/(1 + eB6)]3/2 Gyr, where eB6 is Bergfors-6's orbital eccentricity. For substantial Kozai interactions to have taken place during WASP-12's lifetime (Hebb et al. 2009; Chan et al. 2011) we thus require eB6 > 0.7. Long- period binaries with such high eccentricities are rare, but they do exist (Duquennoy & Mayor 1991). Fabrycky & Tremaine (2007) predict that hot Jupiter systems with an additional, widely-separated stellar com- ponent will preferentially exhibit misalignment between their stellar spin and planetary orbital axes. Recent ob- servations of WASP-12 have determined that the sky- projected angle between the spin and orbital axes is 59+20−15deg, strongly suggesting a misaligned system (Al- Fig. 5. -- Keck/NIRSPEC spectrum of Bergfors-6 (black curve, top and middle) and relative photometry of WASP-12 and Bergfors-6 (filled points, bottom), and our best-fit Phoenix/BT- Settl model (red curve). As discussed in Section 4.3, the ensemble of measurements suggests Bergfors-6 is a hot M dwarf located 50% closer to Earth than WASP-12. Top and Middle: The blue curves show our estimated spectroscopic measurement uncertainties; the vertical scale for these curves is indicated at right. Bottom: The open points at bottom are the inferred photometric dilutions of transits or occultations measured in various bandpasses (tabulated in Table 2); we indicate filters used in this analysis with solid lines, while other filters are denoted with dashed lines. The error bars of the open points represent the 68.3% confidence intervals on these dilution estimates, taking into account the uncertainties in our fit. dW12 dB6 [M/H] at each step in the MCMC analysis that follows. In our analysis the spectral type of Bergfors-6 is con- strained almost entirely by the ∼ 2000 spectroscopic data points, while the geometric ratio RB6 is constrained RW12 only by the broadband photometry. The results of this analysis for Bergfors-6 are an effective temperature of 3660+85−60 K and a log g (cgs) of 5.13+0.38−0.22; we show our best-fit model spectrum in Figure 5. Our derived pa- rameters are fully consistent with an M0 dwarf on the main sequence, which would imply a radius of 0.5−0.6R(cid:12) (Torres et al. 2010). The geometric ratio from our analy- sis is 0.520+0.027 −0.037, which implies a radius 50% larger than expected for a main-sequence dwarf lying at the same distance at WASP-12. Our spectral analysis therefore suggests that Bergfors-6 lies approximately 50% closer to Earth than does WASP-12 and that it represents a chance foreground alignment. 4.4. Radial Velocities To estimate the radial velocities of the two stellar com- ponents in our telluric-corrected Bergfors-6/WASP-12 ratio spectrum we cross-correlate with BT-Settl mod- els (Allard et al. 2010). We cross-correlate each echelle order of the ratio spectrum with models with effective temperatures of 6200 K and 3800 K. We also cross- correlate the spectra (before telluric correction) with the high-resolution atmospheric transmission profile of Hin- kle et al. (2003) to establish our observational reference frame. The radial velocity of Bergfors-6 is constrained mainly by the strong CO bands lying redward of 2.29 µm, while WASP-12's is tightly constrained only by the broad 2.2802.2852.2902.2952.300Wavelength [µm]0.60.70.80.91.01.1Normalized Flux0.00.10.20.30.40.5Flux Errors2.3502.3552.3602.3652.3702.375Wavelength [µm]0.60.70.80.91.01.1Normalized Flux0.00.10.20.30.40.5Flux Errors0.40.8124810Wavelength [µm]0.000.020.040.060.080.100.120.140.16FB2/FW12 Re-evaluating WASP-12b 11 brecht et al. 2012). This may be circumstantial evidence that WASP-12 has migrated via Kozai interactions and that a long-period bound companion is required (e.g., Ibgui et al. 2011). Ultimately, high-resolution imaging can most quickly determine whether WASP-12 and Bergfors-6 truly ex- hibit common proper motion and are gravitationally bound, and whether Bergfors-6 is a single or multiple sys- tem. Based on WASP-12's proper motion (∼ 8 mas yr−1; Zacharias et al. 2004), speckle or seeing-limited astrom- etry of the type presented here will not be sufficient for this purpose. However, a two-year baseline of large- aperture adaptive optics imaging (Yelda et al. 2010) could suffice to confirm or rule out common proper mo- tion. 5. REVISING PAST TRANSIT AND SECONDARY ECLIPSES Because Bergfors-6 was not noted in previous transit and secondary eclipse observations of WASP-12b, these flux diminutions were diluted by this faint star's con- stant baseline flux. This effect is largest in the infrared; the results of optical observations change by only a few percent, less than their typical uncertainties. Although a full re-evaluation of WASP-12's system parameters is beyond the scope of this work, we correct the depth mea- surements for the contamination effect and present re- vised transit and occultation depths below. We propa- gate the uncertainties in Bergfors-6's effective tempera- ture into our estimates of the photometric dilution caused by Bergfors-6, which we list in Table 2. In some sec- ondary eclipses, and in all transits, the corrections we apply change the previously reported depths by > 1σ. Essentially all the light from Bergfors-6 lies within the apertures of ground-based observations (Hebb et al. 2009; L´opez-Morales et al. 2010; Croll et al. 2011; Chan et al. 2011; Maciejewski et al. 2011; Zhao et al. 2012), but Spitzer/IRAC analyses use narrower apertures (Campo et al. 2011; Cowan et al. 2012) and so only a portion of Bergfors-6's starlight contaminates these secondary eclipse measurement. To estimate the IRAC contami- nation fraction we generate 10× super-sampled PSFs for all four IRAC channels12, using a 6300 K blackbody spec- trum simulated at the center of the instrument field of view. We then compute aperture photometry using the reported photometric aperture diameters (Campo et al. 2011; Cowan et al. 2012) at a position offset by 1.05" from the PSF center to estimate how much of Bergfors-6's flux fell into the WASP-12 aperture in these analyses. We ig- nore possible time-variable illumination caused by the intrapixel effect (Charbonneau et al. 2005). Note that the phase curve observations of Cowan et al. (2012) are also diluted by Bergfors-6, and must be revised upward by the same factors as indicated in Table 2. This in turn increases the ellipsoidal variation inferred from these measurements, placing the IRAC 4.5 µm results in even stronger conflict with those from WFC3 (Swain et al. 2012). The correction of optical transit measurements is com- plicated by the common, but deplorable, practice of re- porting a single transit depth when using observations taken in different bandpasses (Hebb et al. 2009; Chan 12 Using Tiny Tim; available at http://ssc.spitzer.caltech. edu/ et al. 2011). Atmospheric characterization via transit observations depends on the fundamentally wavelength- dependent planetary radius during transit, and so we recommend that future analyses of multiple photomet- ric data sets report the transit depths measured in each bandpass in addition to a single, achromatic value. In addition, such analyses are not always clear about the relative weighting of data points from separate observa- tions. Nonetheless we attempt to estimate the relative weightings and derive appropriate correction factors for prior multi-band optical transit observations of WASP- 12. The analysis of Maciejewski et al. (2011) used only a single bandpass (Johnson R) and so their Johnson R transit depth is the most reliable optical transit measure- ment in Table 2. The analysis of Chan et al. (2011) uses two transit data sets: 671 V band and 470 i' band obser- vations with residual RMS values of 2.0 and 1.2 mmag, respectively. This work uses an additional multiplicative term to increase the per-point data uncertainties (1.48 and 1.57, respectively). Assuming that the statistics of the final reported transit depth behaves similarly to a weighted mean, we estimate that the two data sets con- strain the transit parameters with roughly equal weight. We use these weights to determine a weighted aver- age of the correction factors for each bandpass. Such a weighted average is only a rough approximation to the true correction factor, so the correction factors for these analyses are somewhat less certain. We thank the referee for pointing out that the situation is even more muddled for the discovery paper (Hebb et al. 2009), which uses three transit data sets: 227 B band, 614 z' band, and 6393 SuperWASP (roughly V band) observations. Be- cause Hebb et al. (2009) do not report the residual RMS scatter of the SuperWASP photometry, we do not at- tempt to determine the relative weighting of these sev- eral transit observations and we do not report a mean dilution-corrected transit depth for these data. The dilution correction factors, and dilution-corrected transit and occultation depths, are listed in Table 2. The optical transit depths in particular are increased by a few percent, and the planetary radius increases by only half this factor. Owing to current uncertainties in stellar properties, such an effect is smaller than current uncer- tainties on WASP-12b's physical radius. 6. EMISSION SPECTRUM AND ATMOSPHERIC PROPERTIES We now return out attention to the nature of WASP- 12b's atmosphere as constrained by its spectral energy distribution (SED). In our narrow bandpass we find a brightness temperature of 3640 K ± 230 K by using the model stellar spectrum described in the following sec- tion, modeling WASP-12b's emission in our bandpass as a blackbody, and propagating the 1σ uncertainties in our occultation measurement. This brightness temperature is rather higher than the planet's equilibrium tempera- ture of 2990 ± 110 K (computed using the Bond albedo and recirculation efficiencies from Cowan et al. 2012) and is higher than inferred in any broad photometric band- pass (L´opez-Morales et al. 2010; Croll et al. 2011; Campo et al. 2011; Cowan et al. 2012). The CFHT/WIRCam Ks filter cuts off just where the NB2315 filter cuts in, so our measurement is not in conflict with this previous 12 Crossfield et al. secondary eclipse observation (Croll et al. 2011). In the following, we use the weighted averages of the two sets of IRAC (3.6 µm and 4.5 µm) measurements (Campo et al. 2011; Cowan et al. 2012, ; see also Table 2). Two possible 4.5 µm secondary eclipse depths were re- ported by Cowan et al. (2012), and we use their "null hypothesis" consistent with no ellipsoidal variations (as implied by the constraints placed on WASP-12b's shape by Swain et al. 2012). We first introduce the improved, model-independent constraints these observations place on WASP-12b's bolometric luminosity in Sec. 6.1, and we then discuss our efforts to generate a coherent model of the planet's emission spectrum in Sec. 6.2. 6.1. Bolometric Luminosity WASP-12b has one of the best-determined bolomet- ric luminosities of any extrasolar planet (Cowan & Agol 2011). In an earlier work (Crossfield et al. 2012) we dis- cussed the current constraints on the bolometric lumi- nosity of WASP-12b. Here we update this analysis in light of our dilution-corrected occultation measurements and recent Spitzer/IRAC and Hubble/WFC3 observa- tions (Cowan et al. 2012; Swain et al. 2012). Current measurements now constrain the planet's day- side to have Lbol = (3.6 − 5.0) × 1030erg s−1, where the lower limit assumes the case of zero emission between the observed bandpasses. Here we have followed the same approach as our previous calculation, but we now use the WFC3 spectrum (Swain et al. 2012) from 1.1 -- 1.485 µm (replacing the J band eclipse, but retaining the H band eclipse, of Croll et al. 2011). In this analysis our narrowband measurement, and the controversy over the IRAC 4.5 µm measurement, affect the luminosity by only 0.1 × 1030erg s−1. On the basis of thermal occultation and phase curve measurements, WASP-12b's bolometric albedo has been inferred to be AB = 0.25±0.1 (Cowan et al. 2012), which implies that the planet absorbs (3.8 ± 0.8) × 1030erg s−1 from its host star. Following the approach of Crossfield et al. (2012), we find that this value constrains the night- side luminosity to be < 1.6×1030erg s−1, consistent with the nightside luminosity inferred by Cowan et al. (2012) of 0.06+0.12−0.02 × 1030erg s−1. This last point further as- sumes that the night side (like the day side; see below) emits approximately like a blackbody. We are thus nearing a bolometric luminosity suffi- ciently well-constrained that we can test evolutionary models of this planet. The current uncertainty in Lbol is dominated by measurements at the shortest wavelengths, suggesting that occultation observations at wavelengths < 1 µm may be the best next step toward an even more tightly constrained bolometric luminosity. 6.2. Atmospheric Models The first systematic effort to retrieve WASP-12b's at- mospheric parameters (using infrared broadband sec- ondary eclipse photometry) inferred a high C/O ratio (> 1) and ruled out any strong temperature inversion at the pressures probed (0.01-2 bar; Madhusudhan et al. 2011). That study published several representative mod- els, all of which had χ2 ∼ 10 with seven measurements and ∼ 10 free parameters (BIC∼ 32). These models also predicted a strong absorption feature (depth (cid:46) 0.2%) at 2.315 µm: our narrowband secondary eclipse measure- ment rules out this absorption feature at > 3σ. This narrowband result, our characterization of Bergfors-6 and correction for its diluting effect, and the alterna- tive Spitzer/IRAC 4.5 µm eclipse depth of Cowan et al. (2012) all indicate the need for a new analysis of WASP- 12b's atmospheric properties. To better understand the nature of the planet's dayside emission we constructed a variety of atmosphere mod- els for WASP-12b, following Barman et al. (2001, 2005). We compared model SEDs to the corrected occultation depths from Table 2. For this exercise we adopt the Cowan et al. (2012) 4.5 µm value without ellipsoidal vari- ations (their "null hypothesis"). As discussed by Cowan et al., when ellipsoidal variations are allowed, the inferred planet elongation in this band is substantially larger than expected. Indeed, recent NIR HST/WFC3 observations exclude the large ellipsoidal variations inferred from the Spitzer/IRAC analysis (Swain et al. 2012). Further- more, the measured primary transit radius at 4.5 µm differs dramatically from radii at other wavelengths as well as predicted transit spectra. Given that two very unexpected physical properties are required to explain the ellipsoidal variations, the 4.5 µm secondary eclipse value without ellipsoidal variations is likely more reliable (Cowan 2012, private communication). Finally, our ul- timate result -- that the high C/O result is not justified by current photometric data -- does not change even if we use the average of the Campo et al. and Cowan et al. results. By far the model that best matches the broad band occultation photometry is a 3000 K black body (upper panel of Fig. 6, and Fig. 7), with χ2 = 15 with only two free parameters (BIC= 19). Taking the mean of the Campo et al. and Cowan et al. results gives a black body model with higher χ2 = 25 and BIC= 29, but this BIC value is still smaller than that for the high C/O models. Discussed above. This lower BIC value indicates that current observations do not justify the use of these more complicated models. Our best-fit black body temper- ature matches the expected equilibrium temperature of the planet if redistribution of heat to the night side is ex- tremely inefficient. The close match to a blackbody also indicates that the planet's photosphere is nearly isother- mal at the depths probed by these wavelengths. Our default irradiated atmosphere model assumes zero redistribution of heat to the nightside (consistent with the black body analysis) and metal abundances matching those of the host star ([Fe/H] = 0.3). This model pre- dicts a temperature inversion above an isothermal region (Fig. 7); it also has χ2 = 15; since it has substantially more free parameters than a black body, the BIC for this model fit is substantially worse than for the simpler model in the previous paragraph. In such a model the inversion extends partially across the IR photosphere re- sulting in modestly inverted CO and H2O bands (lower panel of Fig. 6). However, the model band-integrated fluxes are similar to the black body and match the broad band observations equally well (Fig. 6). Naturally, then, the blackbody model provides a substantially lower BIC than the radiative-transfer models (both our model and the models presented by Madhusudhan et al. 2011). Re-evaluating WASP-12b 13 simple blackbody model (within 2σ), we additionally in- vestigated what mechanisms might cause such a narrow band to exhibit a substantially higher flux. In an attempt to reproduce the narrow band flux, we made a number of ad-hoc modifications to the temperature-pressure (T- P) profile of the default atmosphere model. The base of the inversion was moved to higher and lower pressures, the steepness of the inversion was altered and several Gaussian temperature perturbations were added. None of these models could match the narrow-band flux with- out negatively impacting the comparison at other wave- lengths. A dozen different carbon and oxygen abun- dances (ranging from 7.7 − 9.4 on the standard base- 10 logarithmic scale, where H2 abundance is exactly 12; these abundances overlap the best-fitting regions from Madhusudhan et al. 2011) were also explored, under the assumption of chemical equilibrium, with similarly neg- ative results. Finally, we constructed a dayside model by dividing the planet's hemisphere into 10 concentric regions centered on the substellar point. These regions were modeled separately, each receiving flux along the line-of-sight to the star. The outgoing intensities dur- ing secondary eclipse, along the line-of-sight to the ob- server, were integrated to produce a dayside spectrum following Barman et al. (2005). The predicted limb-to- substellar temperature structure implies a horizontal as well as vertical temperature inversion (i.e., along a path from the substellar point to the terminator at a con- stant radius or pressure the temperature decreases, then increases again), but the surface-integrated fluxes from this model did not significantly differ from the default one-dimensional model. Our main conclusion from this modeling exercise is that it is very unlikely that strong narrow band emission could be reproduced by a mod- ification to the thermal profile alone, given the strong constraints placed on the SED by the broad band pho- tometry. As discussed above, the default, 1D, model predicts a steep temperature inversion. A direct consequence of this inversion is a swath of narrow emission lines that form in the inverted atmospheric layers (grey-shaded spectrum in Fig. 6). While the narrow band filter does not encom- pass the complete CO band, it does cover many of the predicted narrow emission lines including the strong 3-1 band. In the default model, these lines are very bright relative to the continuum but are too narrow to signifi- cantly impact the photometry. The model exercise above weakens the case for flux increases in the pseudocontin- uum, leaving the lines as a potential explanation for the deep narrow-band secondary eclipse. If the line fluxes are increased by a factor of 2 to 4, the narrow band pho- tometry can be reproduced. In this ad-hoc scenario, the lines would be extremely bright compared to those of the default model (see comparison in Fig. 6 inset). In the default atmosphere model, CO is the fourth most abundant molecule throughout most of the atmosphere including much the inversion region, as shown in Fig. 7. Previous studies have shown that local thermodynamic equilibrium (LTE) is achieved for CO in the atmospheres of isolated late-type stars (Ayres & Wiedemann 1989; Schweitzer et al. 2000), meaning that the CO lines should map the gas temperature. However, Barman et al. (2002) concluded that Na was well out of LTE in the upper at- mosphere of the modestly irradiated planet HD 209458b, Fig. 6. -- WASP-12b emission spectrum; see Sec. 6 for a full discussion. Solid points are the dilution-corrected photometric sec- ondary eclipse depths listed in Table 2; at 4.5 µm we plot the results of both Campo et al. (2011) and Cowan et al. (2012) The lower panel is the black body comparison (open symbols are the model, band-integrated, points). The upper panel includes several spec- tra: the dark black spectrum (with open black symbols) is our solar abundance, 2π redistribution model spectrum (smoothed for plot- ting purposes by convolving with a Gaussian of FWHM = 100A). The grey spectrum reproduces the black model but is plotted at 100 times higher resolution to show the many narrow emission lines predicted by this model. The inset shows the narrow band region and a model with CO lines scaled up (red curve). Within the inset plot, the "standard" high-res model is also shown in the same grey color as in the main figure. Note the scale: these lines are narrow but very bright. If such strong, narrow emission lines are present, they should be easily discerned by future observations. Fig. 7. -- Temperature-pressure profile (top) and molecular abun- dances (bottom) used to construct the model emission spectrum of WASP-12b shown in Fig. 6. The middle panel shows the normal- ized contribution functions for the indicated filters. Although our narrowband measurement is fit by the 14 Crossfield et al. resulting in strong emission cores for the Na D line pro- files. Nevertheless, the LTE assumption has not been tested for molecules in highly irradiated atmospheres. If such strong, narrow emission lines were present in the atmosphere of a hot Jupiter, they should be easily dis- cerned by future observations. Note that while CH4 also has a strong band head at this wavelength, (as observed in the telluric transmission spectrum and in spectra of L dwarfs; cf. Hinkle et al. 2003; Cushing et al. 2005), at the high temperature of WASP-12b atmospheric condi- tions would have to be much farther from equilibrium for significant CH4 opacity to be apparent. However, a careful non-LTE study is beyond the scope of this pa- per and further observational confirmation of the narrow- band emission is needed. The claim of strong, non-LTE CH4 emission in HD 189733b (Swain et al. 2010), which was based on single-slit IRTF/SpeX spectroscopy, was disputed on the basis of contamination by telluric effects (Mandell et al. 2011). However, the types of spectroscopic contamina- tion discussed by Mandell et al. (2011) do not apply to our narrowband relative photometry. Changes in telluric absorption are common mode over our narrow field of view and should be removed when we divide the flux from WASP-12 by the comparison star flux. Variable telluric emission should be removed by the combination of global sky frame subtraction and local background subtraction in the aperture photometry process (Section 2.2). As dis- cussed in Section 2.3, though the telluric absorption fea- ture intersecting our narrowband filter could explain the systematic photometric ramp (seen shortly after WASP- 12 crosses the meridian; Figure 2 and Sec. 2.3), it seems unlikely that telluric variations could strongly corrupt the secondary eclipse depth while maintaining such a consistent eclipse duration and time of center (Sec. 3.3). Our analysis of the observations to date support a planet with little to no redistribution of flux to the night side, consistent with Cowan et al. (2012). We also find that photometric observations are well-reproduced by a black body and are not yet sufficiently precise to justify the use of more complicated models. If WASP-12b has a near-isothermal photosphere, then secondary eclipse data will be poorly suited to reveal significant compo- sitional information. Other highly irradiated giant plan- ets have also been observed to host nearly isothermal infrared photosphere, including the similarly hot WASP- 18b (Nymeyer et al. 2011) and perhaps the cooler TrES- 2b and TrES-3b (O'Donovan et al. 2009; Croll et al. 2010; Cowan & Agol 2011). If WASP-12b does indeed largely radiate like a black body, previous conclusions about composition and the the presence or absence of a temper- ature inversion are significantly weakened. If our narrow band flux measurement is confirmed at higher precision and this flux is produced by emission lines, there may be hope for high-resolution spectroscopic studies to infer the planet's atmospheric composition and thermal pro- file (e.g. Barnes et al. 2007; Mandell et al. 2011; Rodler et al. 2012; Brogi et al. 2012). 7. CONCLUSIONS We have presented a deeper-than-expected secondary eclipse (0.45 % ± 0.06 %) of the very hot Jupiter WASP- 12b in a narrow band centered at 2.315 µm. The planet's brightness temperature at this wavelength is 3640 K± 230 K, only marginally consistent with WASP- 12b's equilibrium temperature of 2990 ± 110 K (Cowan et al. 2012). Our precision is lower than expected because of an unanticipated systematic trend affecting both sky background and stellar photometry, but we are able to exclude data affected by this trend. The duration and timing of the eclipse we measure from these data are consistent with a circular orbit and with previous mea- surements (Hebb et al. 2009; Croll et al. 2011; Campo et al. 2011; Cowan et al. 2012). Using NIR photometry and high-resolution spec- troscopy, we find that Bergfors-6, a previously identi- fied object only 1" from WASP-12 (Bergfors et al. 2011, 2012), is an M dwarf star with Tef f =3660+85−60 K. If this object is an unresolved binary with two components of equal mass it could lie at the same distance from Earth as does WASP-12. However, Keck/NIRSPEC spectroscopy shows no evidence for binarity. If single, it likely lies closer to Earth than does WASP-12. Adaptive optics imaging on large-aperture telescopes will be necessary to conduct the proper motion studies necessary to dis- criminate between these two scenarios. If WASP-12 and Bergfors-6 are gravitationally bound, further simulations (e.g., Ibgui et al. 2011) should be undertaken to de- termine whether Kozai interactions with an object with Bergfors-6's characteristics could have caused WASP- 12b's inward migration and, through tidal pumping, have inflated the planet's radius (Bodenheimer et al. 2001). Bergfors-6 has heretofore passed unnoticed in previous transit and occultation analyses, which has caused the measured depths of these past events to be underesti- mated. We use our constraints on Bergfors-6 to infer and correct for the dilution of these past observations, in several cases increasing depths by > 1σ. Thus WASP- 12b is rather hotter and slightly larger (by 1 − 2%) than previously reported. These changes emphasize the im- portance of high-resolution imaging surveys in the vicin- ity of newly discovered transiting planets. The ensemble of dilution-corrected secondary eclipse measurements suggests that WASP-12b's atmosphere is largely isothermal across the pressures probed by eclipse observations (Figs. 6 and 7), with a photospheric tem- perature of roughly 3000 K. This result implies that pre- vious claims of a high carbon to oxygen ratio for this planet (Madhusudhan et al. 2011) are not yet justified by the current photometric data. Further observations of the planet's 4.5 µm secondary eclipse depth is certainly warranted to resolve the discrepancy between previous results at this wavelength (Campo et al. 2011; Cowan et al. 2012), as our modeling efforts indicate that achiev- ing such a dramatically lower temperature in the 4.5 µm channel would require a major unexpected change in the opacity source(s) across this bandpass. Regardless, our narrowband measurement alone excludes the mod- els used to infer this high C/O ratio at > 3σ. The lack of absorption at a wavelength where CO, a dominant species in any atmospheric model, should exhibit strong absorption is further evidence for a near-isothermal pho- tosphere. Thus secondary eclipse observations are ill- suited to determine WASP-12b's atmospheric composi- tion, and ultimately transmission spectroscopy may be a more successful approach in pursuit of this goal. WASP-12b is clearly an unusual object, and further Re-evaluating WASP-12b 15 observations are clearly warranted. Aside from the need for additional IRAC 4.5 µm occultation photometry as described above, any or all of narrowband photometry (from the ground or, if available, using HST/NICMOS), single- or multi-object spectroscopy (Swain et al. 2010; Bean et al. 2010; Mandell et al. 2011; Berta et al. 2012), or perhaps by high-resolution phase curve spectroscopy (Barnes et al. 2007; Brogi et al. 2012; Rodler et al. 2012) could be of great utility. Finally, a better measurement of the planet's three-dimensional shape is also highly desir- able, especially given the apparent disagreement between the degree of prolateness inferred by Cowan et al. (2012) and Swain et al. (2012). If WASP-12b is substantially prolate, three-dimensional models, ideally coupled with a general circulation model of the planet's atmospheric dynamics, may also provide important clues toward un- raveling the mystery of WASP-12b's atmospheric struc- ture and composition. ACKNOWLEDGEMENTS We thank N. Cowan for many fruitful discussions about the WASP-12 system, M. Swain and J. Bean for discus- sions about the systematic trend apparent in our rel- ative photometry, K. Stevenson and J. Harrington for reiterating to us the importance of partial pixels in high- precision aperture photometry, C. Bergfors for discus- sions of the object she discovered, and our anonymous referee for detailed comments and sugestions which im- proved the quality of this paper. I.C. was supported by the UCLA Dissertation Year Fellowship and by EACM. B.H. is supported by NASA through awards issued by JPL/Caltech and the Space Telescope Science Center. T.K. acknowledges the fi- nancial support by a Grant-in-Aid for the Scientific Re- search (No. 21340045) by the Japanese Ministry of Ed- ucation, Culture, Sports, Science and Technology, with which the NB2315 filter was made. This research has made use of the Exoplanet Orbit Database at http: //www.exoplanets.org, the Extrasolar Planet Encyclo- pedia Explorer at http://www.exoplanet.eu, and free and open-source software provided by the Python, SciPy, and Matplotlib communities. We will gladly distribute our raw data products, or many of our algorithms, to interested parties upon request. Facilities used: Subaru, IRTF, Keck, Spitzer REFERENCES Albrecht, S. et al. 2012, ArXiv e-prints, ADS, 1206.6105 Allard, F., Homeier, D., & Freytag, B. 2010, ArXiv e-prints, ADS, 1011.5405 Asplund, M., Grevesse, N., Sauval, A. J., & Scott, P. 2009, ARA&A, 47, 481, ADS, 0909.0948 Assef, R. J. et al. 2010, ApJ, 713, 970, ADS, 0909.3849 Ayres, T. R., & Wiedemann, G. R. 1989, ApJ, 338, 1033, ADS Barman, T. S., Hauschildt, P. H., & Allard, F. 2001, ApJ, 556, 885, ADS, arXiv:astro-ph/0104262 -- . 2005, ApJ, 632, 1132, ADS, arXiv:astro-ph/0507136 Barman, T. S., Hauschildt, P. H., Schweitzer, A., Stancil, P. C., Baron, E., & Allard, F. 2002, ApJ, 569, L51, ADS, arXiv:astro-ph/0203139 Barnes, J. R., Leigh, C. J., Jones, H. R. A., Barman, T. S., Pinfield, D. J., Collier Cameron, A., & Jenkins, J. S. 2007, MNRAS, 379, 1097, ADS, 0705.0272 Bean, J. L., Miller-Ricci Kempton, E., & Homeier, D. 2010, Nature, 468, 669, ADS, 1012.0331 Croll, B., Jayawardhana, R., Fortney, J. J., Lafreni`ere, D., & Albert, L. 2010, ApJ, 718, 920, ADS, 1006.0737 Croll, B., Lafreniere, D., Albert, L., Jayawardhana, R., Fortney, J. J., & Murray, N. 2011, AJ, 141, 30, ADS, 1009.0071 Crossfield, I. J. M., Barman, T., & Hansen, B. M. S. 2011, ApJ, 736, 132, ADS, 1104.1173 Crossfield, I. J. M., Hansen, B. M. S., & Barman, T. 2012, ApJ, 746, 46, ADS, 1201.1023 Cushing, M. C., Rayner, J. T., & Vacca, W. D. 2005, ApJ, 623, 1115, ADS, arXiv:astro-ph/0412313 Daemgen, S., Hormuth, F., Brandner, W., Bergfors, C., Janson, M., Hippler, S., & Henning, T. 2009, A&A, 498, 567, ADS, 0902.2179 de Mooij, E. J. W., & Snellen, I. A. G. 2009, A&A, 493, L35, ADS, 0901.1878 Deming, D., Brown, T. M., Charbonneau, D., Harrington, J., & Richardson, L. J. 2005, ApJ, 622, 1149, ADS, arXiv:astro-ph/0412436 Bergfors, C., Brandner, W., Henning, T., & Daemgen, S. 2011, in Desidera, S., & Barbieri, M. 2007, A&A, 462, 345, ADS, IAU Symposium, Vol. 276, IAU Symposium, ed. A. Sozzetti, M. G. Lattanzi, & A. P. Boss, 397 -- 398, ADS Bergfors, C., Brandner, W., Daemgen, S., Biller, B., Hippler, S., Janson, M., Kudryavtseva, N., Geissler, K., Henning, T., & Kohler, R. 2012, arXiv.org, astro-ph.EP, 1209.4087. MNRAS accepted. Berta, Z. K. et al. 2012, ApJ, 747, 35, ADS, 1111.5621 Bodenheimer, P., Lin, D. N. C., & Mardling, R. A. 2001, ApJ, 548, 466, ADS Bramich, D. M. 2008, MNRAS, 386, L77, ADS, 0802.1273 Brogi, M., Snellen, I. A. G., de Kok, R. J., Albrecht, S., Birkby, J., & de Mooij, E. J. W. 2012, Nature, 486, 502, ADS, 1206.6109 Campo, C. J. et al. 2011, ApJ, 727, 125, ADS, 1003.2763 Carpenter, J. M. 2001, AJ, 121, 2851, ADS, arXiv:astro-ph/0101463 Castelli, F., & Kurucz, R. L. 2004, ArXiv Astrophysics e-prints, ADS, arXiv:astro-ph/0405087 Chan, T., Ingemyr, M., Winn, J. N., Holman, M. J., Sanchis-Ojeda, R., Esquerdo, G., & Everett, M. 2011, AJ, 141, 179, ADS, 1103.3078 Charbonneau, D. et al. 2005, ApJ, 626, 523, ADS, arXiv:astro-ph/0503457 Cowan, N. B., & Agol, E. 2011, ApJ, 729, 54, ADS, 1001.0012 Cowan, N. B., Machalek, P., Croll, B., Shekhtman, L. M., Burrows, A., Deming, D., Greene, T., & Hora, J. L. 2012, ApJ, 747, 82, ADS, 1112.0574 arXiv:astro-ph/0610623 Duquennoy, A., & Mayor, M. 1991, A&A, 248, 485, ADS Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935, ADS, 1005.4415 Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298, ADS, 0705.4285 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2012, ArXiv e-prints, ADS, 1202.3665 Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661, arXiv:astro-ph/0612671 Fortney, J. J., Shabram, M., Showman, A. P., Lian, Y., Freedman, R. S., Marley, M. S., & Lewis, N. K. 2010, ApJ, 709, 1396, ADS, 0912.2350 Fossati, L. et al. 2010, ApJ, 714, L222, ADS, 1005.3656 Gu, P.-G., Lin, D. N. C., & Bodenheimer, P. H. 2003, ApJ, 588, 509, ADS, arXiv:astro-ph/0303362 Hebb, L. et al. 2009, ApJ, 693, 1920, ADS, 0812.3240 Hinkle, K. H., Wallace, L., & Livingston, W. 2003, 35, 1260, ADS Husnoo, N. et al. 2011, MNRAS, 413, 2500, ADS, 1004.1809 Ibgui, L., Spiegel, D. S., & Burrows, A. 2011, ApJ, 727, 75, ADS, 0910.5928 Ichikawa, T. et al. 2006, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 6269, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ADS Jefferies, S. M., & Christou, J. C. 1993, ApJ, 415, 862, ADS 16 Crossfield et al. Jenkins, J. M., Caldwell, D. A., & Borucki, W. J. 2002, ApJ, 564, Press, W. H. 2002, Numerical recipes in C++ : the art of 495, ADS Katsuno, Y., Ichikawa, T., Asai, K., Suzuki, R., Tokoku, C., & Nishimura, T. 2003, Instrument Design and Performance for Optical/Infrared Ground-based Telescopes. Edited by Iye, 4841, 271 Kleinmann, S. G., & Hall, D. N. B. 1986, ApJS, 62, 501, ADS Knutson, H. A. et al. 2009, ApJ, 690, 822, ADS, 0802.1705 Kraus, A. L., & Hillenbrand, L. A. 2007, AJ, 134, 2340, ADS, 0708.2719 scientific computing, ed. Press, W. H., ADS Rayner, J. T., Toomey, D. W., Onaka, P. M., Denault, A. J., Stahlberger, W. E., Vacca, W. D., Cushing, M. C., & Wang, S. 2003, PASP, 115, 362, ADS Redfield, S., Endl, M., Cochran, W. D., & Koesterke, L. 2008, ApJ, 673, L87, ADS, 0712.0761 Rodler, F., Lopez-Morales, M., & Ribas, I. 2012, ApJ, 753, L25, ADS, 1206.6197 Rogers, J. C., Apai, D., L´opez-Morales, M., Sing, D. K., & Li, S.-L., Miller, N., Lin, D. N. C., & Fortney, J. J. 2010, Nature, Burrows, A. 2009, ApJ, 707, 1707, ADS, 0910.1257 463, 1054, ADS, 1002.4608 L´opez-Morales, M., Coughlin, J. L., Sing, D. K., Burrows, A., Apai, D., Rogers, J. C., Spiegel, D. S., & Adams, E. R. 2010, ApJ, 716, L36, ADS, 0912.2359 Lord, S. D. 1992, A new software tool for computing Earth's atmospheric transmission of near- and far-infrared radiation, Tech. rep., ADS Schoedel, R., Yelda, S., Ghez, A., Girard, J. H. V., Labadie, L., Rebolo, R., & Perez-Garrido, A. 2011, ArXiv e-prints, ADS, 1110.2261 Schweitzer, A., Hauschildt, P. H., & Baron, E. 2000, ApJ, 541, 1004, ADS, arXiv:astro-ph/0006049 Seager, S. 2011, Exoplanets, ed. Piper, S., ADS Seager, S., & Sasselov, D. D. 2000, ApJ, 537, 916, ADS, Luhman, K. L., & Mamajek, E. E. 2010, ApJ, 716, L120, ADS, arXiv:astro-ph/9912241 1005.2675 Maciejewski, G., Errmann, R., Raetz, S., Seeliger, M., Spaleniak, I., & Neuhauser, R. 2011, A&A, 528, A65, ADS, 1102.2421 Madhusudhan, N. et al. 2011, Nature, 469, 64, ADS, 1012.1603 Madhusudhan, N., & Seager, S. 2010, ApJ, 725, 261, ADS, 1010.4585 Mandell, A. M., Drake Deming, L., Blake, G. A., Knutson, H. A., Mumma, M. J., Villanueva, G. L., & Salyk, C. 2011, ApJ, 728, 18, ADS, 1011.5507 Mason, B. D., Wycoff, G. L., Hartkopf, W. I., Douglass, G. G., & Worley, C. E. 2001, AJ, 122, 3466, ADS Skrutskie, M. F. et al. 2006, AJ, 131, 1163, ADS Snellen, I. A. G., de Kok, R. J., de Mooij, E. J. W., & Albrecht, S. 2010, Nature, 465, 1049, ADS, 1006.4364 Stevenson, K. B. et al. 2010, Nature, 464, 1161, ADS, 1010.4591 Suzuki, R. et al. 2008, PASJ, 60, 1347, ADS Swain, M. et al. 2012, arXiv.org, astro-ph.EP, 1205.4736 Swain, M. R. et al. 2010, Nature, 463, 637, ADS, 1002.2453 Tokunaga, A. T., Simons, D. A., & Vacca, W. D. 2002, PASP, 114, 180, ADS, arXiv:astro-ph/0110593 Torres, G., Andersen, J., & Gim´enez, A. 2010, A&A Rev., 18, 67, ADS, 0908.2624 McLean, I. S. et al. 1998, in SPIE Conference Series, ed. A. M. Vacca, W. D., Cushing, M. C., & Rayner, J. T. 2004, PASP, 116, Fowler, Vol. 3354, 566 -- 578, ADS Miller-Ricci Kempton, E., & Rauscher, E. 2011, ArXiv e-prints, ADS, 1109.2270 Montalto, M., Santos, N. C., Boisse, I., Bou´e, G., Figueira, P., & Sousa, S. 2011, A&A, 528, L17, ADS, 1102.0464 Nagasawa, M., Ida, S., & Bessho, T. 2008, ApJ, 678, 498, ADS, 0801.1368 Nymeyer, S. et al. 2011, ApJ, 742, 35, ADS, 1005.1017 O'Donovan, F. T., Charbonneau, D., Harrington, J., Seager, S., Deming, D., & Knutson, H. A. 2009, in IAU Symposium, Vol. 253, IAU Symposium, 536 -- 539, 0909.3073 352, ADS, arXiv:astro-ph/0401379 Winn, J. N. 2010, ArXiv e-prints, ADS, 1001.2010 Winn, J. N. et al. 2008, ApJ, 683, 1076, ADS, 0804.4475 Yelda, S., Lu, J. R., Ghez, A. M., Clarkson, W., Anderson, J., Do, T., & Matthews, K. 2010, ApJ, 725, 331, ADS, 1010.0064 Zacharias, N., Urban, S. E., Zacharias, M. I., Wycoff, G. L., Hall, D. M., Monet, D. G., & Rafferty, T. J. 2004, AJ, 127, 3043, ADS, arXiv:astro-ph/0403060 Zhao, M., Monnier, J. D., Swain, M. R., Barman, T., & Hinkley, S. 2012, ApJ, 744, 122, ADS, 1109.5179
1501.05791
1
1501
2015-01-23T13:09:52
Forming chondrules in impact splashes. I. Radiative cooling model
[ "astro-ph.EP" ]
The formation of chondrules is one of the oldest unsolved mysteries in meteoritics and planet formation. Recently an old idea has been revived: the idea that chondrules form as a result of collisions between planetesimals in which the ejected molten material forms small droplets which solidify to become chondrules. Pre-melting of the planetesimals by radioactive decay of 26Al would help producing sprays of melt even at relatively low impact velocity. In this paper we study the radiative cooling of a ballistically expanding spherical cloud of chondrule droplets ejected from the impact site. We present results from a numerical radiative transfer models as well as analytic approximate solutions. We find that the temperature after the start of the expansion of the cloud remains constant for a time t_cool and then drops with time t approximately as T ~ T_0[(3/5)t/t_cool+ 2/5]^(-5/3) for t>t_cool. The time at which this temperature drop starts t_cool depends via an analytical formula on the mass of the cloud, the expansion velocity and the size of the chondrule. During the early isothermal expansion phase the density is still so high that we expect the vapor of volatile elements to saturate so that no large volatile losses are expected.
astro-ph.EP
astro-ph
The Astrophysical Journal, 794:91 (12pp), 2014 October 10 Preprint typeset using LATEX style emulateapj v. 5/2/11 FORMING CHONDRULES IN IMPACT SPLASHES I. RADIATIVE COOLING MODEL Cornelis Petrus Dullemond, Sebastian Markus Stammler Institute for Theoretical Astrophysics, Heidelberg University, Albert-Ueberle-Strasse 2, 69120 Heidelberg, Germany Lund Observatory, Department of Astronomy and Theoretical Physics, Lund University, Box 43, 22100 Lund, Sweden The Astrophysical Journal, 794:91 (12pp), 2014 October 10 Anders Johansen ABSTRACT The formation of chondrules is one of the oldest unsolved mysteries in meteoritics and planet for- mation. Recently an old idea has been revived: the idea that chondrules form as a result of collisions between planetesimals in which the ejected molten material forms small droplets which solidify to be- come chondrules. Pre-melting of the planetesimals by radioactive decay of 26Al would help producing sprays of melt even at relatively low impact velocity. In this paper we study the radiative cooling of a ballistically expanding spherical cloud of chondrule droplets ejected from the impact site. We present results from a numerical radiative transfer models as well as analytic approximate solutions. We find that the temperature after the start of the expansion of the cloud remains constant for a time tcool and then drops with time t approximately as T ≃ T0[(3/5)t/tcool + 2/5]−5/3 for t > tcool. The time at which this temperature drop starts tcool depends via an analytical formula on the mass of the cloud, the expansion velocity and the size of the chondrule. During the early isothermal expansion phase the density is still so high that we expect the vapor of volatile elements to saturate so that no large volatile losses are expected. Subject headings: chondrules, radiative transfer 1. INTRODUCTION The formation of chondrules is one of the funda- mental questions of meteoritics and planet formation. Chondrules are the 0.1· · · 1 mm-size once-molten sili- cate spherules found abundantly in primitive meteorites known as chondrites (Jones, Grossman & Rubin 2005; Sears 2004; Davis et al. 2014). Most of these chondrites in fact consist predominantly of these chondrules, so the melt-producing events that created them must have been extremely common during the first few million years of the solar system. Yet there is confusing and conflicting evidence as to what these events might have been. Boss (1996) published an overview of the status of the discus- sion at that time, though significant new developments have occurred since then. Many theories have been put forward over the last half a century. One theory involves impact melt sprays. Put forward by Urey (1953) and refined by Kieffer (1975) this model states that high-velocity impacts (& 3· · · 5 km/s) could lead to sprays of impact melt that produce droplets which solidify into chondrules. While these required colli- sion velocities are high, a small fraction of impacts might acquire such velocities (Bottke et al. 1994). Perhaps the so-called "Grand Tack"-scenario of Walsh et al. (2011), in which Jupiter temporarily entered the asteroid belt re- gion before migrating back outward, might produce such high velocities. However, Taylor et al. (1983) put forward a number of arguments against the impact-melt scenario, some of which were based the differences between chon- drites and lunar impact regolith. Zook (1980) suggested that if the interiors of the collid- ing bodies are already in a molten state due to 26Al de- cay, then the required impact velocities to create sprays of melt are much lower, and thus more consistent with ex- pectations of the average relative velocities between plan- etesimals. Sanders & Taylor (2005), Hevey & Sanders (2006) and Sanders & Scott (2012) follow up on this idea of pre-molten impactors and back it up with models and meteoritic evidence. Sanders & Scott (2012) argue that this scenario is hard to avoid, given that in the first 2.5 million years most planetesimals were internally nearly fully molten by 26Al decay heat, and that collisions be- tween planetesimals were extremely common. Each col- lision would almost certainly release substantial amounts of molten rock into the nebula in the form of sprays of lava droplets, and it is natural to assume that these may be chondrules. They argue that the near-solar composi- tion of chondrules can be explained by the vigorous con- vection inside the molten planetesimals that may slow down iron/nickel-core formation, thus keeping the melt solar. Some degree of differentiation would then in fact explain the low iron-content L and LL chondrites. Recently, Asphaug et al. (2011) performed Smooth Particle Hydrodynamics simulations for such impacts to demonstrate the dynamics of this scenario. In addition they proposed a simple way to calculate the melt droplet size by equating the released enthalpy after the collision to the surface energy of the droplets. There are many alternative scenarios proposed in the literature. Perhaps the most popular model is the flash heating by nebular shocks (Hood & Hor´anyi, 1991). De- sch & Connolly (2002) and Ciesla & Hood (2002) de- veloped detailed 1-D radiative shock models with dust particles and chondrules interacting both radiatively and frictionally with the gas. They showed how the radia- tive shocks exhibit temperature spikes of mere tens of 2 Dullemond, Stammler, Johansen seconds (with cooling rates > 104 K/hr) that would be good candidates for chondrule-forming events. After the main shock temperature spike their model exhibits a fur- ther cooling at intermediate cooling rate (∼50 K/hr), which cools the chondrules to sufficiently low tempera- ture in a sufficiently short amount of time. This model seems to produce the flash-heating events required for turning "dust bunnies" into chondrules. However, so far the shock scenario still has several unresolved issues (e.g., Desch et al. 2012; Boley et al. 2013; Stammler & Dulle- mond 2014). Also some issues are raised about whether large scale shocks in the optically thick solar nebula can explain the time scales involved in chondrule formation (Stammler & Dullemond 2014). It is noteworthy that there exist several other nebular flash heating models, most notably flash heating by nebular lightning (Gibbard et al. 1997) and flash heating by energy dissipation in current sheets forming in MHD turbulence (Hubbard et al. 2012). One important constraint on chondrule formation models is that these heating events were very short-lived compared to any other nebular time scales. By compar- ing textures of chondrule to those obtained in furnace experiments it can be inferred that chondrules must have cooled from the liquidus temperature down to below the solidus temperature in a matter of hours (Hewins 1983; Hewins & Connolly 1996; see also references in Morris & Desch 2010 and the excellent review paper of Desch et al. 2012). In other words: the chondrule forming events must have been flash-events. On the other hand, un- der optically thin conditions a molten chondrule would radiatively cool in about a second, which would be too fast. Another constraint that a chondrule formation model must fulfill is the retention of volatile elements such as Fe, Mg, Si, Na and K. Chondrules are not observed to have low abundances of these elements. In particu- lar for the highly volatile elements Na and K this is a puzzle, because any heating event that heats a (proto- )chondrule above ∼ 1700 K and keeps these chondrules above that temperature for more than a few minutes will cause most of the Na and K to evaporate out of the chondrule. Alexander et al. (2008) argue that this means that chondrules must have formed in regions that are extremely dense in solids, much more so that typi- cal dust concentration mechanisms in the protoplanetary disk can achieve. According to their calculations such regions are so dense in solids that they must be self- gravitating. Morris et al. (2012) instead propose that chondrules formed behind the bow shock of a fast moving planetary embryo, and that the embryo's atmosphere is rich in such volatile elements caused by outgassing from the embryo's interior, thus providing the necessary vapor pressure to keep the chondrules volatile-rich. In this paper we will focus on the impact splash hy- pothesis, either the low-velocity pre-molten planetesimal version or the high-velocity impact-melt version. We as- sume that after the collision between two planetesimals a cloud of molten droplets of magma was released. Since soon after the ejection of this cloud of magma droplets the internal pressure of the cloud would have dropped to near-zero (there is, to good approximation, only vac- uum between the droplets), the cloud would simply ex- pand ballistically. Initially the density of the cloud is so high that the cloud is completely optically thick and no radiation can escape from its interior. There is also no adiabatic cooling because of the lack of pressure. So the temperature of the magma droplets will initially stay roughly constant in time. As the cloud expands, how- ever, the optical depth drops and eventually the cloud will start to cool radiatively. During the early stages the density is very high and the evaporation of volatile ele- ments will be saturated (as was argued by Alexander et al. 2008). Once the temperature drops below the solidus, volatiles can no longer escape. In this first paper we intend to compute how the tem- perature behaves as a function of time after the impact. To do this we set up a simple model: that of a ho- mologously ballistically expanding homogeneous spher- ical cloud of lava droplets. We compute the temperature of the chondrules as a function of time t and comoving ra- dial location inside the cloud r/rcloud by solving the time- dependent radiative transfer equation. We also present an analytic approximate solution. While the spherical cloud model is not an accurate model of the complex shape of an impact splash, it can be regarded as a model of part of the impact splash. As such we believe that it will give reasonable estimates of the radiative cool- ing behavior of the impact splash as a function of model parameters such as the total mass of the cloud and its ex- pansion velocity. The total mass of the cloud tells some- thing about the masses of the two colliding bodies while the expansion velocity of the cloud tells us something about the impact velocity. 2. EXPANDING CLOUD MODEL When lava droplets are produced in a collision between two planetesimals, they will disperse away from the im- pact site in a ballistic way. Let us call this the "impact splash". If the impact velocity is larger than the escape velocity of the two colliding planetesimals this ballistic (pressureless) expansion will be linear (i.e. the velocity of expansion will not change with time). We regard the impact splash as an expanding cloud of lava droplets (chondrules-to-be). This cloud is not necessarily cen- tered on the impact site; it can also move away from the impact site. The radius of the cloud increases linearly with time t and thus the density of the cloud will drop as 1/t3. Depending on the complexity of the geometry of the impact splash (see e.g. Asphaug et al. 2011) we can also consider the splash to be multiple smaller clouds. In either case, each of these clouds will expand linearly. The scenario is pictographically shown in Fig. 1. If the impact is at low velocity, then the gravitational pull of the surviving (or merged) body can cause some (or most) of the material to re-accrete. In this case the 1/t3 expansion will cease and turn into recompression. Initially, however, the linear expansion and 1/t3 density drop will still be a good description. To compute the temperature of the droplets in these clouds as a function of time after impact we must solve the time-dependent equation of radiative transfer. This is a very non-local problem, because a cooling droplet at location ~x1 radiates away some heat into the form of infrared radiation, which can then be absorbed by an- other droplet at location ~x2 6= ~x1 which is then heated. The droplets are thus radiatively coupled over large dis- tances. Solving this problem in 3-D for complex cloud Forming chondrules in impact splashes 3 Pre−collision After collision: Some parts survive. Debris clouds (e.g. clouds of molten lava droplets) get ejected. The debris clouds (here modeled as spherical clouds) expand linearly as they move away from the impact site. Fig. 1.- Pictographic representation of the model. Left: Two planetesimals approach each other and are about to collide. Middle: After the collision some parts of the planetesimals may survive, but some parts are destroyed and dispersed in a cloud of debris. We model this dispersing cloud as a set of spherical clouds. If the debris consists of melt, the cloud will consist of molten droplets (chondrules). Right: As the clouds move away from the point of impact they expand linearly. The average distance between the chondrules increases linearly with time. geometries is somewhat challenging. the cloud is constant and outside of the cloud is zero. The density of the cloud as a function of time is then We believe, however, that much can already be learned from a simple spherical cloud model, for which the radia- tive transfer problem can be solved to high precision and reliability with the 1-D tangent-ray variable eddington factor method. We will describe the method in Section 4. The three main parameters of the model are: Mcloud , T0 , vexp (1) where Mcloud is the total mass worth of chondrules (lava droplets) in our spherical cloud, T0 is the initial tempera- ture of the chondrules and vexp is the expansion velocity of the outer radius of the cloud. To get a feeling for the numbers it is more convenient to express the mass of the cloud in terms of the radius Rmelt,0 of a ball of magma of the same mass Mcloud: Mcloud = 4π 3 ξchonR3 melt,0 (2) where ξchon is the material density of the chondrule droplets and thus, by definition, the material density of the hypothetical ball of magma. For a linearly expanding cloud the radius of the cloud is Rcloud(t) = vexpt (3) where the time t is the time since the impact and the ejection of the cloud of melt droplets, assuming perfectly ballistic (pressureless) expansion and of course assum- ing large enough t that Rcloud(t) ≫ Rmelt,0 so that the droplets are clearly separated from each other. The ve- locity profile inside the cloud is v(r, t) = vexp r Rcloud(t) (4) which is of course only valid for r ≤ Rcloud(t). Let us define a coordinate r centered on the center of the spherical cloud. We assume that the density within ρcloud(r, t) =(cid:26) 3Mcloud 4πR3 cloud(t) 0 for r ≤ Rcloud(t) for r > Rcloud(t) (5) From here on we write only ρcloud(t) ρcloud(r, t). The time-dependence of ρcloud(t) is instead of ρcloud(t) = 3Mcloud 4πv3 exp 1 t3 (6) Let us consider chondrules (lava droplets) of radius achon and material density ξchon. We take ξchon = 3.3 g/cm3 for our model. The mass of a chondrule is then mchon = 4π 3 ξchona3 chon The number density of chondrules is then nchon(t) = ρcloud(t) mchon (7) (8) The geometric cross section of a chondrule is πa2 chon. Let us assume that the chondrule has zero albedo. Then the absorption cross section equals the geometric cross section. The opacity (i.e. the cross section per gram) is then κ = πa2 chon 4πξchona3 chon/3 = 3 4 1 ξchonachon (9) We will assume that the opacity of any possible vapor between the chondrules will be low compared to that of the chondrules so that it can be ignored. The optical depth from the center of the cloud to the edge is τ (t) = ρcloud(t)κRcloud(t) = 3 4π M κ v2 exp 1 t2 (10) The rate of cooling will strongly depend on this optical depth. 4 Dullemond, Stammler, Johansen vexp Model Rmelt,0 100 m/s 1 km 0.1 km 1000 m/s 10 km 100 m/s 0.01 km 1000 m/s F1 F2 F3 F4 T0 tcool 2000 K 27 min 2000 K 16 sec 2000 K 2000 K 7 h 1 s τcool 92 9.2 366 2.3 TABLE 1 The model parameters (first three columns) for the main ("fiducial") models presented in this paper. The cloud mass Mcloud can be derived from Mcloud = 4π melt,0 (cf. Eq. 2). The last two columns are computed from 3 ξchonR3 these parameters: the cooling time tcool (the time after the impact when the cloud starts to cool, see Eq. 16) and the optical depth at that time τcool ≡ τ (t = tcool) (see Eq. 20). In this paper we will present our results based on a set of fiducial models as well as parameter scans. The parameters of the fiducial models are listed in Table 1. 3. ANALYTIC ESTIMATE OF THE COOLING BEHAVIOR OF THE CHONDRULE CLOUD As the cloud expands it can start to radiate away en- ergy. A fully fledged time-dependent radiative cooling computation will yield temperature as a function of time and space: T (r, t). We will compute this in Section 4. A first estimate of the cooling behavior of the cloud can already be made with pen and paper. The cloud will initially be optically thick: τ ≫ 1. The luminosity of the cloud is then L = 4πR2 cloudσT 4 eff (11) where σ is the Stefan-Boltzmann constant and Teff is the effective surface temperature at r = Rcloud. In the above equation, and from here onward, we omit the (t) for aesthetic reasons, but the time-dependence is still as- sumed. During the cooling phase the effective surface temperature will be lower than the central temperature. A commonly used estimation of this effect is: L ≃ 4πR2 cloudσT 4 1 τ (12) where T is now the central temperature and 1/τ is the correction factor to account for the lower temperature of the surface Teff, and is based on radiative diffusion theory which states that T ≃ τ 1/4Teff (for τ ≫ 1). To compute the change of the temperature as a result of the radiative loss given by Eq. (12) we must use the heat capacity formula. The total thermal energy stored in the chondrule cloud (approximating it, for the moment, as an isothermal cloud) is: E = McloudcmT (13) where cm is the mass-weighted specific heat of the lava droplet. A value of cm = 107 erg g−1 K−1 is a reasonable value which we will adopt here. The radiative loss time scale can now be defined as trad(t) = E L = Mcloudcmτ 4πR2 cloudσT 3 (14) becomes smaller. It is therefore to be expected that at early times the temperature of the chondrules remains constant. Note that since the cloud consists of liquid drops that stay at a constant volume and move away from each other, there is no adiabatic cooling involved here. Any potential vapor may adiabatically cool, but we will ignore this effect in this paper, assuming that the total mass in vapor is always small compared to the mass in droplets. At some point in time, however, trad becomes smaller than t and the chondrules start to cool. Based on the results of the true radiative transfer calculations of Sec- tion 4 it turns out that near the center of the cloud the transition from the initial constant temperature phase to the temperature-decline phase occurs roughly at a time tcool defined by trad(tcool) = 5 tcool Inserting Eqs.(14,12,13,10) into Eq. (15) yields tcool =(cid:18) 1 5 3 (4π)2 M 2 cloudcmκ expσT 3 v4 0 (cid:19)1/5 (15) (16) Note that in Eq. (15) the factor of 5 is purely empirical. The estimate we make here is essentially a dimensional analysis in which proportionality factors have to be cal- ibrated against more exact calculations (in our case the full radiative transfer calculation). Roughly for t < tcool the temperature stays constant while for t > tcool the temperature drops with time. From now on we shall define the tcool as the one cal- culated for the center of the cloud. Near the surface the cooling sets in earlier. The temperature decline with time can also be es- timated, at least up until the point where the cloud becomes optically thin, after which our approximation breaks down. The way to make this estimate is to solve the cooling equation dT (t) dt = − T (t) tcool(t) (17) where tcool(t) is the cooling time given by Eq. (16) but with T0 replaced by T (t). The reasoning behind this is that if we would instead use trad(t) (which might look more reasonable at first sight) we will quickly cool to temperatures for which condition trad . t is again no longer fulfilled. This is because trad ∝ T −3 and thus very rapidly rises with declining temperature. If we insert Eq. (16) with T0 replaced by T into Eq. (17) we obtain a differential equation that can be solved by separation of variables. The solution is T (t > tcool) = T0(cid:20) 3 5 t tcool + 2 5(cid:21)−5/3 (18) where here tcool is again the original one with T0. We assume that for t < tcool we have: T (t < tcool) = T0 (19) This time scale varies with time: it is very long at early times because the cloud is then still extremely optically thick. At late times the cloud is optically thinner and the radiation can more freely escape (Eq. 12) and trad The solution Eqs. (18,19) holds true for the central tem- perature. But most of the mass of a homogeneous sphere resides in the outer parts. For the temperature at r = 0.8Rcloud and r = 0.9Rcloud (roughly the radii Forming chondrules in impact splashes 5 which is the time between the start of the cooling at T = T0 and the time when the temperature has dropped below the solidus temperature 1400 K. To first order we can write t1400 ≃ T0 − 1400K T (t = tcool) =(cid:18)1 − 1400K T0 (cid:19) tcool (22) cloud and as v−4/5 The quantities T and t1400 can be compared to the con- straints coming from the analysis of textures of chon- drules, which put time limits on the cooling process of order of hours, or more precisely: cooling rates in the range 10 · · · ∼ 3000 K/hr (see Morris & Desch 2010 and Desch et al. 2012 and references therein, though see also Miura & Yamamoto 2014 for a different view based on theoretical modeling of crystallization), indicated by the grey area in Fig. 3. According to Eq. (16) tcool goes as M 2/5 exp . There is not very much freedom of choice of T0: it must lie somewhere between 1770 and 2120 K (see Morris & Desch 2010 and references therein). The κ does not have too much wiggle room either: the radii of chondrules are known. This means that con- straints on tcool from textural analysis directly set lim- its on the ratio Mcloud/v2 exp or equivalently on the ratio R3 If, for example, we choose achon = 0.03 cm and T0 = 2000 K we obtain κ = 7.6 cm2/g. If we require, for instance, the cooling time scale to be tcool = 10 minutes, then we find that Mcloud/v2 exp ≃ 107 g s2 cm−2. This would be fulfilled e.g. for vexp = 1 km/s and Mcloud = 1017 g (a mass corresponding to a ball of magma of roughly Rmelt,0 =2 km radius). It would also be fulfilled by a smaller mass and smaller velocity: e.g. for vexp = 10 m/s and Mcloud = 1013 g (Rmelt,0 =0.1 km radius). Such small mass could either mean that upon impact only a small fraction of the debris is in the form of chondrule droplets, or it could simply mean that the largest closed unit of the droplet splash (subcloud) is so small, but many such expanding cloudlets of chondrules are ejected. The elongated streams of debris found in Asphaug et al. (2011) could be regarded as a string of such smaller mass cloudlets. melt,0/v2 exp. This is shown more directly in Fig. 4 which shows the parameter space of the model, with the grey area again showing the models which give cooling rates between 10 and 3000 K/hr. Fig. 2.- The temperature evolution of the expanding cloud model according to the analytical estimate of Eqs. (18,19), shown in black. In grey are, for comparison, the full radiative transfer so- lutions from Section 4 for the fiducial model F1 (see Table 1). The temperatures are shown at three positions in the cloud: the center, at 80% of the radius and at 90% of the radius (near the surface of the cloud). The time scale is scaled to tcool (Eq. 16). The analytic solutions are only valid as long as τcool ≫ 1 (cf. Eq. 20). of half mass and of 75% mass respectively) the best fit- ting solution to the real solutions of Section 4 are similar to Eq. (18) but with the term 2/5 replaced by 3/5 and 3.8/5 respectively. These analytic estimates of the tem- perature as a function of time are plotted in Fig. 2. As one can see, although the time after the impact at which the cooling stars is different depending on whether you look at the center or at the edge of the cloud, the typical cooling rate after the cooling begins is similar throughout the cloud. It is important to remind ourselves that the cooling normally starts well before the cloud becomes optically thin. Optically thin cooling would be extremely fast: a single chondrule would cool within about 1 second. The protracted cooling over a time scale of hours occurs be- cause the high optical depth (even after t = tcool) acts as a kind of blanket that keeps the chondrules warm, al- beit a blanket that is thinning over time as the optical depth drops. To get a feeling for this we can calculate the optical depth at the time t = tcool from Eq. (10): (20) 4. TIME-DEPENDENT RADIATIVE TRANSFER So far we have only made an estimate of the cooling behavior of the cloud of liquid chondrules. Let us now calculate the temperature profile with a full treatment of time-dependent radiative transfer. The time-dependence comes in only due to the time it takes the lava droplets to convert their heat into radiation, not in the form of light- travel time (which typically only plays a role for media at temperatures above 105K). This can be easily verified by comparing the radiative energy density aT 4 to the material energy density ρcmT , which in our case always satisfies aT 4 ≪ ρcmT . Therefore the radiative transfer equation itself is stationary, and the time-dependence is in the equation for heating/cooling of the chondrules. 4.1. Equations of time-dependent radiative transfer τcool ≡ τ (tcool) = 33/552/5M 1/5 cloudκ3/5σ2/5T 6/5 0 (4π)1/5v2/5 expc2/5 m So Eqs. (18,19), which are based on the assumption of the "blanket effect" of high optical depth, are expected to be valid for the case τcool ≫ 1. In table 1 one can see that τcool is well above 1 for all models, except the most extreme model F4 which has anyway a much too small cooling time to be consistent with chondrule textures. The analytic model of Eqs. (18,19) tells us that for t > tcool the temperature drops off suddenly on a typical time scale of tcool. The temperature decay rate at t = tcool is: T (t = tcool) ≡ = − T0 tcool (21) dT dt (cid:12)(cid:12)(cid:12)(cid:12)t=tcool This is plotted, together with tcool, in Fig. 3. It is also useful to define another kind of cooling time scale t1400 6 Dullemond, Stammler, Johansen Fig. 3.- The cooling time tcool (Eq. 16) as a function of param- eters of the model for T0 = 2000 K and achon = 0.03 cm. Note that the time to cool down from 2000 K to 1400 K (t1400) is, ac- cording to Eq. (22), about four times smaller than tcool. On the right axis the corresponding cooling rate dT /dt in K/hr is plotted. The grey area is the observed range of chondrule cooling rates (see main text). Note that the values of tcool are (by definition) for the center of the cloud. The regions of the cloud more closely to the surface will start to cool earlier, but the cooling rate is almost the same. to the geometric cross section, independent of frequency, because the chondrules are much larger than the wave- length of infrared radiation. In this case the frequency- dependence of the radiative transfer does not need to be considered. Eq. (23) is called the "formal transfer equation", and describes the transport of radiation along rays of direction ~n. It is easy to integrate if the value of B(~x) = B(T (~x)) is known everywhere, but since we want to calculate the temperature T (~x), this function is part of the thing we want to solve. In our 1-D spherically symmetric setting Eq. (23) can be written as µ dIµ(r) dr + 1 − µ2 dIµ(r) r dµ where we shortened = α(B(T (r)) − Iµ(r)) (25) ρκ =: α (26) Here µ = cos(θ) where θ is the angle between the ray along which the radiation is followed and the radially out- ward pointing unit vector. The Eddington factor method of solving this equation relies on the definition of the first three angular moments of the intensity at each location r: J(r) = H(r) = K(r) = 1 1 1 −1 2Z +1 2Z +1 2Z +1 −1 −1 Iµ(r)dµ Iµ(r)µdµ Iµ(r)µ2dµ (27) (28) (29) Here J is called the mean intensity and H is the flux divided by 4π. By integrating Eq. (25) over 1 2 dµ after multiplying it by 1 and by µ respectively one obtains the first two moment equations: 1 r2 d(r2H) dr 3f − 1 = α(B(T ) − J) dK dr + r J = −αH (30) (31) Fig. 4.- The parameter space of the model. The grey area shows where the solutions have cooling rates in the range observed for chondrules. In the bottom-right the region is shown where τcool < 10, where the analytic solution becomes invalid. The formal radiative transfer equation is (see e.g. Mi- halas & Mihalas 1999): ~n · ~∇I(~x, ~n) = ρκ(B(~x) − I(~x, ~n)) (23) where I is the frequency-integrated mean intensity in erg s−1 cm−2 ster−1, ~x is the position vector in space, ~n is the unit vector of direction of the radiation, and finally B the frequency-integrated Planck function given by B(T ) = σ π T 4 (24) We will use frequency-integrated quantities because the opacity of a chondrule is expected to be roughly equal These are two equations with three unknowns. To close this set of equations we write K(r) = f (r)J(r) (32) where f is called the "Eddington factor". If f (r) is known, then the two moment equations are in closed form and can be solved for J(r) and H(r). From the moment equations alone we cannot know what f (r) is for all r, but for certain limiting cases we know their values: For optically thick media f = 1/3. Outside of the cloud for r → ∞ we will get f → 1, which is the free-streaming limit. However, for general r and for gen- eral values of the optical depth we must employ another method of calculating f (r). We do this by integrating, for the current value of T (r) (and thus B(T (r))), the formal transfer equation Eq. (25) using the "tangent ray method" (see e.g. Mihalas & Mihalas 1999). By inte- grating the resulting intensities over angle we calculate the current estimates of J(r) and K(r). Let us call these Forming chondrules in impact splashes 7 Jfte(r) and Kfte(r) (where "fte" stands for "formal trans- fer equation"). Our current estimate of f (r) is then f (r) = Kfte(r) Jfte(r) (33) This is then the f (r) function we stick into Eq. (32), so that the moment equations can be solved for J(r) and H(r). We also need to impose boundary conditions at the inner and outer edge. At the inner edge we take H = 0 (zero flux). At the outer edge we could set H = J, which is valid if the outer edge of our computational domain is at r → ∞. For finite outer radius we set instead H(rout) = Hfte(rout) Jfte(rout) J(rout) (34) Note that we can choose to set the outer edge of our computational domain rout at rout = Rcloud, but we can also set it at rout > Rcloud, as long as we properly set the boundary condition at r = rout according to Eq. (34). We can now combine the two moment equations into the following form: 1 α 1 r2 d dr (cid:26) r2 α (cid:20) d(f J) dr + 3f − 1 r J(cid:21)(cid:27) = J − σ π T 4 (35) This can be solved, together with the boundary condi- tions, using a matrix equation. The details of this are discussed in Appendix A. Next we must compute how the temperature T (r) re- acts to the radiation field, or in other words, how T (r) radiatively cools with time. The energy equation is ρcm dT dt = 4πα(cid:16)J − σ π T 4(cid:17) (36) where ρ is given by Eq. (6). The d/dt operator is the comoving derivative, i.e. the time derivative is computed for a given chondrule. Eq. (36) is a local equation at each location. The non-locality of the heating/cooling is only through the non-locality of the equation for J(r) (Eq. 35). The way we solve these equations numerically is by setting up a radial grid of N gridpoints ri (with 1 ≤ i ≤ N ) upon which we define the mean intensity Ji and the temperature Ti. Since the cloud is expanding we let the gridpoints move along with the material (Lagrangian approach) so that the location of grid point i at time t is ri(t) = ηivexpt where η is the dimensionless radial coordinate η = r Rcloud(t) = r vexpt (37) (38) The values of ηi do not change with time. Each chondrule stays at constant η, and so the comoving time deriva- tive d/dt used in e.g. Eq. (36) becomes simply the time derivative of that quantity at a given ηi grid point. In principle one could imagine a simple time-dependent integration method for the combination of numerical Eqs.(35,36). We know J n i at time t = tn, and we can solve for T n+1 by taking an explicit Euler inte- gration step of Eq. (36). Then we can solve the moment equations (because now we have Bn+1 )) and i and T n = B(T n+1 i i i i obtain J n+1 . The problem with this scheme is that the required time steps can become very small due to nu- merical stiffness. A more robust method is to integrate the complete system Eqs.(35,36) using an implicit inte- gration scheme. This is discussed in appendix A. 4.2. Results of the time-dependent radiative transfer models The results of the time-dependent radiative transfer calculations are shown in Fig. 5. These results confirm the estimates made in Section 3. We find that the shape of the cooling curves is mostly the same for all models, except for the time scaling, which we know from the ana- lytically expression of tcool (Eq. 16). Only for the rather extreme (and presumably unrealistic) case such as fidu- cial model F4 with Rmelt,0 = 0.01 km and vexp = 1000 m/s we find from the full radiative transfer model that the cooling is slower than the analytic solution. This is because in that case the condition that τcool (Eq. 20) is ≫ 1 is broken and the cooling time scale will then not be determined by the optical depth decline but by the effec- tiveness by which a chondrule can convert its heat into radiation. These cases, however, imply typically time scales of seconds rather than hours, and these are any- way too short to be consistent with chondrule textures. It seems that the analytic solutions are fairly accurate when τcool & 10, which is always true for the interesting regime of parameter space. From Figs. 2 and 5 we see the model systematically predicts a period of constant temperature with a dura- tion similar to the later cooling phase. This is true at least near the center of the cloud. Near the surface this temperature plateau is shorter, but the cooling time is similar. In Fig. 6 we show, for model F1, how the temperature looks as a function of radial position in the cloud, for different times after the collision. Since the cloud is ex- panding, which would make it difficult to compare the temperature profiles at different times, we use the radial coordinate relative to the outer cloud radius. One sees that the outer regions start cooling earlier because the optical depth to the surface is smaller. Also one sees that a negative radial temperature gradient is produced which causes the radiative diffusive energy transport to the surface. The cooling curve from our model is significantly dif- ferent from the ones predicted from shock heating (e.g. Morris & Desch 2010, Morris et al. 2012). The shock heating models predict a radiative preheating phase, a temperature spike with rapid cooling, and depending on geometric conditions followed by a slower cooling phase. In our model the temperature is high from the start (or for high-speed collisions: upon impact), stays relatively constant for some time and then drops over a similar time scale. 5. DISCUSSION The scenario of splashes of melt droplets originating from colliding pre-molten planetesimals has been dis- cussed at length in several papers by Sanders and cowork- ers (e.g. Sanders & Scott 2012) and the paper by Asphaug et al. (2011). These papers discuss how this scenario holds up agains the many meteoritic constraints. We will not repeat these arguments here, but refer for those 8 Dullemond, Stammler, Johansen Fig. 5.- The temperature evolution of the fiducial expanding cloud models F1 (top-left), F2 (top-right), F3 (bottom-left) and F4 (bottom-right). See Table 1 for the parameters. The temperatures are shown at three positions in the cloud: the center, at 80% of the radius and at 90% of the radius (near the surface of the cloud). The vertical line shows the time tcool of Eq. (16). 5.1. Splash geometry Our model describes a simple spherical expanding cloud as a model for (part of) the splash resulting from colliding bodies. It is obvious that this is a drastic simpli- fication. The SPH model of Asphaug et al. (2011) shows a much more realistic geometry involving, among other complexities, tube-like geometries. A proper 3-D time- dependent radiative transfer simulation done in concert with the gravito-hydrodynamic simulation of the impact splashing is required to know precisely what the cooling behavior is. Nevertheless we predict that our result can be used to estimate the results of those detailed compu- tations: the tube-like geometry can be approximated as a homologously expanding cylinder, expanding not only in the 2 perpendicular directions but also in the longitu- dinal direction. Its properties are then not too much dif- ferent from a chain of spherically expanding clouds with diameters similar to the diameter of the tube, except that these spherical cloud cannot cool in all directions (4π) but only in part of the sky (in the other part it will heat its neighbors and be heated by its neighbors). We would then expect that the cooling time tcool will be a bit longer but not by a large factor. Also, one can expect some parts of the splash to be ejected at slow speed compared to the escape speed, so that it may fall back to the remaining (unsplashed) Fig. 6.- The radial temperature temperature structure of fidu- cial model F1 for different times. The horizontal axis is the radius relative to the outer cloud radius. Note that the temperature pro- file extends beyond the outer radius of the cloud because even if that region is empty, a chondrule residing there would have a well- defined temperature. For our purposes, however, only the temper- ature within the cloud is relevant. discussions to those papers. Instead, we will focus on the discussion of aspects related to our splash cooling model. Forming chondrules in impact splashes 9 bodies and accrete on them. This evidently breaks the assumption of homologous expansion, and as a result will also deform any initially spherical sub-cloud into a more pancake-like shape. The idea of regarding the splash cloud as a collection of spherical sub-clouds is then clearly no longer valid, not even in an approximative way. In such cases a full 3-D radiation-gravito-hydrodynamic simulation of the splash is unavoidable. However, multi- dimensional radiation-hydrodynamics is an enormously challenging problem, in particular for problems involving such extremely complex geometries. It is an interesting challenge for the future. For getting order-of-magnitude estimates we may be forced to stick, for the time be- ing, with simple models such as the one presented in this paper. 5.2. Chondrule sizes According to the model by Asphaug et al. (2011), by which the lava drop size is calculated from the initial pressure under which the magma was in the pre-molten planetesimals before the collision, the planetesimal sizes must have been 10 km or larger. Since our cloud of lava droplets can be one of a number of such clouds ejected from the impact, the Rmelt,0 of our model cannot be di- rectly compared to that > 10 km size lower limit from Asphaug. But it does give an indication that, if we adopt their model of the chondrule sizes, it seems reasonable to be looking to values of roughly the order of Rmelt,0 ≃ 10 km or larger. To elucidate our reasoning: If we collide two fully molten planetesimals of 20 km radius, and we model the splash with 10 spherically expanding droplet clouds, then we would get Rmelt,0 ≃ 12 km. In Fig. 3 this means we are looking at the region of the diagram above the dotted line. There is a caveat, however: What tells us that it is just 10 clouds, not a 1000? This brings us to the complicated issue of the actual geometry of the splash. If we "de- compose" the splash geometry into many small spherical clouds we must assure that these clouds can radiatively cool to the outside. In other words: we cannot consider a big spherical cloud as consisting of many smaller sub- clouds, because they will irradiate each other and not cool much. But if the cloud geometry is a long extended streak, it can be considered as consisting of a chain of spherical subclouds. We are here confronted with the limitation of our spherical expanding cloud model. 5.3. Partial pre-melting The reason why pre-melting seems to be necessary for the impact splash scenario to work is that very high speed collisions (& 5 km/s) needed to produce melt from cold planetesimals are presumably rare. However, pre-melting may lead to differentiation of the planetesimals, causing the iron and nickel to sink into the core and produc- ing a basaltic mantle. Chondrules formed from such ob- jects would be very non-solar in composition. Although Sanders & Scott (2012) argue that convection reduces the efficiency of the differentiation and that some amount of differentiation is in fact consistent with e.g. L and LL chondrites, it seems that this issue is not yet conclusively solved and more detailed and quantitative modeling of the differentiation process in these molten planetesimals may need to be done. However, a middle way is if planetesimals were pre- heated to elevated temperatures below the melting point. This is to be expected in particular for later generation planetesimals (i.e. planetesimals formed at a time when a substantial fraction of the 26Al has already decayed). Lower velocity collisions might then be sufficient to add the remaining energy needed to produce melt. The prob- lem with this scenario is that it requires fine-tuning: The planetesimal must have formed at a time when the re- maining abundance of 26Al is low enough not to melt the planetesimal but high enough to heat it to tempera- tures only slightly below the melting temperature. This might be too much coincidence. On the other hand, we know that some planetesimals were strongly differentiated: the parent bodies of iron and basaltic meteorites. If chondrules formed from col- lisions between planetesimals, one would expect some chondrules to have been formed from collisions with such differentiated planetesimals. Some chondrules should have compositions that are very iron poor, presumably more iron-poor than LL-chondrites. This raises the ques- tion where are these? One possible answer is that chon- drules produced in these very early phases (< 1 million years after CAI formation) were either accreted into the sun or reincorporated into other bodies, where they were again molten. But if this process of elimination is so ef- ficient, why did some CAIs survive and get incorporated into the (later) chodrites? This shows that while the im- pact splashing model is appealing, there is more work to be done to answer such questions. 5.4. Compound chondrules In our homologously expanding cloud model the melt droplets move away from their nearest neighbor at ex- tremely low velocity (millimeters per minute). Small in- homogeneities (which are of course unavoidable in such a messy process as an impact) would lead to some droplets actually colliding. When they do so during the initial constant temperature phase, they presumably coalesce and form a larger melt droplet. However, when they col- lide during the cooling phase, they may have become cool enough to be plastic but not liquid anymore. They will then produce a compound chondrule. It would be useful if we could quantify the percentage of chondrules that become compound in this way. How- ever, at present we see no way how we can estimate the random relative velocities between adjacent chondrules in the expanding clouds. Many compound chondrules, however, clearly show one chondrule having remained rigid, while the other having been indented. So clearly they were of different temper- ature. This is difficult to achieve in our simple spherical expanding cloud model. For this it will be necessary to have other debris (e.g. another cloud-fragment emerging from the impact site) to interdisperse with our cloud of droplet, so as to have chondrules from different thermal environments to mix. And this must happen on a time scale shorter than the time scale of thermal equilibration between the chondrules from these different origins. To make this more quantitative would be, however, a chal- lenge. It would likely require us to contemplate complex splash geometries. Metzler (2012) found that some chondrites in fact con- tain entire large scale clusters of compound chondrules. 10 Dullemond, Stammler, Johansen A possible explanation could be that some of the im- pact debris may have fallen back onto the remainder of the biggest of the two colliding planetesimals. If this happens quickly enough, so that this re-accretion occurs before the chondrules have cooled below the temperature at which the chondrules are still sticky, then it is reason- able to assume that such clusters may form as chondrules simply fall onto each other. 5.5. What about relict grains and rims? Another problem with our simple spherically symmet- ric expanding plume model is that it does not account for any possible "pollution" of the melt droplets with other (non-molten) debris. Some chondrules contain relict grains (Jones 1996), which must have been inserted while the chondrule was still liquid, but the chondrule must have cooled very soon thereafer. As in the case with the compound chondrules (Section 5.4) this would require debris from a different part of the impact debris cloud to interdisperse with our cloud of droplets. This foreign de- bris must have originated from non-molten parts of the original planetesimals, e.g. the regolith or pristine dust covering the crust of the original planetesimals. Only those grains from that non-molten debris that entered late enough can enter the chondrules shortly before they cool below the solidus, and thus survive. Like with the compound chondrules, however, getting quantitative es- timates of this process is challenging. 5.6. Producing chondritic parent bodies The impact model may provide a plausible scenario for the formation of chondrules, but how do these chondrules (and matrix) accrete to form another planetesimal (the parent body of the later chondritic meteorite)? A nice aspect of the low-velocity collision of pre-molten plan- etesimals model is that much of the debris may, in fact, reaccrete onto one of the surviving original planetesimals (or what remains of it). The idea is that due to the high gas density in the protoplanetary disk the stochas- tic planetesimal motions are of low velocity so that when they collide, they collide essentially at velocities that are only just above their mutual escape speed. Since much of the energy is dissipated in the inelastic impact, most of the debris will not have enough kinetic energy to es- cape the system and will (after perhaps a flyby or two) eventually accrete onto a single body or a binary body. However, there is evidence that inside a single chondrite there exist chondrules of different ages (Villeneuve et al. 2009). And some chondrules experienced multiple heating events. Somehow chondrules produced in multi- ple chondrule-forming impacts must have mixed together and form a single chunk of rock that later is observed as a chondritic meteorite. The immediate re-accretion sce- nario may make this somewhat difficult because once the collision releases most of the magma, and much of this is reaccreted as chondrules (for the molten part) and ma- trix (for the non-molten part) the remaining body will no longer be a sphere of magma. If, however, the collisions would be high-speed enough that much of the debris escapes, the newly formed chon- drules will disperse into the protoplanetary disk. There they mix with the pool of other chondrules of various ages, and then later accrete into a chondrite parent body. An important question in this scenario is how aerody- namic drag and the resulting radial drift affects this: will the chondrules stay long enough in the asteroid-belt- region before they dift away or not? Jacquet et al. (2012) study this question in detail. Perhaps more likely is a combination of the two. In particular the cluster chondrites of Metzler (2012, see Section 5.4) can be easily understood in terms of the reaccretion scenario while the multiple ages seems more to point toward a dispersion and later accretion scenario. 5.7. Volatile elements One of the main characteristics of most chondrules is their "normal" abundance of volatile elements such as Na, K etc. If a chondrule would stay at a high tempera- ture (e.g. 2000K) for more than a few minutes, these el- ements would have evaporated out of the chondrule and leave a volatile-depleted chondrule behind. We believe that the impact splash scenario may naturally resolve this issue, because during the hot phases of the expand- ing impact splash the density of the cloud is so high that the vapor quickly reaches the saturation pressure and evaporation stops. However, this idea must be verified by actual modeling: Will the evaporation indeed satu- rate? Will the temperature drop below the evaporation temperature before the density of the expanding cloud becomes too low to saturate the pressure? Answering these questions with explicit calculations is the topic of paper II in this series. 5.8. What kind of chondrules qualify for this scenario? So far we have not made any distinction between differ- ent kinds of chondrules or chondrites. Given that chon- drule properties vary considerably between chondritic classes, the natural question arises: do all chondrules form through the same mechanism or are different chon- drule formation scenarios responsible for different kinds of chondrules? And for which kind, if at all, could the impact splash scenario of this paper be applicable? A special class of chondrites and chondrules is the CB and CH class. It has been recognized for some time that these chondrules may indeed have originated from collid- ing planetesimals or planets (see Desch et al. 2012 for a discussion). The evidence for this includes (1) the cryp- tocrystalline structure of CB chondrules, (2) their strong depletion of volatile elements, (3) zoned metal spherules that appear to have been condensed out of the gas phase in a highly energetic single-staged process (Krot et al. 2005) and (4) their young age, 5 Myr after CAIs. One can interpret these properties as meaning (1) that these chondrules must have cooled more rapidly (seconds to minutes) than "normal" chondrules (hours) to explain their textures, (2) that the vapor-melt plume of the im- pact must have rapidly become of low enough density to allow volatile elements to get out of the chondrules and not recondense, (3) the impact must have been of very high velocity (several km/s) to produce metal vapor, (4) which appears consistent with the young age which means low gas densities in the protoplanetary disk and thus allows for high-speed impacts. Arguments (1) and (2) seem to require a rapid reduction of the density and optical depth, which can be achieved by a high impact velocity (also required for (3) and consistent with (4)) Forming chondrules in impact splashes 11 and a low mass of the impactor. Krot et al. (2005) ar- gue, instead, for a very high mass collision to explain CB chondrules. Perhaps the adiabatic cooling of the impact- shock-compressed rock is enough to explain the fast cool- ing, or only a fraction of the mass from the giant impact actually gets converted into a plume of chondrules. In constrast to the case for the CB/CH chondrules our model may explain the properties of "normal" chon- drules: (1) barred or porphyritic textures indicating slower cooling (hours), (2) little depletion of volatiles (see Section 5.7), (3) possibly lower temperatures and (4) earlier times after CAIs. The slower cooling rates come naturally out of our model because the optical depth of the cloud keeps the chondrules warm for a while. The little depletion of volatiles may be achieved by the high densities (though a quantitative analysis has to wait until paper II in this series). And the lower temperatures are natural for lower impact velocities which are expected in the earlier phases of the protoplanetary disk. The melt is the molten interior of the 26Al-heated planetesimals, which also is expected during the earlier phases. We might also be able to apply our radiatively cool- ing expanding cloud model to the high impact velocity scenario for CB/CH chondrules, but it would require us to include more physics, in particular a proper equation of state for the shock-heated rock and the subsequent adiabatic expansion and adiabatic cooling. Among the "normal" chondrules there is also the issue of radial and barred textures versus porphyritic textures. The latter are the majority (about 85% of chondrules, Gooding & Keil 1981). Radial textures require that no nucleation sites were present in the melt, so that crys- tallization in the supercooled melt droplet starts from a single site and progresses radially outward from that site. In the splash scenario this means that the melt in the original pre-molten planetesimals was well above the liquidus, so that any previously existing crystals were molten. Conversely, for porphyritic chondrules numer- ous nucleation sites must have been present in the melt, and thus must have presumably remained present in the magma of the colliding planetesimals, meaning that that magma was likely slightly below the liquidus tempera- ture. The impact may have heated the melt through shock-heating above the liquidus, but if the cooling sets in quickly enough some of these nucleation sites may sur- vive. The question of porphyritic textures and the origin of the nucleation sites in the splash scenario remains, however, a tricky one. Sanders & Scott (2012) propose instead that these nucleation sites may have entered into the melt droplets as pollution from e.g. the surface re- golith that may have been mixed in with the chondrules. The problem with this (as with the relict grain issue of Section 5.5) is that the timing of mixing must be ideal: too early and the temperature is still above the liquidus and these regolith particles will also melt; too late and the chondrules have already formed radial or barred tex- tures. Finally there is the issue of Fe-content of chondrules: the H, L and LL classes. If the pre-molten planetesimals have experienced strong differentiation, it is possible that the mantles of these planetesimals are poor in iron and nickel. If the masses of the planetesimals are low enough and convection might be present, then perhaps all too strong loss of Fe in the mantle can be avoided. The issue of H, L and LL chondrites in the context of the splash scenario is further discussed in Sanders & Scott (2012). 6. SUMMARY AND CONCLUSION This paper studies the radiative cooling of an ex- panding cloud of lava droplets originating from the low- velocity collision between two partly or fully pre-molten planetesimals. The model is also valid for high-speed col- lisions between cold planetesimals, in which the melt is produced by the collision itself. The main result of this paper is a self-consistent cool- ing profile for chondrules based on the initial conditions of the expanding plume of melt droplets (chondrules). These cooling profiles are characterized by a time scale tcool given by Eq. (16) and are well approximated by a constant temperature for t < tcool and a powerlaw dropoff (Eq. 18) for t > tcool. The cooling rate is given by Eq. (17). The cooling time scale tcool depends on the total mass of the plume and on its expansion veloc- ity. Both parameters are related to the conditions of the collision between the two planetesimals. A large total mass of the expanding cloud of chondrules means that the colliding planetesimals must have been at least of that mass, but presumably considerably more massive since not all of the matter is likely to end up in a single expanding plume. The expansion velocity of the droplet cloud is related to the collision velocity of the two planetesimals: it is unlikely that it is higher than the impact velocity. Note that the expansion velocity is different from (and presumably much less than) the velocity by which the cloud moves away from the impact site. To relate both parameters (Mcloud and vexp) to paramters of the colliding planetesimals it is necessary to perform 3-D hydrodynamic simulations of the impact, such as those performed by Asphaug et al. (2011). The results of our model, when compared to known cooling time scales of chondrules, put constraints on Mcloud and vexp, and thus on the conditions under which planetesimal collisions can produce chondrules. Acknowledgements: We would like to thank An- dreas Pack, Knut Metzler, Alessandro Morbidelli, David Lundberg, Addi Bischoff, Neal Turner, Stephan Henke and Mark Swain for useful discussions and feedback. This paper benefitted a lot from detailed comments and suggestions by an anonymous referee, for which we are very grateful. S. Stammler is funded by the Deutsche Forschungsgemeinschaft (DFG) grant Du 414/12-1. Alexander, C.M.O. 1996, Chondrules and the protoplanetary Alexander, C.M.O., Grossman, J.N., Ebel, D.S., Ciesla, F.J., disk, 233 2008, Science, 320, 1617 Alexander, C.M.O., Ebel, D.S., 2012, Meteoritics and Planetary Asphaug, E., Jutzi, M., Movshovitz, N., 2011, Earth and Science, 47, 1157 Planetary Science Letters, 308, 369 REFERENCES 12 Dullemond, Stammler, Johansen Boss, A. 1996 in Proceedings "Chondrules and the protoplanetary Krot, A. N., Amelin, Y., Cassen, P. & Meibom, A. 2005, Nature, disk", pp 257. 436, 989992 Bottke, W.F. 1994, Icarus, 107, 255 Ciesla, F.J., Hood, L.L., 2002, Icarus, 158, 281 Connelly, J.N., Bizzarro, M., Krot, A.N., Nordlund, A., Wielandt, D., Ivanova, M.A., 2012, Science, 338, 651 Cuzzi, J., Alexander, C.M.O., 2006, Nature, 441, 483 Davis, A.M., Alexander, C.M. O'D., Ciesla, F.J., Gounelle, M., Krot, A.N., Petaev, M.I., Stephan, T., 2014, in Protostars and Planets VI, eds. Beuther, Klessen, Dullemond, Henning. Lambrechts, M., & Johansen, A., 2012, A&A, 544, A32 Libourel, G., Chaussidon, M., Krot, A.N., 2008, in 39th Lunar and Planetary Science Conference, vol. 39, pp 2017 Metzler, K., 2012, Meteoritics and Planetary Science, 47, 2193 Metzler, K., Bischoff, A. & Stoeffler, D. 1992, Geochimica et Cosmochimica Acta, 56, 2873 Mihalas, D. & Mihalas, B, "Foundations of Radiation Hydrodynamics" Courier Dover Publications, 1999 Desch, S.J., Connolly, H.C., 2002, Meteoritics and Planetary Miura, H. & Yamamoto, T. 2014, The Astronomical Journal, 147, Science, 37, 183 54 Desch, S.J., Morris, M.A., Connolly, H.C. & Boss, A.P. 2012, Morris, M. A. & Desch, S. J. 2010, The Astrophysical Journal, Meteoritics and Planetary Science, 47, 11391156 Gibbard, S. G., Levy, E. H. & Morfill, G. E. 1997, Icarus, 130, 517 Gooding, J.L, Keil, K., 1981, Meteoritics, 16, 17 Hevey, P.J., Sanders, I.S., 2006, Meteoritics and Planetary Science, 41, 95 Hewins, R.H., 1983, In: Chondrules and their origins, Houston, pp 122 722, 1474 Morris, M. A., Boley, A. C., Desch, S. J. & Athanassiadou, T. 2012, ApJ, 752, 27 Neumann, W., Breuer, D. & Spohn, T., 2012, A&A, 543, 141 Ormel, C. W., & Klahr, H. H., 2010, A&A, 520, A43 Sanders, I.S., Scott, E.R.D., 2012, Meteoritics and Planetary Science, 47, 2170 Hewins, R.H. & Connolly, H. C., Jr. 1996, in Chondrules and the Sanders, I.S., Taylor, G.J., 2005, Chondrites and the Protoplanetary Disk (Cambridge: Cambridge Univ. Press), Hood, L.L., Horanyi, M., 1991, Icarus, 93, 259 Hubbard, A., McNally, C. P. & Mac Low, M.-M. 2012, ApJ, 761, 58 Jacquet, E., Gounelle, M. & Fromang, S. 2012, Icarus, 220, 162 Jones, R.H., Grossman, J.N., Rubin, A.E. 2005, in Chondrites and the Protoplanetary Disk, ASP Conference Series, Vol 341, Eds: Krot, Scott & Reipurth Kieffer, S., 1975, Science, 189, 333 Protoplanetary Disk, 341, 915 Sears, D., 2004 "The Origin of Chondrules and Chondrites", Cambridge University Press Stammler, S.M. & Dullemond, C.P. 2014, Icarus in press. Taylor, G. J., Scott, E. R. D. & Keil, K. 1983, IN: Chondrules and their origins (A85-26528 11-91). Houston, 262 Villeneuve, J., Chaussidon, M., Libourel, G., 2009, Science, 325, 985 Urey, H.C., Craig, H. 1953, Geochimica et Cosmochimica Acta, 4, 36 Zook, H.A., 1980, Meteoritics, 15, 390 TIME-DEPENDENT RADIATIVE TRANSFER: DISCRETIZATION AND IMPLICIT INTEGRATION Solving the radiative transfer moment equations coupled to the time-dependent heating/cooling equation requires an implicit integration scheme. The two coupled equations to solve are: APPENDIX F1(J, T ) := 1 α 1 r2 d dr (cid:26) r2 α (cid:20) d(f J) dr + 3f − 1 r and The inner boundary condition is: F2(J, T ) := ρcv dT dt − 4πα(cid:16)J − σ π σ π T 4 J(cid:21)(cid:27) − J + T 4(cid:17) = 0 where rin is the smallest radius of our grid, which we take rin ≪ Rcloud. This condition translates into F1(J, T, r = rin) := H(r = rin) = 0 (cid:20) d(f J) dr + 3f − 1 r J(cid:21)r=rin = 0 The outer boundary condition is F1(J, T, r = rout) := H(r = rout) − hJ = 0 with h = Hfte(rout)/Jfte(rout), which translates into F1(J, T, r = rout) :=(cid:20) d(f J) dr + 3f − 1 r J + hαJ(cid:21)r=rout = 0 (A1) (A2) (A3) (A4) (A5) (A6) The only time-derivative in the equations is the one on the temperature. The radiation field is assumed to immediately adapt. We put J(r, t) and T (r, t) on a spatial grid {r1, · · · , rN }. So for time step n we have the values J n 1 , · · · , J n N N . The above equations are also to be evaluated at these grid points: F1,1 = 0, · · · , F1,N = 0 and 1 , · · · , T n and T n F2,1 = 0, · · · , F2,N = 0. Let us define and F (Q) = (F1,1(Q), F2,1(Q), F1,2(Q), F2,2(Q), · · · , F1,N (Q), F2,N (Q))T Qn = (J n 1 , T n 1 , J n 2 , T n 2 , · · · , J n N , T n N )T (A7) (A8) Forming chondrules in impact splashes 13 The objective of the time-integration of these equations is to find the values of Qn+1: the values of Q at the next time step. For numerical stability we express all instances of Q in the equations F in their future form: Qn+1, except when a time-derivative of Q is used, which is written as ∂Q/∂t = (Qn+1 − Qn)/∆t and where the present value of Qn is required. We thus get as our set of equations: F (Qn+1) = 0 (A9) which we need to solve for Qn+1. If F were linear in Qn+1 then one could write the vector F as a matrix multiplication with the vector Qn+1 plus a contant vector R: Then the solution to F = 0 would be just a matrix inversion: Qn+1 = −M invR F = M Qn+1 + R (A10) (A11) However, F is non-linear in Qn+1 because the Planck function B(T ) = (σ/π)(T n+1)4. And so we have a non-linear function F (Qn+1). So let us start with an initial guess Qn+1 (0) . We typically then have Ideally we want to find a Qn+1 (1) F (Qn+1 (1) ) using first order Taylor expansion: F (Qn+1 (0) ) 6= 0 (A12) for which F (Qn+1 (1) ) = 0, but we can only use Newton's method by approximating or more in general for the k-th iteration: F (Qn+1 (1) ) ≃ F (Qn+1 (0) ) + F (Qn+1 (k+1)) ≃ F (Qn+1 (k) ) + ∂F ∂Q ∂F ∂Q · (Qn+1 (1) − Qn+1 (0) ) = 0 · (Qn+1 (k+1) − Qn+1 (k) ) = 0 If we define we can solve for ∆Qn+1 (k) : and then we obtain the new: ∆Qn+1 (k) = Qn+1 (k+1) − Qn+1 (k) ∆Qn+1 (k) = −(cid:18) ∂F ∂Q(cid:19)inv F (Qn+1 (k) ) Qn+1 (k+1) = ∆Qn+1 (k) + Qn+1 (k) Now let us write out the discrete equations F1(J, T ) and F2(J, T ) explicitly: F1,i(J, T n+1) ≡ 1 αir2 i ∆ri(cid:18) r2 i+1/2 αi+1/2 fi+1Ji+1 − fiJi ∆ri+1/2 fiJi − fi−1Ji−1 r2 i−1/2 αi−1/2 (3fi+1/2 − 1)(Ji+1 + Ji)ri+1/2 ∆ri−1/2 2αi+1/2 (3fi−1/2 − 1)(Ji−1 + Ji)ri−1/2 2αi−1/2 (T n+1 i )4 = 0 (cid:19) − + − σ π − Ji + For the gas equation we get: F2,i(J, T n+1) ≡ ρcv T n+1 i − T n i ∆t − 4παi(cid:16)Ji − σ π (T n+1 i )4(cid:17) − q = 0 Boundary conditions only have to be applied to the first equation. The inner boundary condition: F1,1(J, T n+1) = f2J2 − f1J1 α3/2∆r3/2 + 1 2 (J1 + J2) 3f3/2 − 1 α3/2r3/2 = 0 (A13) (A14) (A15) (A16) (A17) (A18) (A19) (A20) 14 Dullemond, Stammler, Johansen The outer boundary condition: F1,N (J, T n+1) = fN JN − fN −1JN −1 αN −1/2∆rN −1/2 + 1 2 (JN + JN −1)(cid:18) 3fN −1/2 − 1 αN −1/2rN −1/2 + h(cid:19) = 0 The energy equation can be rescaled to: F2,i(J, T n+1) ≡ T n+1 i − T n i − The matrix coefficients then become: 4παi∆t ρcv σ π (cid:16)Ji − (T n+1 i q∆t ρcv )4(cid:17) − = 0 i−1/2fi αi−1/2∆ri−1/2 1 αir2 i ∆ri ( r2 i ∆ri ( r2 i ∆ri ( r2 1 αir2 1 αir2 i+1/2fi+1 αi+1/2∆ri+1/2 i−1/2fi−1 αi−1/2∆ri−1/2 + (3fi+1/2 − 1)ri+1/2 r2 i+1/2fi αi+1/2∆ri+1/2) − 1 ) ) 2αi−1/2 2αi+1/2 (3fi−1/2 − 1)ri−1/2 + + (cid:18) ∂F1,i ∂Ji (cid:19) = − (cid:18) ∂F1,i ∂Ji+1(cid:19) = ∂Ji−1(cid:19) = (cid:18) ∂F1,i (cid:18) ∂F1,i (cid:19) = (cid:18) ∂F2,i ∂Ji (cid:19) = − (cid:19) = 1 + (cid:18) ∂F2,i ∂T n+1 ∂T n+1 4σ π i i (T n+1 )3 i 4παi ρicv ∆t 4παi ρicv 4σ π (T n+1 i )3∆t The matrix coefficients for the boundary conditions become: f2 ∆r3/2(cid:27) ∆r3/2(cid:27) f1 (cid:18) ∂F1,1 ∂J2 (cid:19) = (cid:18) ∂F1,1 ∂J1 (cid:19) = (cid:18) ∂FN,1 ∂JN (cid:19) = ∂JN −1(cid:19) = (cid:18) ∂FN,1 1 1 + − 2r3/2 2r3/2 α3/2(cid:26) 3f3/2 − 1 α3/2(cid:26) 3f3/2 − 1 αN −1/2 (cid:18) 3fN −1/2 − 1 αN −1/2 (cid:18) 3fN −1/2 − 1 2rN −1/2 2rN −1/2 1 1 + − fN ∆rN −1/2(cid:19) + ∆rN −1/2(cid:19) + fN −1 h 2 h 2 (A21) (A22) (A23) (A24) (A25) (A26) (A27) (A28) (A29) (A30) (A31) (A32)
1504.01678
1
1504
2015-04-07T17:26:01
A $\sim$32-70 K formation temperature range for the ice grains agglomerated by comet 67P/Churyumov-Gerasimenko
[ "astro-ph.EP" ]
Grand Canonical Monte Carlo simulations are used to reproduce the N$_2$/CO ratio ranging between 1.7 $\times$ 10$^{-3}$ and 1.6 $\times$ 10$^{-2}$ observed {\it in situ} in the Jupiter family comet 67P/Churyumov-Gerasimenko by the ROSINA mass spectrometer aboard the Rosetta spacecraft, assuming that this body has been agglomerated from clathrates in the protosolar nebula. Simulations are done using an elaborated interatomic potentials for investigating the temperature dependence of the trapping within a multiple guest clathrate formed from a gas mixture of CO and N$_2$ in proportions corresponding to those expected for the protosolar nebula. By assuming that 67P/Churyumov-Gerasimenko agglomerated from clathrates, our calculations suggest the cometary grains must have been formed at temperatures ranging between $\sim$31.8 and 69.9 K in the protosolar nebula to match the N$_2$/CO ratio measured by the ROSINA mass spectrometer. The presence of clathrates in Jupiter family comets could then explain the potential N$_2$ depletion (factor up to $\sim$87 compared to the protosolar value) measured in 67P/Churyumov-Gerasimenko.
astro-ph.EP
astro-ph
A ∼32-70 K formation temperature range for the ice grains agglomerated by comet 67P/Churyumov-Gerasimenko S. Lectez1, J.-M. Simon1, O. Mousis2, S. Picaud3, K. Altwegg4, M. Rubin4, J.M. Salazar1 Received ; accepted 1Laboratoire Interdisciplinaire Carnot de Bourgogne, UMR 6303 CNRS-Universit´e de Bourgogne, Dijon, France [email protected] 2Aix Marseille Universit´e, CNRS, LAM (Laboratoire d'Astrophysique de Marseille) UMR 7326, 13388, Marseille, France 3Institut UTINAM, UMR 6213 CNRS-Universit´e de Franche Comt´e, Besan¸con, France 4Physikalisches Institut, University of Bern, Sidlerstrasse 5, CH-3012 Bern, Switzerland -- 2 -- ABSTRACT Grand Canonical Monte Carlo simulations are used to reproduce the N2/CO ratio ranging between 1.7 × 10−3 and 1.6 × 10−2 observed in situ in the Jupiter family comet 67P/Churyumov-Gerasimenko by the ROSINA mass spectrometer aboard the Rosetta spacecraft, assuming that this body has been agglomerated from clathrates in the protosolar nebula. Simulations are done using an elaborated interatomic potentials for investigating the temperature dependence of the trapping within a multiple guest clathrate formed from a gas mixture of CO and N2 in proportions corresponding to those expected for the protosolar nebula. By assuming that 67P/Churyumov-Gerasimenko agglomerated from clathrates, our calculations suggest the cometary grains must have been formed at temperatures ranging between ∼31.8 and 69.9 K in the protosolar nebula to match the N2/CO ratio measured by the ROSINA mass spectrometer. The presence of clathrates in Jupiter family comets could then explain the potential N2 depletion (factor up to ∼87 compared to the protosolar value) measured in 67P/Churyumov-Gerasimenko. Subject headings: astrobiology -- comets: general -- comets: individual (67P/Churyumov-Gerasimenko) -- solid state: volatile -- methods: numerical -- 3 -- 1. Introduction The thermodynamic conditions prevailing in many bodies of the solar system suggest that clathrates could exist in the Martian permafrost (Thomas et al. 2009; Swindle et al. 2009; Mousis et al. 2013), on Titan (Choukroun & Sotin 2012; Mousis et al. 2014), as well as in the interiors of other icy satellites (Kieffer et al. 2006; Hand et al. 2006). It has also been suggested that the activity observed in some cometary nuclei results from the dissociation of these crystalline structures (Marconi & Mendis 1983; Smoluchowski 1988; Marboeuf et al. 2010, 2011, 2012; Mousis et al. 2012). Several indirect evidences suggest that clathrates probably participated in the formation of planetesimals and the building blocks of giant planets in the outer solar system (Lunine & Stevenson 1985; Mousis et al. 2010, 2014). In the absence of experimental data existing at low-temperature (20 -- 200 K) and low-pressure conditions (10−13 -- 10−3 bar), which are typical of those encountered in planetary environments (Lunine & Stevenson 1985; Sloan & Koh 2008), clathrates are characterized via theoretical modeling. The approach usually employed in planetary science is based on the statistical mechanics model initially proposed by van der Waals & Platteeuw (1959), which makes use of simplified intermolecular potentials calibrated on equilibrium measurements performed at relatively high temperatures. In consequence, the use of these potentials at thermodynamic conditions relevant to those of the protosolar nebula (hereafter PSN) for predicting the composition of clathrates deserves to be confronted with more sophisticated approaches. In this present work, we aim at reproducing the N2/CO ratio ranging between 1.7 × 10−3 and 1.6 × 10−2 observed in situ in the Jupiter family comet 67P/Churyumov- Gerasimenko (hereafter 67P) by the ROSINA mass spectrometer aboard the Rosetta spacecraft (Balsiger et al. 2007), which is found to be depleted by a factor up to ∼87 compared to the value of 0.148 hypothesized for the protosolar nebula (see below). By -- 4 -- assuming that 67P has been agglomerated from clathrates, it is possible to derive the temperature range of formation of these crystalline structures in the PSN by mean of Grand Canonical Monte Carlo (GCMC) simulations based on elaborated interatomic potentials. These allowed us to investigate the temperature dependence of the trapping within a multiple guest (hereafter MG) clathrate formed from a gaseous mixture of CO and N2 in proportions corresponding to those expected for the protosolar nebula. 2. Computational details and simulation procedure We assume multiple guest (MG) clathrate formation from a gaseous mixture composed of N2 and CO in proportions specified from a plausible protosolar gas phase composition. For this, we have assumed that both all C and N are in form of CO and N2 in the initial disk's gas phase. This assumption is in agreement with thermochemical models of the protosolar nebula (Lewis & Prinn 1980; Prinn & Fegley 1989; Mousis et al. 2002). The use of C and N protosolar elemental abundances from the compilation of Lodders et al. (2009) allowed us to derive a gaseous mixture with mole fractions of 0.871 for CO and 0.129 for N2, giving a N2/CO ratio of 0.148 in the PSN. The MG clathrate composition has been computed along its equilibrium pressure curve Peq,M G given by (Thomas et al. 2007; Mousis & Schmitt 2008): (cid:35)−1 (cid:34)(cid:88) i yi Peq,i Peq,M G = , (1) where yi is the mole fraction of species i in the gas phase (here CO or N2) and Peq,i the equilibrium pressure of single guest clathrate formed from component i only. The equilibrium pressures Peq,i (in bar) are determined by using an equation based on an Arrhenius law (Miller 1961): -- 5 -- ln Peq,i = Ai T + Bi, (2) where Ai and Bi are constant parameters depending on the nature of the species trapped in the clathrate and T is the temperature (K). Ai and Bi have been determined by fitting the available theoretical and laboratory data (Mousis et al. 2008) and their values are given in Table 1. The calculated values of Peq,i and partial pressures of CO (PCO) and N2 (PN2) are represented as a function of the inverse temperature in Fig. 1. Here, the equilibrium pressure of MG clathrate is in the ∼5.2 ×10−10 -- 2.9 ×10−3 bar range when T is varied between ∼52 and 100 K, respectively. We have calculated the composition of N2 -- CO clathrates via Monte-carlo (MC) simulations in the Grand Canonical ensemble (GCMC) (Frenkel & Smit 2002) for temperatures ranging from 52 to 100 K (the choice of this temperature range is explained in Sec. 3, but it is worth noting that, below 50 K, the equilibration time is in fact too long for the GCMC simulation), with an increment of 4 K between each computation. All our calculations have been performed in the case of Structure I (SI) clathrates (see structural details in Sloan & Koh (2008)). This choose has been motivated by the fact that CO is the dominating species in the gaseous mixture and is known to form Structure I single guest clathrate (Mohammadi 2005). In our system, the considered crystal size consists in 125 cubic unit cells (5×5×5), corresponding to 5,750 water molecules. The dimension of one parameter of the cubic simulation box is set equal to 60.15 A during all our simulations. Periodic boundary conditions are applied to mimic infinite crystal. The water molecules are modeled using the well-known TIP4P/2005 model (Abascal & Vega 2005), allowing them to translate and rotate during the simulation. Models for N2 and CO molecules are taken from Potoff & Siepmann (2001) and Piper et al. (1984), respectively. One hundred million MC steps were -- 6 -- performed including insertion, deletion, translation and rotation of the molecules. Only the last 50 million steps were used to compute the data. The first 50 million steps have been discarded from the analysis and were only used to equilibrate the system. From the knowledge of the partial pressure of each species (see Fig. 1), GCMC simulations have been performed to compute the composition of the MG clathrate. However, the number of MC steps needed to equilibrate the system strongly increases when temperature decreases. Simulation tests showed that below a temperature of 52 K, this number becomes even larger than the 50 million steps we simulated. At these low temperatures, statistically relevant results on the mole fractions of encaged molecules in our MC calculations thus become too time consuming. In consequence, below 52 K, these quantities have been evaluated via a thermodynamic extrapolation, described below, and fitted to the results computed at higher temperatures. After the equilibration period, chemical equilibrium takes place between gas and clathrate. It is then possible to write an equilibrium constant Ki for each species i in the form: Ki = ai Pi/P 0 , (3) where ai is the activity of species i in clathrate defined from the mole fraction of species i in clathrate (xi) and its activity coefficient (γi ) as ai = γixi. The evolution of Ki with temperature should obey the Van't Hoff relation: dlnKi dT = ∆Ei RT 2 , (4) where ∆Ei is the entrapping energy of CO or N2, and R the ideal gas constant. Under the simulated conditions, the occupancy of the clathrate cages is still maximum. -- 7 -- We thus, obtained an amount of 8 molecules of CO and N2 per unit cell, implying that the total number of molecules trapped in clathrate is constant, independently of the temperature. By Neglecting the non ideal terms of the activity coefficient at first approximation, this allowed us to get γi = 1 and to rewrite Eq. 4 as follows: dln(Ni/Pi) d(1/T ) = −∆Ei R , (5) with Ni the number of encaged molecules of type i. Our simulations allowed retrieving the values of NCO and NN2 at temperatures higher than 52 K and provided us with a direct access to the different values of the quantity ln(Ni/Pi). Figure 2 represents these quantities plotted for CO and N2 as a function of 1/T . They show a linear behavior with a correlation coefficient higher than 0.999 for each species, in agreement with Eq. 5. The computed data have been fitted via the use of linear equations, allowing to find ln(NCO/PCO) = 1685.145/T - 15.617 and ln(NN2/PN2) = 1554.469/T - 15.972. The entrapping energies comes directly from the fit, ∆ECO = −1685.145×R and ∆EN2 = −1554.469×R. These two linear equations have been used to estimate the NN2/NCO ratio at temperatures lower than 52 K. Note that, for simplification, we will use below the abbreviation N2/CO for this ratio. 3. Results Figure 3 shows the evolution of N2/CO ratio as a function of temperature in the 20 -- 100 K range. This ratio monotonically increases with the growing temperature. The figure exhibits a linear regime in the 50 -- 100 K range whereas at lower temperatures, the curve trend is less steep and the ratio converges smoothly towards zero. The N2/CO ratio found is equal to ∼1.5 × 10−4 at 20 K. This value is 50 times smaller than at 50 K. In the temperature range considered here, the calculated N2/CO ratio is significantly lower than -- 8 -- the ratio of ∼0.15 in the coexisting gas phase. This indicates that the clathrate formation favors the CO entrapping at the expense of N2. This behavior is related to the difference in entrapping energy between CO and N2. Indeed, ∆ECO being lower, the entrapping of CO is selectively favored when the temperature decreases. Figure 4 is a zoom of a portion of Fig. 3 given for easy reading of the correspondence between the N2/CO ratio measured in 67P by the ROSINA instrument and the formation temperature of the ice grains from which the comet agglomerated. Taking into account the strong variation of the N2/CO measurement between 0.17 to 1.6% depending on the position of the Rosetta spacecraft above the surface of the comet nucleus (Rubin et al. 2015), we find that the ice grains at the origin of 67P formed at temperatures ranging between ∼31.8 and 69.9 K in the protosolar nebula, with corresponding equilibrium pressures ranging between 6.0 × 10−19 and 2.1 × 10−6 bar. For the sake of information, the mean N2/CO ratio of 0.57% corresponding to the averaging of the 138 spectra obtained by Rubin et al. (2015) is represented on Fig. 4. The corresponding formation temperature of the ice grains is of ∼45 K in the PSN, with an equilibrium pressure of 3.3 × 10−12 bar. 4. Discussion and Conclusions The composition of a MG clathrate formed from a gaseous mixture of N2 and CO in proportions corresponding to those expected for the protosolar nebula (87.1 % for CO and 12.9% for N2) has been investigated in the 20 -- 100 K temperature range. Above 50 K, the clathrate composition has been computed via Grand Canonical Monte-Carlo simulations for pressures ranging from ∼5.2 ×10−10 to 2.9 ×10−3 bar. Below 50 K, the clathrate composition has been extrapolated via the use of a Van't Hoff relation. The results show that, at thermodynamic conditions relevant to those of the protosolar nebula, CO has a much higher propensity than N2 to be trapped in clathrates. Assuming that -- 9 -- 67P agglomerated from clathrates, our calculations suggest that the cometary grains must have formed at temperatures ranging between ∼31.8 and 69.9 K in the protosolar nebula to match the N2/CO ratio measured by the ROSINA mass spectrometer (Rubin et al. 2015). Whilst narrower, the range of formation temperatures inferred from our model for the grains of 67P is consistent with the one (∼22 -- 80 K) found from the reading of Fig. 2 of Mousis et al. (2012) who performed calculations of planetesimals compositions based on the classical statistical mechanics model of van der Waals & Platteeuw (1959). In the absence of experiments at this temperature range, the fact that these two different approaches lead to similar conclusions, namely that clathrates can explain the N2/CO ratio observed in 67P, suggest that this scenario is plausible. The presence of clathrates in Jupiter family comets could then explain the apparent N2 depletion (factor up to ∼87 compared to the protosolar value) measured in 67P. Financial support from the BQR Bourgogne Franche -- Comt´e is gratefully acknowledged. O.M. acknowledges support from CNES. This work has been partly carried out thanks to the support of the A*MIDEX project (no ANR-11-IDEX-0001-02) funded by the "Investissements d'Avenir" French Government program, managed by the French National Research Agency (ANR). -- 10 -- REFERENCES Abascal, J. L. F., & Vega, C. 2005, J. Chem. Phys., 123, 234505 Balsiger, H., Altwegg, K., Bochsler, P., et al. 2007, Space Sci. Rev., 128, 745 Choukroun, M., & Sotin, C. 2012, Geophys. Res. Lett., 39, L04201 Frenkel, D. & Smit, B., 2002, Understanding molecular simulation: from algorithms to applications, Elsevier Hand, D. P., Chyba, C. F., Carlson, R. W., & Cooper, J. F. 2006, Astrobiology, 6, 463 Kieffer, S. W., Lu, X., Bethke, C. M., et al. 2006, Science, 314, 1764 Lewis, J. S., & Prinn, R. G. 1980, ApJ, 238, 357 Lodders, K., Palme, H., & Gail, H.-P. 2009, Landolt Bornstein, 44 Lunine, J. I. & Stevenson, D. J. 1985, The Astrophysical J. Supplement Series, 58, 493 Marboeuf, U., Schmitt, B., Petit, J.-M., Mousis, O., & Fray, N. 2012, A&A, 542, AA82 Marboeuf, U., Mousis, O., Petit, J.-M., et al. 2011, A&A, 525, AA144 Marboeuf, U., Mousis, O., Petit, J.-M., & Schmitt, B. 2010, ApJ, 708, 812 Marconi, M. L., & Mendis, D. A. 1983, ApJ, 273, 381 Miller, S. L. 1961, Proceedings of the National Academy of Science, 47, 1798 Mohammadi, A.H., Anderson, R., & Tohidi, B. 2005, Am. In. Chem. Eng., 51, 2825 Mousis, O., Lunine, J. I., Fletcher, L. N., et al. 2014, ApJ, 796, LL28 Mousis, O., Chassefi`ere, E., Lasue, J., et al. 2013, Space Sci. Rev., 174, 213 -- 11 -- Mousis, O., Lunine, J. I., Picaud, S., & Cordier, D. 2010, Faraday Discussions, 147, 509 Mousis, O., & Schmitt, B. 2008, ApJ, 677, L67 Mousis, O., Alibert, Y., Hestroffer, D., et al. 2008, MNRAS, 383, 1269 Mousis, O., Guilbert-Lepoutre, A., Lunine, J. I., et al. 2012, ApJ, 757, 146 Mousis, O., Gautier, D., & Bockel´ee-Morvan, D. 2002, Icarus, 156, 162 Piper, J., Morrison, J. A., & Peters, C. 1984, Molecular Physics, 53, 1463 Potoff, J.J & Siepmann, I. 2001, AICHE J., 47, 1676 Prinn, R. G. P., & Fegley, B., Jr. 1989, Origin and Evolution of Planetary and Satellite Atmospheres, 78 Rubin, M., Altwegg, K., Balsiger, H., et al. 2015, Science, DOI:10.1126/science.aaa6100 Sloan, E. D. & Koh, C. A. 2008, Clathrate Hydrates of Natural Gases, 3rd ed.; CRC Press, Taylor & Francis Group, Boca Raton Smoluchowski, R. 1988, MNRAS, 235, 343 Swindle, T. D., Thomas, C., Mousis, O., Lunine, J. I., & Picaud, S. 2009, Icarus, 203, 66 Takeuchi, F., & Hiratsuka, M., Ohmura, R., Alavi, S., Amadeu, K. S. et al. 2013, J. Chem. Phys., 138, 124504 Thomas, C., Mousis, O., Picaud, S., & Ballenegger, V. 2009, Planet. Space Sci., 57, 42 Thomas, C., Mousis, O., Ballenegger, V., & Picaud, S. 2007, A&A, 474, L17 van der Waals, J. H., & Platteeuw, J. C., 1959, Advances in Chemical Physics, 2, 1 This manuscript was prepared with the AAS LATEX macros v5.2. -- 12 -- Table 1: Parameters of the equilibrium curves of the considered single guest clathrates Molecule type i Ai / (K) Bi CO N2 -1685.54 10.9946 -1677.62 11.1919 -- 13 -- Fig. 1. -- Calculated pressure and partial pressures of CO and N2 in the gas as a function of the inverse temperature based on the Arrhenius law and using parameters in Table 1. 0.010.0120.0140.0160.0180.021/T (K-1)1e-61e-40.011100Pressure (Pa)Peq,MGPN2PCO -- 14 -- Fig. 2. -- Evolution of the number of encaged molecules in the cases of CO (a) and N2 (b) as a function of the inverse temperature. A linear fit was performed on the data with correlation coefficients higher than 0.999 in both cases (see text). The figure shows the results of the GCMC simulation between 52K and 100K. 0.0120.0160.021/T (K)051015ln (Ni,clat./Pi,gas)0.0120.0160.021/T (K)051015ln(Ni,clat./Pi,gas)CON2(a)(b) -- 15 -- Fig. 3. -- N2/CO ratio in clathrate as a function of formation temperature. The results are derived from the GCMC simulations performed above 52K. Below this temperature, they are based on the entrapping energies derived from the GCMC simulations. 2030405060708090100T (K)00.0040.0080.0120.0160.020.0240.028N2 / CO -- 16 -- Fig. 4. -- Minimum and maximum N2/CO ratios measured in 67P and corresponding for- mation temperatures for the ice grains. The results are derived from the GCMC simulations performed above 52K. Below this temperature, they are based on the entrapping energies derived from the GCMC simulations. 283236404448525660646872T (K)00.0020.0040.0060.0080.010.0120.0140.016N2 / COT = 45.0 KT = 31.8 KT = 69.9 K
1701.03795
2
1701
2017-04-25T21:48:35
The Eccentric Kozai-Lidov mechanism for Outer Test Particle
[ "astro-ph.EP", "astro-ph.SR" ]
The secular approximation of the hierarchical three body systems has been proven to be very useful in addressing many astrophysical systems, from planets, stars to black holes. In such a system two objects are on a tight orbit, and the tertiary is on a much wider orbit. Here we study the dynamics of a system by taking the tertiary mass to zero and solve the hierarchical three body system up to the octupole level of approximation. We find a rich dynamics that the outer orbit undergoes due to gravitational perturbations from the inner binary. The nominal result of the precession of the nodes is mostly limited for the lowest order of approximation, however, when the octupole-level of approximation is introduced the system becomes chaotic, as expected, and the tertiary oscillates below and above 90deg, similarly to the non-test particle flip behavior (e.g., Naoz 2016). We provide the Hamiltonian of the system and investigate the dynamics of the system from the quadrupole to the octupole level of approximations. We also analyze the chaotic and quasi-periodic orbital evolution by studying the surfaces of sections. Furthermore, including general relativity, we show case the long term evolution of individual debris disk particles under the influence of a far away interior eccentric planet. We show that this dynamics can naturally result in retrograde objects and a puffy disk after a long timescale evolution (few Gyr) for initially aligned configuration.
astro-ph.EP
astro-ph
DRAFT VERSION OCTOBER 1, 2018 Preprint typeset using LATEX style emulateapj v. 12/16/11 THE ECCENTRIC KOZAI-LIDOV MECHANISM FOR OUTER TEST PARTICLE MACARENA ZANARDI3,4, GONZALO CARLOS DE ELÍA3,4, ROMINA P. DI SISTO3,4 SMADAR NAOZ1 & GONGJIE LI2 Draft version October 1, 2018 ABSTRACT The secular approximation of the hierarchical three body systems has been proven to be very useful in address- ing many astrophysical systems, from planets, stars to black holes. In such a system two objects are on a tight orbit, and the tertiary is on a much wider orbit. Here we study the dynamics of a system by taking the tertiary mass to zero and solve the hierarchical three body system up to the octupole level of approximation. We find a rich dynamics that the outer orbit undergoes due to gravitational perturbations from the inner binary. The nominal result of the precession of the nodes is mostly limited for the lowest order of approximation, how- ever, when the octupole-level of approximation is introduced the system becomes chaotic, as expected, and the tertiary oscillates below and above 90◦, similarly to the non-test particle flip behavior (e.g., Naoz 2016). We provide the Hamiltonian of the system and investigate the dynamics of the system from the quadrupole to the octupole level of approximations. We also analyze the chaotic and quasi-periodic orbital evolution by studying the surfaces of sections. Furthermore, including general relativity, we show case the long term evolution of individual debris disk particles under the influence of a far away interior eccentric planet. We show that this dynamics can naturally result in retrograde objects and a puffy disk after a long timescale evolution (few Gyr) for initially aligned configuration. 1. INTRODUCTION The hierarchical three body secular dynamics has been studied extensively in the literature and was shown to be very effective in addressing different astrophysical phenom- ena (see for review Naoz 2016, and reference therein). In this hierarchical setting the inner binary is orbited by a third body on a much wider orbit, the outer binary, such that the secular approximation can be applied (i.e., phase averaged, long-term interaction). The gravitational potential is then expanded in semi-major axis ratio (a1/a2, which, in this approximation, remains constant), where a1 (a2) is the semi-major axis of the inner (outer) body (Kozai 1962; Lidov 1962). This ra- tio is a small parameter due to the hierarchical configuration. The lowest order of approximation, which is proportional to (a1/a2)2 is called the quadrupole-level. Most of these studies focus on the gravitational perturba- tions that a far away perturber exerts on the inner binary. In early studies of high-inclination secular perturbations (Kozai 1962; Lidov 1962), the outer orbit was assumed to be circular and it was assumed that one of the inner binary members is a massless test particle. In this situation, the component of the inner orbit's angular momentum along the z-axis (which is set to be parallel to the total angular momentum, i.e., the invariable plane) is conserved, and the lowest order of the ap- proximation, the quadrupole approximation, is valid. How- ever, relaxing either one of these assumptions leads to qual- itative different behavior (e.g., Naoz et al. 2011; Lithwick & Naoz 2011; Katz et al. 2011). Considering systems beyond the test particle approximation, or a circular orbit, requires [email protected] Angeles, CA 90095, USA 1 Department of Physics and Astronomy, University of California, Los 2 Harvard Smithsonian Center for Astrophysics, Institute for Theory and Computation, 60 Garden Street, Cambridge, MA 02138, USA 3 Instituto de Astrofísica de La Plata, CCT La Plata-CONICET-UNLP Paseo del Bosque S/N (1900), La Plata, Argentina 4 Facultad de Ciencias Astronómicas y Geofísicas, Universidad Na- cional de La Plata Paseo del Bosque S/N (1900), La Plata, Argentina the next level of approximation, called the octupole -- level of approximation, which is proportional to (a1/a2)3 (e.g. Har- rington 1968, 1969; Ford et al. 2000; Blaes et al. 2002). In the octupole level of approximation, the inner orbit eccentric- ity may reach extreme values (Ford et al. 2000; Naoz et al. 2013a; Li et al. 2014c; Teyssandier et al. 2013). In addition, the inner orbit can flip its orientation, with respect to the total angular momentum (i.e., z-axis), from prograde to retrograde (Naoz et al. 2011). Here we study the secular evolution of a far away test par- ticle orbiting an inner massive binary. In this case, the in- ner orbit is fixed, and effectively carries all of the annular momentum of the system, while the outer orbit undergoes a dynamical evolution. This situation has large range of appli- cations from the gravitational perturbations of binary super massive black holes on the surrounding stellar distribution to the effects of planetary orbits on debris disks, Oort cloud and trans-neptunian objects. From N-body simulations, Zanardi et al. (2017) analyzed the long term evolution of test particles in the presence of an interior eccentric planet. Such an study produces particles on prograde and retrograde orbits, as well as particles whose orbital plane flips from prograde to retro- grade and back again along their evolution. We note that Ziglin (1975) investigated the oscillations of an outer circumbinary planet in the context of the restricted elliptical three body problem. Later Verrier & Evans (2009) and Farago & Laskar (2010) studied the stability of high in- clined planet around in this situation using a combination of numerical and perturbation theory up to the quadruple level of approximation approaches (see also, Li et al. 2014a; de la Fuente Marcos et al. 2015). Furthermore, Gallardo (2006) and Gallardo et al. (2012) studied the effects of the Kozai-Lidov for trans-neptunian objects near mean motion resonance with Neptune. However, here, we do not allow for mean motion resonances to allow for the double averaging process (see Naoz et al. 2013a, appendix A2 for the canonical transfor- mation which describes the averaging process).We provide a general treatment for the outer-test particle case, up to the 2 S. NAOZ ET AL octupole -- level of approximation in the secular theory. The paper is organized as follow: We begin by describing the outer test particle Hamiltonian and equations of motion (§2), and continue to discuss the quadrupole-level of approx- imation (§3) where we also drive the relevant timescales, and then we study the role of the octupole-level of approxima- tions, and provide surface of section maps (§4). We also dis- cuss the role of general relativity precession in §5. We then consider one study case in the form of the long term evolution debris disk particles in §6. Finally, we offer our discussions in §7. 2. THE EQUATIONS OF MOTION We solve the orbit of an exterior massless test particle to an eccentric planet (m2), both orbiting a star (m1), including only secular interactions expanded to octupole order. The planet is on a fixed eccentric orbit (i.e., e1 = const) and the outer particle's orbit is specified by four variables: e2, ω2, θ, Ω2 , (1) where e2 is the test particle eccentricity, θ = cosi and i is the inclination of the test particle with respect to the inner or- bit, and ω2 and Ω2 are the argument of periapse and longi- tude of ascending node of the outer orbit, relative to the inner planet's periapse (Murray & Dermott 2000). Specifically, We set ϖ1 = 0, see Appendix A for the coordinate transformation. We kept the subscript "2" in ω2 and Ω2 for consistency with the comparable masses treatments. From e2 and θ we can de- fine the canonical specific momenta (cid:113) (cid:113) J2 = 1 − e2 2 J2,z = θ 1 − e2 2 Figure 1. The quadrupole-level of approximation evolution. We show the time evolution of the inclination in the left panels and the cross section trajectory in the inclination-Ω2 plane, in the right panels. We consider two cases, the top panels are for e1 = 0.9 and the bottom panels are for e1 = 0.3. We show the following examples: in the circulating mode (setting initially Ω2 = 0): i = 20◦ (cyan), 60◦ (blue), and in the librating mode (setting initially Ω2 = 90◦): i = 20◦ (brown) and 60◦ (magenta). The separatrix corresponds to i = 90◦ for the two different inner eccentricities is shown in red. Note that in the case of e1 = 0.3 there is no librating mode for i = 20◦. For consistency we adopt the following orbital parameters: m1 = 1 M(cid:12), m1 = 1 MJ, a1 = 3 AU and a2 = 40 AU, ω2 = 90◦ and e2 = 0. and f = fquad + δ foct . (10) We note that fquad has the same functional form as the inner test particle Fquad presented in Lithwick & Naoz (2011) up to the (1 − e2 2)3/2 which is not constant in our case. The equations of motion may be expressed as partial deriva- tives of an energy function f (e2, ω2, θ, Ω2) dJ2 dτ dJ2,z dτ dω2 dτ dΩ2 dτ = ∂ f ∂ω2 = ∂ f ∂Ω2 = ∂ f ∂e2 = − ∂ f ∂θ J2 e2 1 J2 + ∂ f ∂θ θ J2 (11) (12) (13) (14) where τ is proportional to the true time (see Eq. [A1]). Unlike inner orbit test particle M is not constant while e1 is constant (in other words, the angular momentum of the inner orbit is conserved). The equations of motion were tested successfully compared to the general equations of motions, presented in Naoz et al. (2013a). We also test the evolution compared to N-body in Appendix B. 3. QUADRUPOLE-LEVEL OF APPROXIMATION 3.1. General Analysis (2) (3) (4) (5) , (6) The Hamiltonian for which m3 → 0 is f = fquad + M foct , M = m1 − m2 m1 + m2 a1 a2 e2 1 − e2 2 where and fquad = (2 + 3e2 1)(3θ2 − 1) + 15e2 2)3/2 (1 − e2 1(1 − θ2)cos(2Ω2) (cid:20) 1)θ(1 − θ2)sin ω2 sin Ω2 10(1 − e2 15e1 foct = 4(1 − e2 2)3/2 + 1 {2 + 19e2 1)θ2 − 35e2 2 × (θ sin ω2 sin Ω2 − cos ω2 cos Ω2) 1 − 5(2 + 5e2 (cid:21) 1(1 − θ2)cos(2Ω2)} (7) Note that unlike the inner test particle approximation M is not constant during the motion, and the constant parameter during the evolution is: δ = m1 − m2 m1 + m2 a1 a2 e1 (8) and the Hamiltonian up to the octupole level of approximation can be defined as: foct = e2 1 − e2 2 foct e1 (9) EKL FOR OUTER ORBIT TEST PARTICLE 3 (cid:18) a1 (cid:19)2 2π The quadrupole-level of approximation is integrable and thus provides a good starting point. Unlike the quadrupole- level approximation for the inner orbit test particle, here the z-component of the particle's angular momentum is not con- served, as the Hamiltonian depends on Ω2. However, at this level J2 is conserved, and thus the outer orbit eccentricity e2 remains constant as the Hamiltonian does not depends on ω2. The equation of motion for the inclination takes a simple form: dθ dt (cid:12)(cid:12)(cid:12)(cid:12)quad (15) = −15 8 m1m2 (m1 + m2)2 e2 1 1 − θ2 (1 − e2 a2 P2 2)2 sin2Ω2 , where we consider the time evolution and not the scaled evo- lution for completeness, and P2 is the period of the outer orbit. We can find the maximum and minimum inclination by setting θ = 0. Thus, we find that the values of the longitude of ascend- ing nodes that satisfy this condition are Ω2 = nπ/2, where n = 0,1,2... In other words, Ω2 has two classes of trajecto- ries, librating and circulating. The trajectories in the librating region are bound between two values of Ω2, while the circu- lation region represents trajectories where the angles are not constrained between two specific values. On circulating tra- jectories, at Ω2 = 0, the inclination (i < 90◦) is largest (where the i > 90◦ case is a mirror image of the i < 90◦ one). The extrema points for the librating mode are located at Ω2 = 90◦. The time evolution of Ω2 for the quadrupole-level of approxi- mation is given by: = − m1m2 (m1 + m2)2 2π P2 (cid:18) a1 a2 (cid:19)2 3θ(cid:0)2 + 3e2 1 − 5e2 1 cos2Ω2 8(1 − e2 2)2 (16) (cid:1) In Figure 1 we show the evolution associated for the quadrupole-level of approximation. The two librating and cir- culating trajectories are considered, where we folded the Ω2 angle to be between 0 − 180◦. The librating mode gives the nominal precession of the nodes, at which the inclination os- cillates between the i90 inclination (the inclination for which Ω2 = 90◦) and 180◦ − i90. The precession of the nodes was noted before in the literature (e.g., Innanen et al. 1997). From the latter Equation and Equation (15) we have (cid:12)(cid:12)(cid:12)(cid:12)quad dΩ2 dt (cid:115) Note that this expression can be also achieved by considering the conservation of energy between the minimum and maxi- mum cases. Setting the initial conditions for the energy, Equa- tion (6), we can find the extrema points as a function of the initial conditions. A special case can be considered when Ω2 is initially set to be zero, and then the maximum inclination is the initial inclination i0, in other words: (1 − e2 1) (1 + 4e2 1) for Ω2,0 = 0◦ . sinimin = sini0 (20) In other words we can set imax → 90◦ and for a given e1 find the largest imin allowed, which corresponds to the separatrix. This relationship is also apparent in figure 14 in Zanardi et al. (2017) numerical results. As can be seen, for the circulat- ing mode depicted in Figure 1, setting initially i = 20◦ corre- sponds to imin = 16.25◦ and 10.56◦ for e1 = 0.3 and e1 = 0.9, respectively, consistent with Equation (20). A comparison be- tween the flip criterion and the numerical results can be seen in Figure 9. For the librating mode we set imin which associated with the Ω2 = 90◦ case, and thus integrating over Equation (17) between i = 90◦ to imin (and Ω2 = Ω2,min to Ω2 = 90◦, respec- tively), we get, 2 + 3e2 1 − (2 + 8e2 1)sin2 imin . 5e2 1 cos2Ω2,min = (21) Thus, setting initially Ω2 = 90◦ as in the examples depicted in Figure 1 we find that the minimum value that Ω2 can achieve in the e1 = 0.9 case is 59.23◦ for the imin = 60◦ and 15.95◦ for the imin = 20◦ example, consistent with the numerical results. Hence, the range of which Ω2 is librating on is 2 × (90◦ − Ω2,min) . 3.2. Timescale The timescale associated with the evolution can be esti- mated from the equation of motion for Ω2 at the quadrupole- level [Equation (16)], for the circulating mode, by setting dΩ2 → π and taking the terms in the parenthesis to be roughly order of unity (which is achieved, by setting Ω2 → 0): tquad ∼ 4 3 2)2 (m1 + m2)2 m1m2 (cid:18) a2 circulating. P2(1 − e2 (cid:19)2 (22) a1 For the example system depicted in Figure 1, this equation gives a timescale of about 6 × 107 yr, for initial inclination of 60◦, which agrees with the circulating mode (although we note that different e1 give slightly different timescales). We also estimate the timescale in the librating mode by setting dΩ2 → 2×(90◦ − Ω2,min), which we have found earlier. Thus, (cid:19)2 tquad ∼ 2× 2(π/2 − Ω2,min) (cid:18) a2 (m1 + m2)2 2π m1m2 a1 P2(1 − e2 2)2 × 8 3 librating. (23) The 2 pre-factor of here comes from numerical comparisons to the examples depicted in Figure 1. Note the e1 dependency that rises from Ω2,min. This timescale is consistent with the examples depicted in Figure 1 by less than a factor of two. dΩ2 dθ = θ 2 + 3e2 5e2 1 − 5e2 1 cos2Ω2 1(1 − θ2)sin2Ω2 Integrating the two sides we have: (cid:90) θb . θ (17) dθ , (18) (cid:90) Ωa Ωb sin2Ω2 1 − 5e2 1 cos2Ω2 2 + 3e2 dΩ2 = 5e2 1(1 − θ2) θa where Ωa,b is the longitude of ascending nodes that is associ- ated with the inclination value of θa,b = cosia,b. For the circulating mode, we find that, after integrating over Equation (17) from imax to imin (and Ω2 = 0 to Ω2 = 90◦, re- spectively) we get (cid:115) sinimin = sinimax 1 − e2 1 1 + 4e2 1 . (19) 4. THE ROLE OF THE OCTUPOLE LEVEL OF APPROXIMATION 4 S. NAOZ ET AL Figure 2. The role of octupole We consider two cases, the secular evolution up to the quadrupole-level of approximation (blue lines) and up to the octupole level of approximation (red lines). We show the time evolution of the inclination and the outer orbital eccentricity (which remains constant at the quadrupole-level of approximation). We also consider the inclination evolution as a function of Ω2. Left panels: We consider the following system: m1 = 1 M(cid:12), m2 = 1 M j, a1 = 0.4 AU, a2 = 7 AU, e1 = 0.65 and e2 = 0.4. We initialize the system with ω2 = Ω2 = 0◦ and i = 91◦. Right panels: m1 = 1 M(cid:12), m2 = 1 M j, a1 = 3 AU, a2 = 50 AU, e1 = 0.85 and e2 = 0.7. We initialize the system with ω2 = 0◦, Ω2 = 40◦ and i = 20◦. The octupole level of approximation can significantly affect the overall dynamics of the general hierarchical three body system (see Naoz 2016, and reference therein). Specifically, in the inner test particle case, the inner orbit's z component of the angular momentum is not conserved anymore and the orbit is allowed to flip (for large range of initial inclinations Lithwick & Naoz 2011; Li et al. 2014c,b). In our case, the J2,z is not conserved at the quadrupole-level, but J2 is. Thus, the octupole level of approximation in this case allows for variations of e2 and introduces higher level resonances, which may result in a chaotic behavior (see below). In Figure 2 we consider two representative example for which we compare the quadrupole (blue lines) and octupole (red lines) levels of approximation, where we consider the time evolution of the eccentricity and inclination. In both of these examples we consider a 1 M(cid:12) star orbited by an ec- centric Jupiter, with a test particle on a far away orbit. One can consider such a setting to represent a result of a scattering event for example. On the left set of panels of Figure 2 we consider a Jupiter at 0.4 AU with e1 = 0.65 and a test particle at 7 AU with e2 = 0.4, initialized on a retrograde orbit (i = 91◦). With the introduc- tion of the octupole level of approximation to the calculation, the test particle eccentricity starts to oscillate, though in this case it never increases pass its initial value (due to choice of initial conditions here). More notably, the test particle incli- nation, with respect to the total angular momentum, oscillates from retrograde (> 90◦ which was the initial condition) to prograde (< 90◦). As in the more general case, there is no ap- parent associated timescale for this flipping modulation and it seems chaotic in nature (see below). While the quadrupole- level is circulatory in nature (see i− Ω2 plot) a libration behav- ior emerges at the octupole level. On the right set of panels of Figure 2 we consider a Jupiter at 3 AU with e1 = 0.85 and a test particle at 50 AU with e2 = 0.7, and the system is initialized on a prograde orbit (i = 20◦). Here, like the previous example the outer test parti- Figure 3. We consider the system: m1 = 1 M(cid:12), m2 = 1 M j, a1 = 0.5 AU, a2 = 10 AU, e1 = 0.4 and e2 = 0.6. We initialize the system with ω2 = Ω2 = 0◦ and i = 85◦. We show the time evolution of the orbital parameters, i.e., argument of pericenter ω2, longitude of ascending node Ω2, inclination i and the outer orbital eccentricity, e2. We also consider the inclination evolution as a function of Ω2. Here both Ω2 and ω2 were folded to achieve the 0− 180◦ symmetry. cle eccentricity begins to oscillate, and even grows above the initial value. However, unlike the previous example this sys- tem does not flip. The inclination does oscillate with a long scale modulation and we show the long scale evolution that captures about four octupole cycles. The system does not ex- hibit a chaotic behavior in this case, and it remains in a circu- EKL FOR OUTER ORBIT TEST PARTICLE 5 latory trajectory even after the inclusion of the octupole level of approximation to the calculation. In Figure 3 we zoom-in on the evolution of a different example, and also provide the time evolution of ω2 for the octupole level itself. As in the general hierarchical secu- lar three body problem we find the short scale (associated with the quadrupole) oscillations, that are modulated by the higher level octupole approximation. The octupole modula- tions take place on timescales which is between few×tquad to few tens×tquad. Its interesting to note that the inclination flips shown in Fig- ures 2 and 3 are qualitatively different from the ordered, back and forth oscillation of the quadrupole-level of approximate evolution. The latter produces a simple, ordered oscillation of the inclination angle between i0 and 180◦ −i0, for the librating regime. However, in the presence of the octupole-level of ap- proximation the system behaves similarly to the general flips discussed in Naoz et al. (2011), where the inclination oscil- lates for sometime at the prograde (i < 90◦) regime, and then flips to retrograde configuration (i > 90◦). The eccentricity, e2, gives rise to an additional complication as in essence the outer test particle eccentricity can grow so much until the orbits will cross. We adopt the nominal stabil- ity criterion  = a1 a2 e2 1 − e2 2 < 0.1 (24) to guid us when the system leaves stability. We discuss this stability criterion in the context of N-body comparisons in ap- pendix B. To explore the chaotic nature of the system and the differ- ent dynamical regimes we use surface of sections. The outer test particle approximation reduces the general hierarchical three body system from six degrees of freedom to four de- grees of freedom. In addition, in the test particle limit, the in- ner orbit is stationary and reduces the system to two degrees of freedom. In this system f and δ are the only conserved parameters and ω2 and Ω2 are the only coordinates that can change with time. For a two-degrees-of-freedom system, the surface of section projects a four-dimensional trajectory on a two-dimensional surface, where we select intersections of the trajectories on the surfaces when ω2 and Ω2 moves in the positive directions. For simplicity, we separate the two initial conditions into three characteristic parameters: e1, the inner orbit eccentricity, which remains constant during the evolu- tion, the energy, or initial value of the reduced Hamiltonian f , and δ = m1 − m2 m1 + m2 a1 a2 e1 . Note that the energy depends on δ and e1, so in fact, although we choose to characterize the surface of sections by three pa- rameters, there are only two independent ones. In Figures 4-6 we consider the surface of sections for var- ious values of the f , e1 and δ, in the J2 − ω2 plane (top row in each figure) and in J2,z − Ω2 plane (bottom row in each fig- ure). In both planes we identify the resonances at which the momenta and angles undergo bound oscillations. The trajec- tories in this region are quasi-periodic, and the system is in the libration mode. The circulation region represents trajecto- ries where the angles are not constrained between two specific values. Both librating and circulatory trajectories are mapped onto a one-dimensional manifold on the surface of section and they form lines on the section. However, chaotic trajectories are mapped onto a higher dimensional manifold and they are filling an area on the surface. We note that in some of the trajectories in the Figures, due to sampling limitation, seem as dashed lines, but they actually represent a one-dimensional manifold. In all of the maps we indicate the instability regime (light orange stripe) for which  > 0.1. We intersect the trajectories at Ω2 = π/2 to produce the sur- faces in the J2 − ω2 plane, in order to capture the librating cases. The empty regions at large J2 at the parameter spaces in the far left and right panels in Figure 5 (i.e,. for the param- eters: e = 0.4, δ = 0.02 and f = ±10) and the far left panel in Figure 6 ( for the parameters: e = 0.9, δ = 0.1 and f = −10) correspond to regions with no physical solutions. The vari- abilities in J2 are mostly small in the stable regime. We see that there are regular behavior, i.e., trajectories which fill one- dimensional lines on the surface of section in most of the sta- ble region. We find the emergence of chaos in parts of the unstable zones, in particular when J2 is low (e2 is high). Considering the J2,z − Ω2 plane, we intersect the trajectories at ω2 = 0. The system exhibits a chaotic behavior across all pa- rameter regime of e1, δ and f . Most of the circulation region, associated with curve, non chaotic one-dimensional manifold, are typically associated with J2,z ∼> 0.3. The outer orbits can flip (J2,z shifts signs) in most of the parameter space. Resonances can be easily identified in a few of the maps. Specifically, in the J2 − ω2 plots, the resonances can be found centered near ω2 = π (e.g., e1 = 0.4, δ = 0.02, f = −10, and e1 = 0.9, δ = 0.1, f = −10, etc.), and ω2 = π/2 (e.g., e1 = 0.4, δ = 0.02, f = 10). The dynamics is quite complicated when e1 is higher and when δ is larger, and higher order resonances (appearing as small liberating islands) emerge in the surface of section in the J2 − ω2 plane when e1 = 0.9, δ = 0.1 and f = 0. Resonances can also be identified in the J2,z − Ω2 plane, such as e1 = 0.9, δ = 0.1, f = 10. 5. THE ROLE OF THE GENERAL RELATIVITY As noted previously in many studies, General Relativity (GR) precession tends to suppress the inner orbit eccentricity excitations associated with the Eccentric Kozai-Lidov (EKL) mechanism, and thus suppress the flips (e.g., Naoz et al. 2013b). In our secular case, the inner orbit is massive and the outer orbit is a test particle, so practically the inner orbit does not feel the outer orbit gravitational interactions. However, the inner orbit can still precess due to GR with the nominal precession rate (e.g., Naoz et al. 2013b): = 3k3(m1 + m2)3/2 a5/2 1 c2(1 − e2 1) , (25) where k2 is the gravitational constant and c is the speed of light. However, in our frame of reference, where the in- ner orbit carries all the angular momentum, we are basically working in the rotating frame of the inner orbit. Therefore, since we set ω1 = −π − Ω2, (see Appendix A), GR precession of ω1 translates to a precession of Ω2. Thus, using our coor- dinate transformation we find: (cid:12)(cid:12)(cid:12)(cid:12)GR,inner dω1 dt (cid:12)(cid:12)(cid:12)(cid:12)GR,Ω2 dΩ2 dt = −3k3(m1 + m2)3/2 a5/2 1 c2(1 − e2 1) , (26) which can suppress the inclination oscillations. The timescale associated with that precession is the nominal GR one: tGR,Ω2 ∼ 2π a5/2 1 c2(1 − e2 1) 3k3(m1 + m2)3/2 . (27) 6 S. NAOZ ET AL Figure 4. Surface of section. We consider e1 = 0.4, δ = 0.1 and, from left to right f = −0.2,0 and 0.2. We note that the cases f = 10 and f = −10 gives a similar map to the f = 0.2 and f = −0.2 respectively and thus, were not depicted, to avoid clutter. The light orange stripe marks the unstable regime, for which  > 0.1. Figure 5. Surface of section. We consider e1 = 0.4, δ = 0.02 (compared to Figure 4, this means changing the factor (m1 − m2)a1/a2/(m1 + m2) by factor of 5) and, from left to right f = −10,−0.2,0,0.2 and 10. The light orange stripe marks the unstable regime, for which  > 0.1. In Figure 7 we consider two examples, one for which tGR,Ω2 ∼ 5× 106 yr is smaller than tquad ∼ 6× 106 yr (top panel), and the other of which tGR,Ω2 ∼ 4.4×108 yr is a bit longer than the corresponding tquad ∼ 2×107 yr (bottom panel). Both of these examples had a Sun size star and a Jupiter planet, orbited by a far away test particle. The Jupiter has a non-negligible eccen- tricity, that perhaps can be a result of either a scattering event or a high eccentricity migration. In the top panel the Jupiter was set at a1 = 0.5 AU with e1 = 0.4 and the test particle was set at a2 = 10 AU and e2 = 0.6. The system in the absence of GR was in libration mode and as noted before exhibited a chaotic nature. However, the GR precession bound the system into a circulatory regime, and suppressed the flips. In the bot- tom panel the Jupiter was set at a1 = 3 AU and e1 = 0.9, while the test particle was set at a2 = 40 AU with e2 = 0.65. Note that in this latter case, although the Jupiter is rather far from the host star and the GR precession timescale is longer than the secular precession timescale, the GR changes the dynam- ics. Specifically, before the inclusion of GR precession, the EKL FOR OUTER ORBIT TEST PARTICLE 7 Figure 6. Surface of section We consider e1 = 0.9, δ = 0.1 and, from left to right f = −10,−0.2,0,0.2 and 10. The light orange stripe marks the unstable regime, for which  > 0.1. Figure 7. The role of GR. We consider two cases, the evolution without GR (red lines) and the evolution with GR (blue lines). Top panel: We consider the following system: m1 = 1 M(cid:12), m2 = 1 M j, a1 = 0.5 AU, a2 = 10 AU, e1 = 0.4 and e2 = 0.6. We initialize the system with ω2 = Ω2 = 0◦ and i = 85◦. For this system we find that tGR,inner ∼ 5× 106 yr which is much shorter than the quadrupole timescale. Bottom panel: We consider the following system: m1 = 1 M(cid:12), m2 = 1 M j, a1 = 3 AU, a2 = 40 AU, e1 = 0.9 and e2 = 0.65. We initialize the system with ω2 = 90◦ and Ω2 = 100◦ and i = 20◦. The GR precession timescale for this system is estimated as tGR,inner ∼ 4.4× 108 yr, which is longer than the quadrupole timescale. This is a typical situation to an individual debris disk particle or a icy body reservoir object (see section 6). system was in a libration mode and seemed quasi-periodic, however, after the inclusion of GR the system exhibits both libration and circulation, and the emergence of chaotic behav- ior seems to take place. The dramatic change in dynamical behavior with the inclusion of GR precession, even if takes place on longer timescales than the secular timescales, was noted previously in the general and inner test particle case in Naoz et al. (2013b). Figure 8. Debris disk particles. We consider the time evolution of a test particle located at 55 AU from a M(cid:12) star due to the gravitational perturba- tions from an eccentric Jupiter a1 = 5 AU and e1 = 0.85. We initialized the system with e2 = 0.5, Ω2 = 90◦, ω2 = 0◦ and i = 20◦. We consider from top to bottom the inclination, eccentricity and Ω2. The transition between librating and circulating can clearly be seen in the bottom panel. When the angle is in circulation mode, it increases in value as a function of time. We note that in all of our calculation below we also take into account the outer orbit GR precession (e.g., Naoz et al. 2013b): (cid:12)(cid:12)(cid:12)(cid:12)GR,outer dω2 dt = 3k3(m1 + m2)3/2 a5/2 2 c2(1 − e2 2) , (28) which typically takes place on much larger timescales. 6. A STUDY CASE APPLICATION: INDIVIDUAL DEBRIS DISK PARTICLES Debris disks mark the late end stages of planet formation, and are made of the leftover material of rocks and ices. The 8 S. NAOZ ET AL gravitational interactions between these particles and interior or exterior companion can leave a distinct imprint on the mor- phology of the disk and can cause dust production (e.g., Rodi- gas et al. 2014; Matthews et al. 2014; Nesvold & Kuchner 2015; Lee & Chiang 2016; Nesvold et al. 2016). Many of these studies typically focus on few million years of integra- tion to allow for comparison of observations which usually can detect young systems. Here we allow for longer integra- tion timescales and investigate the evolution of a test particle under the influence of an eccentric Jupiter. In Figure 8 we show an example system, where we consider an eccentric Jupiter at 5 AU, with 0.85 eccentricity orbiting a one solar mass star. The test particle is located at 55 AU and initialized with an eccentricity of 0.5. We integrate the octupole level equations of motion in the presence of GR pre- cession for both the inner and outer orbits. As in the example depicted in Figure 7, which also considered an icy body or comet reservoirs analogs, the orbit switches between libration and circulation as can be seen in the bottom panel. We consider the effect of the planet's eccentricity, e1, and the test particle inclination by surveying the parameter space of e1 and initial inclination for a given system, where we set a2 = 55 AU, with e2 = 0.5, and set the system initially with Ω2 = 90◦ and ω2 = 0◦. Many giant exoplanets have high eccen- tricities, specifically for giant planets (msini > 0.1 MJ, with separation > 0.05 AU) the average eccentricity is ∼ 0.2, and a maximum value of 0.975. Thus, an eccentric Jupiter doesn't seem like an unlikely configuration for a planetary system. In the example depicted in Figure 8 the time evolution of the test particle's inclination that starts with a moderate eccen- tricity oscillates between extreme values 20◦ ∼ 160◦ without increasing its eccentricity. The system depicted in Figure 8 is, of course, just one ex- ample for a particular choice of the orbital parameters. To study the effects of planet's eccentricity and initial inclination we have systematically explored the e1-i parameter space in Figure 9. We choose a Jupiter like system (a1 = 5 AU) and for a range of eccentricities, with an outer orbit at 55 AU. The test particle orbit was initialized with e2 = 0.5, Ω2 = 90◦, ω2 = 0◦ and a range of inclinations. We show the maximum inclina- tion reached during the evolution as a function of the initial inclination and eccentricity in Figure 9. In the left panel we depict the initial inclination vs the initial Jupiter's eccentricity, where the color code marks the maximum inclination reached. The solid line in the left panel follows Equation (19), which is consistent that the resonance associated with the quadrupole- level of approximation is indeed the main driver for the dy- namical evolution of the system. In the figure we depicted the initial inclination regime to the prograde case (iinitial ≤ 90◦) to avoid clutter. However, we show in the right panel the maxi- mum inclination reached during the evolution as a function of the initial inclination going all the way to 180◦ this time. As a proof of concept we depict in Figure 10 the behavior of a narrow debris disk after 4 Gyr of evolution. This inclination represents the instantaneous inclination at 4 Gyr of evolution. The system, of course, continue to oscillate, and the disk of particles will remain puffed. The inclination and eccentric- ity of the system at this snapshot are qualitatively different from the initial conditions assumed. This hypothetical sys- tem shows the orbital configuration of a disk located between 55 − 65 AU, with an interior eccentric Jupiter (e1 = 0.85) at 5 AU around a solar mass star. The system was set initially 5 Taken from The Exoplanet Orbit Database (Wright et al. 2011). with a mutual inclination of 20◦, e2 = 0.3 and Ω2 and ω2 are chosen from a random uniform distribution between 0− 360◦. While the particles in the disk have eccentric values, the disk does not appear as a coherent eccentric ring as the values of Ω2 and ω2 are random. At the end of the integration, the parti- cles in the disk became slightly more eccentric (with an aver- age eccentricity of ∼ 0.34) and there is a clear trend of Ω2 as a function of inclination. The particles with inclination above the initial 20◦ have a value of Ω2 close to zero (or with the symmetric value 180◦), while the particles with inclination below 20◦ have Ω2 values closer to 90◦. The behavior is sin- gular to Ω2 and is not manifest itself in ω2 (does not depicted here to avoid clutter). We note that some retrograde parti- cles were formed as well, in line with Zanardi et al. (2017) numerical results of an orbital flip. These behaviors are eas- ily understandable from the surface of section maps depicted above. Its important to note that during the evolution of the system depicted in Figure 10, the maximum  achieved was 0.0625. The average value of  which corresponds to the maximum e2 achieved during the evolution was ∼ 0.4. Thus, the system is kept stable during this evolution, the secular approximation holds, and we do not expect any scattering event. We also note that we have compared a debris disk particles secular and N-body evolution and found a qualitative agreement, which is similar in behavior to Figure 11 left panel, and this is not shown here to avoid clutter. 7. DISCUSSION We have studied the secular evolution of an outer test parti- cle hierarchical system. We presented the three body, outer test body Hamiltonian up to the octupole level of approx- imation in the power series of the semi-major axis ratio. We showed that in the quadrupole-level of approximation, (a1/a2)2, the system has two distinct behaviors, librating and circulating (see Figure 1), where the librating mode gives the nominal precession of the nodes, results for which the incli- nation oscillates between the i90 inclination (the inclination for which Ω2 = 90◦ and 180◦ − i90). Furthermore, the bound values of the liberating mode have a simple analytical expres- sion, Eq. (21). We also found the minimum and maximum inclination that the system can reach in the circulating mode (see Equations (19) and (20)). These conditions are sensitive to the initial inner orbit's eccentricity, and are nicely repro- duced in numerical testing here (see Figure 9) and in Zanardi et al. (2017) numerical experiments (see their figure 14). We also estimated the timescale for oscillations for the two modes (see Section 3.2). We then showed that introducing the octupole level of ap- proximation allows for transition between the two libration and circulation modes (see Figures 2 and 3). This yields that the overall dynamics of the system is similar to the behavior of the general flip behavior in the eccentric Kozai-Lidov mecha- nism. In particular, the dynamics is quite chaotic for parame- ter regions with high e2 and perpendicular mutual inclinations (when J2 is low and when J2,z is near zero), as shown in the surface of sections Figures 4-6. General relativity can play an important role in suppress- ing or exciting the eccentricities in the hierarchical three body problem (e.g., Naoz et al. 2013b). We find here similar behav- ior. Specifically, the inclination excitation will be suppressed for systems with GR precession faster than the quadrupole precession. However, when GR precession takes place on similar (or even somewhat larger) timescales to that of the EKL FOR OUTER ORBIT TEST PARTICLE 9 Figure 9. Debris disk particles. We consider a Sun-Jupiter like system, setting Jupiter at 5 AU. We set the test particle at 55 AU, with e2 = 0.5. We vary systematically the inclination and Jupiter's eccentricity. We set the system initially with Ω2 = 90◦ and ω2 = 0◦. Left panel shows the initial condition map (inclination vs the Jupiter's eccentricity) where the color code depicts the maximum inclination the system reached during its evolution of 4 Gyr. The solid line shows the analytical equation to reach 90◦, Equation (19). The right panel shows the maximum inclination reached as a function of the initial inclination (that goes all the way to 180◦). The color code here marks the inner orbit's eccentricity. a typical example of the evolution of a test particle due to the gravitational perturbations from an eccentric Jupiter (see Fig- ure 8). We systematically varied the Jupiter's eccentricity and the outer orbit's inclination, where we found an agreement between the analytical relation for crossing the 90◦ thresh- old, and the numerical tests. This also suggests that eccentric planet can pump up the inclination of icy body or comet reser- voir analogs (Figure 9). This was further supported by con- sidering the evolution of an initially narrow, thin disk of test particles exterior to an eccentric planet, with initial mutual inclination of 20◦. The disk became puffed with some parti- cles on a retrograde orbits (see Figure 10). A detailed study of the effects of eccentric planets on exterior test particle is presented in Zanardi et al. (2017). This mechanism will also be important for example to circumbinary planetary systems (considered first by Ziglin 1975) and the stars near mergers of black hole binaries, but detailed studies is beyond the scope of this paper. We thank the referee for a quick and detailed report, and es- pecially for his/her inquiry about the surface of sections. We also thank Vladislav Sidorenko for pointing out some missing references. SN acknowledges partial support from a Sloan Foundation Fellowship. GL is supported in part by the Har- vard William F. Milton Award. MZ, GdE, and RPD acknowl- edge the financial support given by IALP, CONICET, and Agencia de Promoción Científica, through the PIP 0436/13 and PICT 2014-1292. Figure 10. Long timescale evolution of a debris disk. The mutual inclina- tion, i, as a function of the semi-major axis, a2, after 4 Gyrs of integration. The color code shows the longitude of ascending nodes Ω2 −90◦ at that time (this presentation emphasizes the symmetry in the system). The right inset shows the histogram of the final eccentricity of the disk, while the left inset shows the histogram of the final mutual inclination of the disk. As a proof of concept we consider a narrow debris disk located between 55 − 65 AU, with an interior eccentric Jupiter (e1 = 0.85) at 5 AU around a solar mass star. The system was initially set with a mutual inclination of 20◦, e2 = 0.3 and Ω2 and ω2 are chosen from a random uniform distribution between 0 − 360◦. The results depicted here achieved by integrating over the equations of motion, Eqs. (11)-(14). Note that GR effects are included here as well. quadrupole precession, the additional precession can produce inclination excitations, in a non-regular manner (see Figure 7). The dynamics of these type of systems can have a wide range of applications, from stars around supermassive black hole binaries to the evolution of individual debris disk parti- cles. We have chosen the latter as an example and presented APPENDIX A. ORBITAL PARAMETERS AND THE SCALED TIME One might have expected that ω1 = const when the outer particle is massless. But in truth Ω1 is undefined because the reference plane is aligned with the inner orbit. Therefore, the inner planet must only have ω1 + Ω1 = const, and we may choose without loss of generality the constant to equal zero. Hence, elimination of the nodes (i.e., Ω1 − Ω2 = π) implies ω1 = −π − Ω2. 00.20.40.60.810102030405060708090e1,initialiinitial [deg] 204060801001201401600306090120150180020406080100120140160180iinitial [deg] imax [deg] 0.10.20.30.40.50.60.70.80.9imax [deg]e1,initial5556575859606162636465020406080100120140a2 [AU]i [deg] 00.10.20.30.40.50306090120e2N10203040506070800306090120050100150i [deg]NAfter 4Gyrs of evolutionInitial i=20°Initiale2=0.3Ω2−90° [deg] 10 S. NAOZ ET AL Figure 11. Comparison between the N-body results and the secular approximation at the octupole level. We show the inclination, eccentricity and Ω2 + ω2 as a function of time in the left hand side and i − Ω2 in the right hand panel. Red lines correspond to the secular calculation (up to the octupole-level of of approximation) and blue lines corresponds to the N-body calculation. Note that in the i − Ω2 we depict the N-body results as points to allow for an easier comparison. On the left side we consider the following system: m1 = 1 M(cid:12) m2 = 1 MJ, a1 = 0.3 AU, a2 = 40 AU, e1 = 0.9, e2 = 0.65, ω2 = 90◦, Ω2 = 100◦, and i = 20◦. On the right side we consider the following system: m1 = 1 M(cid:12) m2 = 1 MJ, a1 = 0.5 AU, a2 = 10 AU, e1 = 0.4, e2 = 0.6, ω2 = Ω2 = 0◦, and i = 85◦. The evolution of this system was depicted in Figure 3 and here we show this system for shorter evolution timescale to allow for a better comparison between the N-body and secular calculation. Similarly to the treatment done in Lithwick & Naoz (2011) we have rescaled the momenta by an arbitrary constant to achieve the specific angular momentum. The Hamiltonian is rescaled by the same constant, and we find that the rescaled time is: m1m2 GN(m1 + m2) t 16 τ = (A1) where t is the true time. The numerical factor 16 comes by taking m3 → 0 in the general Hamiltonian (see Naoz 2016, for the general form of the hierarchical three body double averaged hamiltonian). There is a choice to be made, to either scale the Hamiltonian by this numerical factor or τ. Here, we choose to absorb this number in τ to be consistent with the inner test particle Hamiltonian. (m1 + m2)2 (m1 + m2)2 a3 2 a2 a2 , m1m2 = t 16 2π P2 (cid:115) (cid:18) a1 (cid:19)2 (cid:18) a1 (cid:19)2 B. COMPARISON WITH N-BODY In this section, we compare the secular approximation at the octupole level in the test particle limit with the N-body simulation, using Mercury code (Chambers & Migliorini 1997). In this comparison we did not include GR. Good agreements can be reached when the apocenter distance of the inner binary is much smaller than the pericenter distance of the outer binary. In particular, we include an illustrative example here in Figure 11, where we consider an eccentric (e = 0.9 Jupiter at 3 AU around a solar like object. The test particle is set at 40 AU with e2 = 0.65. The system is set initially with i = 20◦, ω2 = 90◦ and Ω2 = 100◦. As shown in Figure 11, the secular approximation (shown as red lines) agrees qualitatively well with the N-body results for the eccentricity and inclination oscillations. This is likely due to the double averaging process, but nonetheless, the maximum and minimum of the orbital parameters are conserved in both N-body and secular calculations. It is interesting to note that similarly to (Lithwick & Naoz 2011) the approximation holds as long as  < 0.1. However, unlike the inner-test particle case, e2 can change and increase during the evolution which may break the validity of the approximation during the evolution, this is shown in Figures 12. REFERENCES Blaes, O., Lee, M. H., & Socrates, A. 2002, ApJ, 578, 775, arXiv:astro-ph/0203370 Chambers, J. E., & Migliorini, F. 1997, in Bulletin of the American Astronomical Society, Vol. 29, AAS/Division for Planetary Sciences Meeting Abstracts #29, 1024 -- + de la Fuente Marcos, C., de la Fuente Marcos, R., & Aarseth, S. J. 2015, MNRAS, 446, 1867, 1410.6307 Farago, F., & Laskar, J. 2010, MNRAS, 401, 1189, 0909.2287 Ford, E. B., Kozinsky, B., & Rasio, F. A. 2000, ApJ, 535, 385 Gallardo, T. 2006, Icarus, 181, 205 Gallardo, T., Hugo, G., & Pais, P. 2012, Icarus, 220, 392, 1205.4935 Harrington, R. S. 1968, AJ, 73, 190 -- -- . 1969, Celestial Mechanics, 1, 200 Innanen, K. A., Zheng, J. Q., Mikkola, S., & Valtonen, M. J. 1997, AJ, 113, 1915 Katz, B., Dong, S., & Malhotra, R. 2011, ArXiv e-prints, 1106.3340 Kozai, Y. 1962, AJ, 67, 591 Lee, E. J., & Chiang, E. I. 2016, The Astrophysical Journal, 1605.06118 Li, D., Zhou, J.-L., & Zhang, H. 2014a, MNRAS, 437, 3832 Li, G., Naoz, S., Holman, M., & Loeb, A. 2014b, ArXiv e-prints, 1405.0494 EKL FOR OUTER ORBIT TEST PARTICLE 11 Figure 12. Comparison between the N-body results and the secular approximation at the octupole level for systems around  ∼ 0.1. We show the inclination, eccentricity. Red lines correspond to the secular calculation (up to the octupole-level of of approximation) and blue lines corresponds to the N-body calculation. Left side: we consider the following system: m1 = 1 M(cid:12) m2 = 1 MJ, a1 = 1 AU, a2 = 15 AU, e1 = 0.8, e2 = 0.8, ω2 = 0◦, Ω2 = 90◦, and i = 16◦. These parameters imply an initial  of 0.148. Right side: we consider the following system: m1 = 1 M(cid:12) m2 = 1 MJ, a1 = 3 AU, a2 = 40 AU, e1 = 0.9, e2 = 0.65, ω2 = 90◦, Ω2 = 100◦, and i = 20◦. The latter systems initialize with  = 0.084 however, as time goes by the eccentricity grows and the approximation breaks. Li, G., Naoz, S., Kocsis, B., & Loeb, A. 2014c, ApJ, 785, 116, 1310.6044 Lidov, M. L. 1962, planss, 9, 719 Lithwick, Y., & Naoz, S. 2011, ApJ, 742, 94, 1106.3329 Matthews, B. C., Krivov, A. V., Wyatt, M. C., Bryden, G., & Eiroa, C. 2014, in Protostars and Planets VI, 521 -- 544, arXiv:1401.0743v1 Murray, C. D., & Dermott, S. F. 2000, Solar System Dynamics, ed. Murray, C. D. & Dermott, S. F. Naoz, S. 2016, ARA&A, 54, 441, 1601.07175 Naoz, S., Farr, W. M., Lithwick, Y., Rasio, F. A., & Teyssandier, J. 2011, Nature, 473, 187, 1011.2501 -- -- . 2013a, MNRAS, 431, 2155, 1107.2414 Naoz, S., Kocsis, B., Loeb, A., & Yunes, N. 2013b, ApJ, 773, 187, 1206.4316 Nesvold, E. R., & Kuchner, M. J. 2015, The Astrophysical Journal, 798, 83 Nesvold, E. R., Naoz, S., Vican, L., & Farr, W. M. 2016, The Astrophysical Journal, 826, 1603.08005 Rodigas, T. J., Malhotra, R., & Hinz, P. M. 2014, The Astrophysical Journal, 780, 65 1310.5048 Teyssandier, J., Naoz, S., Lizarraga, I., & Rasio, F. A. 2013, ApJ, 779, 166, Verrier, P. E., & Evans, N. W. 2009, MNRAS, 394, 1721, 0812.4528 Wright, J. T. et al. 2011, PASP, 123, 412, 1012.5676 Zanardi, M., de Elía, G. C., Di Sisto, R. P., Naoz, S., Li, G., Guilera, O. M., & Brunini, A. 2017, ArXiv e-prints, 1701.03865 Ziglin, S. L. 1975, Soviet Astronomy Letters, 1, 194
1512.00483
2
1512
2016-02-08T19:04:59
Zodiacal Exoplanets In Time (ZEIT) I: A Neptune-sized planet orbiting an M4.5 dwarf in the Hyades Star Cluster
[ "astro-ph.EP" ]
Studying the properties of young planetary systems can shed light on how the dynamics and structure of planets evolve during their most formative years. Recent K2 observations of nearby young clusters (10-800 Myr) have enabled the discovery of such planetary systems. Here we report the discovery of a Neptune-sized planet transiting an M4.5 dwarf (K2-25) in the Hyades cluster (650-800 Myr). The lightcurve shows a strong periodic signal at 1.88 days, which we attribute to spot coverage and rotation. We confirm the planet host is a member of the Hyades by measuring the radial velocity of the system with the high-resolution near-infrared spectrograph IGRINS. This enables us to calculate a distance based on EPIC 210490365's kinematics and membership to the Hyades, which in turn provides a stellar radius and mass to 5-10%, better than what is currently possible for most Kepler M dwarfs (12-20%). We use the derived stellar density as a prior on fitting the K2 transit photometry, which provides weak constraints on eccentricity. Utilizing a combination of adaptive optics imaging and high-resolution spectra we rule out the possibility that the signal is due to a bound or background eclipsing binary, confirming the transits' planetary origin. EPIC 210490365b has a radius ($3.43^{+0.95}_{-0.31}$R$_{E}$) much larger than older Kepler planets with similar orbital periods (3.484 days) and host-star masses (0.29$M_{\odot}$). This suggests that close-in planets lose some of their atmospheres past the first few hundred Myr. Additional transiting planets around the Hyades, Pleiades, and Praesepe clusters from K2 will help confirm if this planet is atypical or representative of other close-in planets of similar age.
astro-ph.EP
astro-ph
Received November 23, 2015; Accepted December 17, 2015 Preprint typeset using LATEX style emulateapj v. 5/2/11 A NEPTUNE-SIZED PLANET ORBITING AN M4.5 DWARF IN THE HYADES STAR CLUSTER ZODIACAL EXOPLANETS IN TIME (ZEIT). I. Andrew W. Mann,(cid:63),1,2 Eric Gaidos,3,4 Gregory N. Mace,2 Marshall C. Johnson,2 Brendan P. Bowler,2,5,6 Daryll LaCourse,7 Thomas L. Jacobs7, Andrew Vanderburg,8,9 Adam L. Kraus,2 Kyle F. Kaplan,2 Daniel T. Jaffe2 Received November 23, 2015; Accepted December 17, 2015 ABSTRACT Studying the properties of young planetary systems can shed light on how the dynamics and structure of planets evolve during their most formative years. Recent K2 observations of nearby young clusters (10-800 Myr) have facilitated the discovery of such planetary systems. Here we report the discovery of a Neptune-sized planet transiting an M4.5 dwarf (K2-25) in the Hyades cluster (650-800 Myr). The light curve shows a strong periodic signal at 1.88 days, which we attribute to spot coverage and rotation. We confirm that the planet host is a member of the Hyades by measuring the radial velocity of the system with the high-resolution near-infrared spectrograph IGRINS. This enables us to calculate a distance based on K2-25's kinematics and membership to the Hyades, which in turn provides a stellar radius and mass to (cid:39) 5 − 10%, better than what is currently possible for most Kepler M dwarfs (12-20%). We use the derived stellar density as a prior on fitting the K2 transit photometry, which provides weak constraints on eccentricity. Utilizing a combination of adaptive optics imaging and high-resolution spectra, we rule out the possibility that the signal is due to a bound or background eclipsing binary, confirming the transits' planetary origin. K2-25b has a radius (3.43+0.95−0.31 R⊕) much larger than older Kepler planets with similar orbital periods (3.485 days) and host-star masses (0.29M(cid:12)). This suggests that close-in planets lose some of their atmospheres past the first few hundred million years. Additional transiting planets around the Hyades, Pleiades, and Praesepe clusters from K2 will help confirm whether this planet is atypical or representative of other close-in planets of similar age. Subject headings: stars: fundamental parameters - stars: individual (K2-25) - stars: late-type - stars: low-mass – stars: planetary systems - stars: statistics 1. INTRODUCTION Planets and their host stars evolve with time, and the first few hundred million years are thought to be the most formative. Final assembly of rocky terrestrial planets is predicted to occur in 10-100 Myr (Morbidelli et al. 2012). Regular accretion of residual planetesimals would continue to influence physical and chemical conditions on these planets (Hashimoto et al. 2007; Abramov & Mojzsis 2009). More rapid rotation and magnetic activity drive elevated X-ray and ultraviolet emission and coronal mass ejections from the host star, potentially eroding the primordial atmospheres of close-in planets on this timescale (Lammer et al. 2014). M dwarfs play a disproportionately large role in the discovery of Earth-size planets, particularly those planets (cid:63) [email protected] 1 Hubble Fellow 2 Department of Astronomy, The University of Texas at Austin, Austin, TX 78712, USA 3 Visiting Astronomer at the Infrared Telescope Facility, which is operated by the University of Hawaii under contract NNH14CK55B with the National Aeronautics and Space Admin- istration 4 Department of Geology & Geophysics, University of Hawaii at Manoa, Honolulu, HI 96822, USA 5 California Institute of Technology, 1200 E. California Blvd., Pasadena, CA 91125, USA. 6 McDonald Fellow 7 Amateur Astronomer 8 Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA 9 NSF Graduate Research Fellow with theoretical equilibrium temperatures permissive of liquid water (Muirhead et al. 2012; Dressing & Charbon- neau 2013; Gaidos 2013; Mann et al. 2013b; Dressing & Charbonneau 2015). This is because M dwarfs have smaller radii than solar-type stars, permitting the de- tection of smaller planets, and much lower luminosities, such that close-in and detectable planets will also be cooler. The dynamical and structural evolution of these systems may be qualitatively different than that of Sun- like stars; M dwarf stars take much longer (∼ 108 yr) to drop onto the main sequence and remain active much longer than their solar-type counterparts (Ansdell et al. 2015; West et al. 2015). These characteristics could have consequences for the climatic states and atmospheric evo- lution of M dwarf planets (Luger & Barnes 2015). While thousands of exoplanets have been discovered, most by the NASA Kepler transiting planet survey mis- sion (Borucki et al. 2010), the vast majority of these orbit old ((cid:29) 1 Gyr) stars. However, the repurposed Kepler spacecraft (K2, Howell et al. 2014) has observed 10-800 Myr-old clusters (i.e., Upper Scorpius, Pleiades, Hyades, and Praesepe). The TESS (Ricker 2014) and PLATO (Rauer et al. 2014) missions will also observe many stars in both young clusters and nearby young moving groups. These surveys will populate the temporal dimension of exoplanet parameter space, allowing us to statistically deduce how their orbits, masses, and atmospheres change with time. Until catalogs from the Gaia mission become avail- able, stars in well-studied clusters are usually easier to 2 Mann et al. characterize than their counterparts in the field. Precise abundances can be determined from the Sun-like members and can be applied to late-type stars where abundance determinations are more complicated. The distances to most of the nearest young clusters are well established (van Leeuwen 2009; Melis et al. 2014). While one cannot use the exact cluster distance for individual members of large clusters like the Hyades that have members > 20 pc from the cluster center, it is still possible to derive an accu- rate (5-10%) "kinematic distance," i.e., the distance that yields Galactic kinematics (U V W ) consistent with the cluster or moving group (e.g., Röser et al. 2011; Malo et al. 2013). For M dwarfs the combination of distance, flux, and temperature can yield radius estimates accurate to 5-10% (Delfosse et al. 2000; Bayless & Orosz 2006; Mann et al. 2015). This is significantly better than the 13− 15% errors for most Kepler M dwarfs based on spectroscopy (e.g., Muirhead et al. 2014; Newton et al. 2015) and is less subject to systematic biases that have plagued stellar characterization of Kepler targets from photometry alone (e.g., Mann et al. 2012; Gaidos & Mann 2013; Gaidos et al. 2013). Perhaps more significant, the common age of all members can be established by applying different methods (e.g., asteroseismology, isochrone fitting, lithium depletion boundary) to the full set of stars. A handful of planets have already been discovered around stars in young open clusters, i.e., by the Doppler radial velocity (RV) method in the Hyades (650-800 Myr) and Praesepe (650-800 Myr) clusters (Sato et al. 2007; Quinn et al. 2012, 2014) and by Kepler in the ∼1 Gyr old NGC 6811 cluster (Meibom et al. 2013). Owing to the sensitivity of the Doppler method, the former sam- ple is limited to Jupiter-mass planets, which represent only a tiny fraction of the total planet population. The latter is limited to two Neptune-sized planets in a single, distant (≈ 1100 kpc, and therefore difficult to study) cluster. K2 is observing much closer clusters with a range of ages and has already proved precise enough to find close-in (P (cid:46) 20 days), Neptune-size and smaller planets (e.g., Crossfield et al. 2015; Foreman-Mackey et al. 2015; Petigura et al. 2015; Vanderburg et al. 2015) Here we present the K2 discovery and our validation and characterization of a Neptune-sized planet tran- siting a mid-M-type dwarf (EPIC 210490365, 2MASS J04130560+1514520, K2-25) in the Hyades cluster. This object was identified by visual inspection of the host star's K2 light curve shortly after public release. In Section 2 we describe our spectroscopic and imaging follow-up, as well as literature photometry of the host star and extrac- tion of the K2 light curve. We use these data and others from the literature in Section 3 to show that this is a true member of the Hyades cluster. Using the member- ship status of K2-25, we derive a kinematic distance to K2-25, which in turn we utilize in Section 4 to derive accurate stellar parameters of the star. Our fit to the transit light curve is described in Section 5. In Section 6 we combine our stellar and planetary parameters with our adaptive optics (AO) and high-resolution observations to confirm the planetary nature of this transit. In Section 7 we conclude with a brief summary and discussion of the tentative implications for this system, the key differences between this planet and those found by Kepler around old M dwarfs, and the need for additional follow-up. 2. ARCHIVAL AND FOLLOW-UP OBSERVATIONS 2.1. Archival Photometry/Imaging We compiled optical BV photometry from the eighth data release of the AAVSO All-Sky Photometric Survey (APASS; Henden et al. 2012), NIR JHKS photometry from The Two Micron All Sky Survey (2MASS; Skrutskie et al. 2006), griz photometry from the Sloan Digital Sky Survey (SDSS; Ahn et al. 2012), and W 1W 2W 3 infrared photometry from the Wide-field Infrared Survey Explorer (WISE; Wright et al. 2010). We retrieved proper motions for K2-25 from Röser et al. (2011), which combined PP- MXL (Roeser et al. 2010) and UCAC3 (Zacharias et al. 2010) proper motions. These basic data on K2-25 are given in Table 1. The SDSS images are sufficiently deep to detect faint background stars (r(cid:48) < 21) that might fall within the K2 aperture (provided they are (cid:29) 1(cid:48)(cid:48) from the star) and contaminate the light curve or generate a false positive (if they are eclipsing binaries). The DSS image complements this, by offering a view directly behind the star when the target was ∼ 7(cid:48)(cid:48) away owing to its large proper mo- tion between the epoch of the Palomar Observatory Sky Survey (POSS, 1953) and K2 (2015) observations. We utilized these images as part of our false-positive analysis described in Section 6. 2.2. K2 light curve K2-25 was observed by the repurposed Kepler satellite (K2, Howell et al. 2014) for (cid:39)71 days from 2015 February 8 to April 20. Telescope pointing for K2 is unstable due to the loss of two reaction wheels, so the telescope drifts slowly. When the roll angle deviates too far from the desired position the thrusters fire to correct the pointing. As the point-spread function of the star shifts during the roll and subsequent thrust the total measured flux from the star changes due to pixel-to-pixel sensitivity variations. The resulting systematic noise is on timescales matching the thrust and roll (∼6 hr). Fortunately, these deviations can be corrected for or at least mitigated (Vanderburg & Johnson 2014; Armstrong et al. 2015). We retrieved the light curve of K2-25 provided by Vanderburg & Johnson (2014), who generate light curves from K2 pixel data accounting for the nonuniform pixel response function by fitting out correlations between the flux levels and spacecraft pointing. We also downloaded the light curve provided by the Kepler and K2 Science Center to test our sensitivity to the corrections applied by Vanderburg & Johnson (2014). 2.3. Optical Spectrum We obtained an optical spectrum of K2-25 with the SuperNova Integral Field Spectrograph (SNIFS; Aldering et al. 2002; Lantz et al. 2004) on the University of Hawaii 2.2m telescope on Mauna Kea. SNIFS provides simulta- neous coverage from 3200 to 9700 Å by splitting the beam with a dichroic mirror onto blue (3200–5200 Å) and red (5100–9700 Å) spectrograph channels. Although the spec- tral resolution of SNIFS is only R (cid:39) 1000, the instrument provides excellent spectrophotometric precision (Mann et al. 2011; Buton et al. 2013). Three optical spectra were taken of the target in succession (each integrated for 900 s) under cloudy conditions. After reduction the A Transiting Planet in the Hyades 3 three spectra were stacked to yield a signal-to-noise ratio (S/N) of (cid:39)200 (per pixel) in the r band. Reduction of SNIFS data was split into two parts; the first part was done by the SuperNova Factory (SNF) pipeline, which performs basic reduction (e.g., bias and flat correction) and extraction of the 1-dimensional spec- trum. The second section, carried out by our pipeline. applies flux and telluric correction and places the star in its rest frame. Extensive details of the SNF pipeline can be found in Bacon et al. (2001) and Aldering et al. (2006); and details of our pipeline can be found in Gaidos et al. (2014) and Mann et al. (2015). 2.4. Near-Infrared Spectrum A near-infrared (NIR) spectrum was taken with the up- dated SpeX spectrograph (Rayner et al. 2003) mounted on the NASA Infrared Telescope Facility (IRTF) on Mauna Kea. SpeX observations were taken in the short cross- dispersed (SXD) mode using the 0.3×15(cid:48)(cid:48) slit, yielding simultaneous coverage from 0.8 to 2.4µm at a resolution of R (cid:39) 2000. The target was placed at two positions along the slit (A and B) and observed in an ABBA pattern in order to subsequently subtract the sky background. We took 6 exposures following this pattern for a total integration time of 717s, which, when stacked, provided a S/N per pixel of 120 in the H and K bands. SpeX spectra were extracted using the SpeXTool pack- age (Cushing et al. 2004), which performed flat-field correction, wavelength calibration, sky subtraction, and extraction of the one-dimensional spectrum. Multiple exposures were combined using the IDL routine xcomb- spec. To correct for telluric lines, we observed the A-type star HD 31295 immediately after the target observations and within 0.1 airmasses. A telluric correction spectrum was constructed from the A0V star and applied using the xtellcor package (Vacca et al. 2003). Separate orders were stacked using the xcombspec tool. Following the method outlined in Mann et al. (2015) we combined and absolutely flux-calibrated the opti- cal and NIR spectra using published photometry (Sec- tion 2.1) with the filter profiles and zero points provided in Fukugita et al. (1996)11 and Mann & von Braun (2015). We show the combined spectrum in Figure 1. 2.5. High Resolution NIR Spectra We observed K2-25 at 10 epochs spread over 36 days with the Immersion Grating Infrared Spectrometer (IGRINS; Park et al. 2014) on the 2.7m Harlan J. Smith telescope at McDonald Observatory. IGRINS provides si- multaneous H- and K-band (1.48-2.48µm) coverage with a resolving power of R (cid:39)45,000. Similar to the SpeX observations, the target was placed at two positions along the slit and observed in an ABBA pattern. At each epoch we took four exposures (one ABBA cycle), each 240-400s (depending on conditions). For each epoch we also observed an A0V star immediately before or after the observations of K2-25 to aid with telluric correction. IGRINS spectra were reduced using version 2.1 of the publicly available IGRINS pipeline package12 (Lee 2015). The IGRINS pipeline performed flat, bias, and dark cor- rections, as well as extraction of the one-dimensional 11 See http://classic.sdss.org/dr7/algorithms/fluxcal.html 12 https://github.com/igrins/plp spectrum of both the A0V standard and target. An ini- tial wavelength solution was derived using the ThAr lines, followed by a full wavelength solution derived from the sky lines. The resulting spectrum at each epoch has a S/N of 30− 60 per pixel in the center of the H and K bands. We used the A0V spectra to correct for telluric lines follow- ing the method outlined in Vacca et al. (2003). Spectra with uncorrected telluric lines were used for measuring radial velocities to improve the wavelength solution (see Section 3). A single high-S/N spectrum was constructed by stacking the ten exposures after shifting them to the same radial velocity. The final stacked spectrum has a S/N of ∼ 120 in the H-band, which we used to search for faint lines from an undetected companion (Section 6) and calculate v sin i∗ (Section 4). 2.6. Adaptive Optics Imaging We obtained natural guide star (NGS; Wizinowich et al. 2000) AO imaging of K2-25 with the facility imager, NIRC2, on Keck II atop Mauna Kea. Observations were taken with the K(cid:48) filter and the narrow camera. In this mode the pixel scale is 9.952 mas pix−1 (Yelda et al. 2010) and the field of view (FOV) for the 1024 × 1024 pixel array is 10.2(cid:48)(cid:48) × 10.2(cid:48)(cid:48). We took seven images, each with five co-adds, and each co-add integrating for 2 s. Basic reduction (dark, flat field, and bad pixel correction) was applied to each of the images, which were then registered and stacked to produce a single, deep image. No sources are visible in the fFOV other than K2-25 down to the resolution limit of the images ((cid:39) 0.07(cid:48)(cid:48)). From the re- duced and stacked images we constructed a contrast curve using the noise maps and the flux from the primary star following Bowler et al. (2015), which we show in Figure 2. Using the age of the 650 Myr for Hyades, with the KS magnitude and distance of K2-25(Section 4), and the models of Baraffe et al. (2015), we converted our AO image into a sensitivity map following the procedure from Bowler et al. (2015), which we show in Figure 2. The result does not significantly change for an age of 800 Myr. This provides a probability of detecting a star of a given separation and mass in the AO images and accounts for the chance of detecting a companion at a random place in its orbit. 3. HYADES MEMBERSHIP K2-25 was previously identified as a member of the Hyades in Röser et al. (2011) based on its proper motion and photometry. Using the subset of their candidates with parallaxes and a comparison of the density of the Hyades and field stars, Röser et al. (2011) estimate that field star contamination within 9 pc of the cluster core is negligibly small, and they find K2-25 to be only 3.5 pc from the core. Similarly, Douglas et al. (2014) calculate a 99% chance that K2-25 is a member of the Hyades. Furthermore, our high-resolution NIR spectra (Section 2.5), combined with the proper motion (Section 2.1), enable a precise determi- nation of this star's kinematics and hence unambiguously confirm K2-25's membership in the Hyades. We calculated a barycentric RV of K2-25 for each of the 10 IGRINS epochs following the method from Mace et al. (2016, in preparation), which takes advantage of the large spectral grasp and stability of IGRINS. The procedure was to; (1) split each of 42 orders into eight suborders and remove the two on each end where the 4 Mann et al. Figure 1. Combined and absolutely flux-calibrated optical and NIR spectrum of K2-25. The spectrum is shown in black. Photometry is shown in red, with the horizontal "error bars" indicating the width of the filter, and vertical errors representing combined measurement and zero point errors. Blue points indicate the corresponding synthetic fluxes from convolving the spectrum with the appropriate filter profile and multiplying by the zero point. Residuals are plotted in the bottom panel in units of standard deviations. The inset panel shows a zoom-in of the blue part of the spectrum including the major Balmer and calcium H & K emission lines. Figure 2. 7-σ contrast curve (left) and sensitivity map (right) for K2-25 constructed from our AO imaging. The registered and stacked AO image is shown as an inset on the left panel. The sensitivity map is a measure of the probability of detecting an object of a given mass and separation based on an age of 650 Myr and the distance to K2-25 (Section 4). See Bowler et al. (2015) for more details. S/N is lowest owing to the drop in the blaze function, (2) cross-correlate the telluric spectrum to find offsets in the wavelength solution due to temperature changes and instrument flexures, (3) invert and cross-correlate the remaining telluric-corrected suborders against a tem- plate (or series of templates) to determine the offset in pixels, (4) convert the pixel offsets into RVs using the instrument dispersion solution, (5) adopt the median of the RV measurement after removing > 4σ outliers, and (6) repeat steps 2-4 for every available template with a known RV, allowing the RV of each template to adjust according to their literature uncertainties. In total we compared each observation with 153 M2–M6 templates with known RVs. For each epoch the assigned RV and error is the mean and standard deviation of the mean for these 153 measurements. We use the mean of the RVs from the 10 epochs as the system RV. Nominally the statistical error on the final barycentric RV is < 50 m s−1, but it is limited by the 153 m s−1 error in the zero point velocity derived from the templates. Our final barycentric velocity is 38.64 ± 0.15km s−1 and is reported in Table 1. We estimated a photometric distance to the star by comparing the spectral energy distribution of K2-25 (Sec- tion 2.1) with that of template M dwarfs with known distances from Mann et al. (2015). For each star in Mann et al. (2015) we calculated the reduced χ2 (χ2 ν) 0.20.40.60.81.0Flux (10−14 erg cm−2 s−1 Å−1)SpectrumPhotometrySynthetic Photometry0.380.400.420.440.460.48Wavelength (µm) 0.51.01.52.0Wavelength (µm)−303Residual (σ)0.11.010.0Separation (")8642∆K′ (mag)10−1∆RA (")−101∆Dec (") NE 0.10.91101001000 0.010.101.00 0.00.20.40.60.81.0Detection ProbabilitySemimajor Axis (AU)Companion Mass (MSun) A Transiting Planet in the Hyades 5 between the template and K2-25's BV griz JHKs pho- tometry with the distance of K2-25 as a free parameter. We calculated the best-fit distance for K2-25 from the me- dian and standard deviation of the 10 stars with χ2 ν < 3. This yielded a distance of 44 ± 7 pc, assuming that the star is not an unresolved binary. We used the coordi- nates, proper motion, RV, and photometric distance to calculate Galactic position XY Z and motion U V W with corresponding errors, which we report in Table 1. We calculated the probability that K2-25 is member of the Hyades following the Bayesian framework from Rizzuto et al. (2011) and Malo et al. (2013). For sim- plicity we only considered the possibility that this star is part of the Hyades or the field. We adopted XY Z and U V W for the Hyades from van Leeuwen (2009) and the XY Z and U V W values for field stars from Malo et al. (2013). For the prior Malo et al. (2013) give equal weights to membership in each of the memberships considered; however, this is overly optimistic for our case as there are far more field stars than members of the Hyades. Instead we followed Rizzuto et al. (2011) and selected a prior equal to the ratio of the number of stars in the Hyades to the number of field stars in the same region of the sky. We identified all stars within 10◦ of the target in APASS that land within 5σ of the color-magnitude diagram of stars in the Hyades drawn from the Röser et al. (2011) and Goldman et al. (2013) catalogs. We conservatively assume that all of these stars not in Röser et al. (2011) and Goldman et al. (2013) are field stars. Plugging the XY Z and U V W values and errors for our target and the Hyades, along with our prior estimated above, into Equation 8 of Malo et al. (2013) gives a 99.99% chance that K2-25 is a member of the Hyades as opposed to a field star. This is relatively insensitive to our choice of prior; we would have to decrease the prior by more than an order of magnitude to drop the membership probability below 99.7% (3σ). Our analysis does not consider the chance that K2-25 is a member of another moving group, stream, or cluster (only the Hyades and the field). However, the XY ZU V W values of K2-25 are not consistent with any other known moving group or cluster, so we consider this possibility to have negligible probability. K2-25 lands only (cid:39) 3.4 pc from the core of the Hyades and therefore is probably still gravitationally bound to the cluster. We show the Galactic position (XY Z) for members of the Hyades taken from Röser et al. (2011) and Goldman et al. (2013) compared to that of K2-25 in Figure 3. 4. STELLAR PARAMETERS Spectral type: We calculated TiO, CaH, and VO molec- ular indices following the definitions from Reid et al. (1995) and Lépine et al. (2013). We then derived a deci- mal spectral type using the empirical relations between the strength of these molecular indices and the spectral type from Lépine et al. (2013). This gives a final spectral type of M4.5 with an internal error of ≤0.3 subtypes. Hα: The optical spectrum shows noticeable Balmer series and Calcium H & K emission (Figure 1), as expected for a mid-M dwarf in the Hyades (West et al. 2008). We calculated an Hα equivalent width of −3.1 Å (negative to denote emission), following the continuum and feature definitions from Lépine et al. (2013). The resulting value is consistent with other Hα measurements of stars in Hyades and Praesepe clusters (Douglas et al. 2014). Metallicity: Our SpeX spectrum enables us to derive the host star's metallicity from atomic indices in the H and K bands (e.g., Terrien et al. 2012; Rojas-Ayala et al. 2012; Mann et al. 2013a; Newton et al. 2014). However, the commonly used Na lines at 2.2µm could be affected by stellar activity and lower surface gravity (Deen 2013), although this effect should be smaller for the Ca and K lines in the H-band (Terrien et al. 2012). The H-band indices from Terrien et al. (2012) and Mann et al. (2013a) give consistent values ([Fe/H] = 0.15±0.10, 0.17±0.08) and both are in agreement with literature values for the cluster (0.12–0.18, Paulson et al. 2003; Brandt & Huang 2015; Dutra-Ferreira et al. 2016). For our analysis we adopted [Fe/H]=0.15±0.03, which captures both our mea- surements and those from the literature within 1σ and is far more precise than can be done on an individual M dwarf. Distance: Because K2-25 is a member of the Hyades, we can derive a kinematic distance more precise than the photometric distance we used in Section 3 (e.g., Kraus et al. 2014). To this end, we recalculated U V W but allowing distance to float between 1 and 100 pc. We then found the distance that gives U V W values consistent with the established value of the cluster (van Leeuwen 2009). We allowed for a variation of 1.2 km/s in the cluster value due to dispersion from internal kinematics (Palmer et al. 2014). Accounting for this, as well as errors in the proper motion and RV, gave a distance of 45.7±3.3 pc. Luminosity: We first calculated the bolometric flux by integrating over our combined and absolutely flux- calibrated optical and NIR spectrum (Section 2). This gives a bolometric flux of 1.301 ± 0.015 × 10−10 erg s−1 cm−2. Errors on the bolometric flux account for random and correlated (e.g., slope errors in the flux calibration) errors in the combined spectrum, as well as measurement and zero point errors in the photome- try, as detailed in Mann et al. (2015). When combined with the kinematic distance, this yields a luminosity of 8.4 ± 1.4 × 10−3L(cid:12). Effective temperature: We derived an effective tem- perature (Teff) from our optical spectrum following the procedure from Mann et al. (2013b). To briefly summa- rize, Mann et al. (2013b) compare optical spectra of M dwarfs with BT-SETTL CIFIST models13 (Allard et al. 2011). By masking out regions of the spectrum that are poorly reproduced by the models, Mann et al. (2013b) re- produced the Teff scale from long-baseline interferometry (Boyajian et al. 2012) to 60 K, which we adopted as the error on our measurement. This procedure yields a Teff of 3180±60 K. Mass, radius and density: To estimate the stellar mass, radius, and density, we used the Mann et al. (2015) rela- tions between absolute Ks-band magnitude (MKs) and metallicity and stellar radius/mass. The radius relation was calibrated using angular diameter measurements from long-baseline optical interferometry (Boyajian et al. 2012). The mass relation is slightly model dependent, but re- produces the empirical mass-luminosity relation from Delfosse et al. (2000), and in combination the mass and 13 https://phoenix.ens-lyon.fr/Grids/BT-Settl/CIFIST2011 6 Mann et al. Figure 3. Galactic position (XY Z) of Hyades members (circles) and K2-25 (five-point star). All points are colored according to their MKS magnitude. The member list, as well as distances and coordinates used in computing XY Z, was taken from Röser et al. (2011) and Goldman et al. (2013). The background contamination rate (false members) is expected to be low (<7.5%) near the core (< 18 pc), but significantly higher (> 30%) for more distant stars. radius relations reproduce the mass-radius relation from low-mass eclipsing binaries (Feiden & Chaboyer 2012; Mann et al. 2015) within errors. Accounting for errors in the distance, [Fe/H], Ks magnitude, and scatter in the relations, we derived a radius of 0.295 ± 0.020 R(cid:12) and a mass of 0.294 ± 0.021 M(cid:12). For the stellar density we also considered that errors are correlated, i.e., if the distance is greater, both the mass and radius increase together. Accounting for this via Monte Carlo simulation, we find a density of 11.3+1.7−1.5ρ(cid:12). Mann et al. (2015) relations are primarily based on stars older than 1 Gyr, while K2-25 is relatively young (650–800 Myr). However, at this age K2-25 is expected to be on the main sequence, and the Mann et al. (2015) relations include some stars with similar activity levels (as measured by Hα) as K2-25. As a check we also derived the stellar radius using the Stefan–Boltzmann law and our above luminosity and Teff. This yields a radius of 0.301± 0.032 R(cid:12), < 1σ from our estimate above. We note that this method is not completely independent of the MKs-R∗ relation, as they both rely on the same distance and Ks measurement. Rotation period: The light curve from Vanderburg & Johnson (2014) shows ∼ 1% amplitude periodic variation due to rotation and spot coverage. We calculated the autocorrelation function of the light curve and found a peak at P = 1.881 ± 0.021 days, with a harmonic at 0.940±0.005 days; we consider the former the rotation period of the star. Errors on the rotation period were determined by fitting the autocorrelation function around the peak as a Gaussian and do not consider (potential) sources of systematic error such as aliasing, differential rotation, or short-lived spots (e.g., Aigrain et al. 2015). However, periodicity does not change over the 71-day ob- serving period from K2 despite changes in the amplitude. Further, this rotation period is consistent with stars of similar mass and age in the Hyades and Praesepe clusters (Figure 4). v sin i∗14: We estimated the level of rotational broad- Figure 4. Rotation period as a function of stellar mass for stars in the Hyades and Praesepe clusters taken from Agüeros et al. (2011), Scholz et al. (2011), and Delorme et al. (2011). Praesepe targets are similar in age to that of the Hyades (both 650-800 Myr) and have more rotation period measurements at low masses in the literature. K2-25 is shown in red. Errors on mass measurements are typically (cid:39)10%. ening in K2-25 from our stacked IGRINS spectrum. We compared our IGRINS spectrum with a BT-SETTL model with a Teff, log g, and [M/H] of 3200, 5.0, and 0.0 (roughly consistent with our calculations above), which we artifi- cially broaden with the IDL code lsf_rotate (Gray 1992; Hubeny & Lanz 2011). Each of the IGRINS orders is fit separately, excluding those for which the S/N is too low (< 20). We normalized each order and the appropriate region of the model with a 150-pixel (much larger than a given feature) running median after masking out regions of strong (> 30%) telluric absorption. To account for in- strumental broadening, we simultaneously fit the telluric lines in each order, which we extract from the A0V star spectrum (see Section 2.5). We assume that instrumental broadening is Gaussian and that the telluric lines have negligible intrinsic width. The instrumental broadening typically has a full-width half-max of 0.3-0.5 Å, consistent 14 i∗ is used for the stellar sky-projected inclination to distinguish it from i, the inclination of the planet's orbit. A Transiting Planet in the Hyades 7 Table 1 Parameters of K2-25 Parameter Value Source with the resolution of the spectrograph. We applied the broadening derived for each order to the model. We then fit the model to the spectrum, letting v sin i∗ float from 0 to 50 km s−1, calculating χ2 at each step. We adopted the median and standard deviation of the v sin i∗ mea- surements across all fit orders as the final measurement and error, which was 7.8 ± 0.5 km s−1. Sky projected inclination: The combination of our v sin i∗, rotation period, and stellar radius enables us to calculate the (sky-projected) rotational inclination (i∗) of K2-25. We first calculate the equatorial velocity (Veq): Veq = 2πR∗ Prot , (1) where Prot is the stellar rotation period. This yields a velocity of 7.9± 0.5km s−1. We follow the formalism from Morton & Winn (2014) to convert v sin i∗ and Veq to a posterior in cos(i∗), which handles regions of the posterior where v sin i∗> Veq (which is physically impossible). The resulting posterior gives a lower limit on stellar inclination of i∗ > 52◦ at 99.7% confidence (3σ), and i∗ > 72◦ at 68.3% (1σ). All derived stellar parameters for K2-25 are listed in Table 1. 5. K2 LIGHT CURVE ANALYSIS For our analysis we relied on the light curve from Van- derburg & Johnson (2014), which used a 3 × 3 pixel aperture centered on K2-25. We also downloaded and re- peated our analysis using the Pre-search Data Condition- ing (PDCSAP; Stumpe et al. 2012) light curve released by the Kepler and K2 Science Center. Both light curves are shown in Figure 5. Our final results were consistent using either light curve, although the residuals of our transit analysis were smaller when using the reprocessed Vanderburg & Johnson (2014) light curve, so we report values from that analysis. K2-25b was initially identified by eye in the PDCSAP light curve, as the periodic 1% dips (Figure 5) can be seen even through the strong rotational variability. 5.1. Search for additional transiting planets We performed a systematic search for other transit sig- nals in the K2 light curve of K2-25. We started with the light curve from Vanderburg & Johnson (2014) and applied several additional corrections to filter stellar ro- tational variability and flaring. First, a power spectrum was generated using the Lomb-Scargle algorithm (Scargle 1981) and significant, isolated peaks were identified with false-positive probability (FAP) < 0.01. These frequencies were filtered from the light curve and a running median with a 1-day window was also removed (the window is much larger than the expected duration of any transit). A robust standard deviation was calculated using the algorithm of Tukey (1977) and > 3σ positive excursions were replaced with median values. This data set was then searched for periodic transit-like signals using the box-least-squares algorithm from Kovács et al. (2002). A second-order trend in the power spectrum was removed, and signals with FAP < 0.01 were identified by calculat- ing the Signal Detection Efficiency (Eqn. 6 in Kovács et al. (2002)) and evaluating the significance using the cumulative Gaussian with the parameters set to the val- ues found by Kovács et al. (2002) for pure noise. These EPIC EPIC Röser et al. (2011) Röser et al. (2011) APASS APASS SDSS SDSS SDSS SDSS 2MASS 2MASS 2MASS W ISE W ISE W ISE Astrometry 04:13:05.61 +15:14:52.00 120.6 ± 3.3 −21.1 ± 3.2 Photometry 17.760 ± 0.289 15.881 ± 0.030 16.730 ± 0.010 15.235 ± 0.031 13.760 ± 0.010 12.820 ± 0.010 11.303 ± 0.021 10.732 ± 0.020 10.444 ± 0.019 8.443 ± 0.023 8.424 ± 0.021 8.322 ± 0.055 1.88 ± 0.02 38.64 ± 0.15 −42.4 ± 1.2 −18.4 ± 3.2 −1.8 ± 2.4 −39.8 ± 6.3 +1.2 ± 0.2 −18.8 ± 3.0 45.7 ± 3.3 −3.1 ± 0.1 7.8 ± 0.5 M4.5 ± 0.3 0.15 ± 0.03 3180 ± 60 0.294 ± 0.021 0.295 ± 0.020 > 72 Derived Properties α R.A. (hh:mm:ss) δ Dec. (dd:mm:ss) µα (mas yr−1) µδ (mas yr−1) B (mag) V (mag) g (mag) r (mag) i (mag) z (mag) J (mag) H (mag) Ks (mag) W 1 (mag) W 2 (mag) W 3 (mag) This paper This paper This paper This paper This paper This paper This paper This paper This papera This paper This paper This paper This paper This paperb This paper This paper This paper (8.4 ± 1.4) × 10−3 This paper This paper Rotation period (days) Barycentric RV (km s−1) U (km s−1) V (km s−1) W (km s−1) X (pc) Y (pc) Z (pc) Distance (pc) EW (Hα) (Å) v sin i∗ (km s−1) i∗ (degrees) Spectral type [Fe/H] Teff (K) M∗ (M(cid:12)) R∗ (R(cid:12)) L∗ (L(cid:12)) ρ∗ (ρ(cid:12)) a The distanced derived from cluster membership and kinematics of K2- 25b (see Section 4). b This is the weighted mean of our own measurements from SpeX and literature measurements for the Hyades cluster (see Section 4). candidate signals were then analyzed in detail and their S/N calculated. Besides the 3.485-day signal investigated here, no other signals with periods between 0.3 and 20 days were detected. 11.3+1.7−1.5 5.2. Simultaneous-fit K2 light curve The Vanderburg & Johnson (2014) pipeline is optimized for producing light curves of slowly rotating stars, and K2- 25's rapid high amplitude photometric variability resulted in uncorrected systematic effects in the original Vander- burg & Johnson (2014) light curve. Once we identified the transit we reprocessed the K2 light curve using the same procedure as Becker et al. (2015), simultaneously fitting for the stellar activity signal, the K2 flat field, and the transits of K2-25b using a Levenberg-Marquardt minimization algorithm (Markwardt 2009). We modeled the stellar variability as a spline with breakpoints every 0.2 days, the K2 flat field as splines in K2 image centroid position with breakpoints roughly every 0.4 arcseconds, and the transits with a Mandel & Agol (2002) model, while taking into account the 30-minute long-cadence in- tegration time. This processing effectively removed the systematics from the K2 pointing jitter and resulted in 8 Mann et al. improved photometric precision. We show both light curves as well as the PDCSAP light curve in Figure 5. 5.3. Transit Fitting We fit the flattened light curve with a Monte Carlo Markov Chain (MCMC) by utilizing the emcee Python module (Foreman-Mackey et al. 2013) and the batman tool (Kreidberg 2015), which utilizes the Mandel & Agol (2002) transit model. We oversampled and binned the model to the Kepler cadence to handle light-curve distor- tion from long integration times (see Kipping 2010, for a discussion of this issue). We sampled the planet-to-star radius ratio (RP /R∗), impact parameter (b), orbital pe- riod (P ), epoch of the first transit midpoint (T0), two √ parameters that describe the eccentricity and argument of periastron ( e cos(ω)), bulk stellar den- sity (ρ∗), and two limb-darkening parameters (q1 and q2). We assumed a quadratic limb-darkening law and use the triangular sampling method of Kipping (2013) in order to uniformly sample the physically allowed region of parameter space. MCMC chains were run using 100 walkers, each with 200,000 steps after a burn-in phase of 10,000 steps. e sin(ω) and √ √ The transit duration is uniquely determined from the other fitted transit parameters following the formulae from Seager & Mallén-Ornelas (2003), but using ρ∗ as the free parameter rather than transit duration enables us to apply a prior on ρ∗ using our values derived in Section 4. This in turn let us constrain e and ω, which are generally hard to measure owing to their minimal impact on the observed light curve (similar methods are used in Dawson & Johnson 2012; Kipping et al. 2012). As explained in Van Eylen & Albrecht (2015), directly exploring e and ω biases the eccentricity to higher values owing to the √ cutoff at zero, while sampling uniformly in e sin(ω) and e cos(ω) between -1 and 1 is unbiased and still provides uniform sampling in e and ω (see Lucy & Sweeney 1971; Ford 2006; Anderson et al. 2011; Eastman et al. 2013, for more detailed discussions of this issue). We applied a prior drawn from the model-derived limb- darkening coefficients from Claret & Bloemen (2011) as- suming a quadratic limb-darkening law. We interpolate our stellar parameters (log g, Teff, [Fe/H]; see Section 4) onto the Claret & Bloemen (2011) grid of limb-darkening coefficients from the PHOENIX models, accounting for errors from the finite grid spacing, errors in stellar pa- rameters, and variations from the method used to derive the coefficient (Least-Square or Flux Conservation). This yielded priors of 0.45 ± 0.1 and 0.35 ± 0.1 for the linear and quadratic limb-darkening coefficients, respectively, which we propagated to q1 and q2 using the formulae in Kipping (2013). We uniformly sampled the impact parameter over -1.2 to +1.2 (to allow for grazing transits), orbital period over 0 to 100 days, and mid-time of the first transit over ±1.5 days (about half the period) from the value identified in our L-S analysis (Section 5.1). √ We fit the light curve twice: once with e sin(ω) and √ e cos(ω) fixed at 0 and no prior on ρ∗, and once with e cos(ω) limited to −1 to +1 and under e sin(ω) and uniform priors, and with a prior on ρ∗ using our values derived in Section 4. We report the results of both transit fits in Table 2. For each parameter we report the median value with the 'errors' as the 84.1 and 15.9 percentile val- ues (corresponding to 1σ for Gaussian distributions). The √ √ model light curve with the best-fit parameters (highest likelihood) is shown for the latter fit in Figure 6 alongside the K2 data. We also show posteriors and correlations for a subset of parameters in Figure 7. Both fits have a noticeable tail in the RP /R∗ distri- bution owing to poor sampling of the transit duration and a degeneracy with impact parameter. The transit duration is only slightly longer than the integration time, so it is difficult to completely rule out partially grazing (b > 0.9) orbital solutions. The MCMC is able to achieve a reasonable fit to the data with large RP /R∗ values by simultaneously adjusting the impact parameter and tran- sit duration. However, such solutions are also disfavored compared to solutions with a lower impact parameter and shorter transit duration (although less so for the fit with the prior on ρ∗). This issue could be further complicated if model limb-darkening parameters turn out to be systematically erroneous for cool stars, although fits of high-quality light curves suggest that the model parameters are at least roughly correct (e.g., Kreidberg et al. 2014; Swift et al. 2015). We note that follow-up observations from the ground at higher cadence could significantly mitigate this degeneracy by resolving the transit duration, particularly NIR observations, where limb-darkening has a smaller impact. The two different MCMC fits give consistent values for all fit values. The main difference is that the fit with the ρ∗ prior favors a higher impact parameter and hence has power at large radii. The ρ∗-prior fit is consistent with e = 0, although a modest value (e (cid:39) 0.3) is slightly favored. Again, better data to resolve the transit duration would help significantly here, as e is highly degenerate with b and τ when using a prior on ρ∗. We adopt the second fit for our planet parameters as the additional constraints on ρ∗ should provide a more accurate picture of the true planet parameters. Further, K2-25b has a tidal circularization timescale of ∼ 1 Gyr (Goldreich & Soter 1966) and hence probably has not lost its initial eccentricity. Combined with our stellar radius from Section 4 the latter fit yields a planet radius of 3.43+0.95−0.31 R⊕. 6. FALSE POSITIVE ANALYSIS 6.1. Background Star We calculated a posterior probability that an unrelated background star could be responsible for the transit signal, i.e., as an eclipsing binary (EB). We followed the proce- dure described in detail in Gaidos et al. (2016), which is summarized here. The Bayesian probability was calcu- lated with a prior based on a model of the background stellar population and a likelihood calculated from ob- servational constraints, i.e., (1) a background star has to be bright enough to produce a transit signal given a maximum possible eclipse depth of 50%, (2) the density of the star must be consistent with the transit duration; (3) the star is not visible in the 1953 POSS red and blue image in which K2-25, owing to its proper motion over 6 decades, is displaced by about 7(cid:48)(cid:48) (Figure 8); and (4) the star is not visible in our NIRC2 AO imaging (Section 2.6). We used a model stellar population calculated by TRI- LEGAL version 1.6 (Vanhollebeke et al. 2009). We cal- culated the stellar population at the position of K2-25 to Kp = 22 equivalent over a field of 10 square degrees (to reduce counting noise). The faint limit is several mag- A Transiting Planet in the Hyades 9 Figure 5. light curve of K2-25 taken by the K2 spacecraft. The top panel is the PDCSAP flux provided by the Kepler and K2 Science Center. The middle panel is the Vanderburg & Johnson (2014) light curve with simultaneous fitting of the K2 flat field, while the bottom panel shows the Vanderburg & Johnson (2014) light curve with simultaneous fitting of the flat field and stellar variability (see Section 5.2). All light curves have been normalized to 1 and the time zeroed to the start of the K2 observations. Some data points (< 1%) that we attribute to flares are off the top of the panels. The shape and features seen in the light curves are insensitive to choice of aperture size, as the region is free of significant contaminating flux from background stars (see Figure 8). Table 2 Transit Fit Parameters Parameter Fit 1a Fit 2a (Preferred) 3.484552+0.000031 −0.000037 0.1065+0.0286 −0.0065 57062.57935+0.00049 −0.00024 0.60+0.29−0.42 0.79+0.09−0.17 88.3+1.2−0.7 0.27+0.16−0.21 62+44−39 3.43+0.95−0.31 R⊕ 3.484552+0.000036 −0.000044 0.1028+0.0080 −0.0037 57062.57935+0.00049 −0.00024 0.35+0.36−0.25 0.74+0.06−0.04 89.5+0.4−0.9 0 (fixed) 0 (fixed) 3.31+0.34−0.25 R⊕ √ Period (days) RP /R∗ T0 (BJDb-2400000) Impact parameter Durationc (hours) Inclinationc (degrees) Eccentricity ω (degrees) d (R⊕) RP a Fit 1 is done with e and ω fixed at 0 and a uniform prior on ρ∗, while fit ecos(ω) limited to −1 to +1 under uniform 2 is done with priors and with a prior on ρ∗ from our analysis in Section 4. b BJD is given in Barycentric Dynamical Time (TBD) format. c For both fits stellar density and impact parameter are the fitted pa- rameters (instead of transit duration and orbital inclination). We report the duration and inclination derived from the other fit parameters here for convenience. d Planet radius is derived using our stellar radius from Section 4. √ esin(ω) and 10 Mann et al. Table 3 Relative Radial Velocities JD-2,400,000 RV (m s−1)a σRV (m s−1) 161 57288.92269 159 57289.89553 161 57293.85944 158 57295.86940 158 57319.84005 159 57320.83534 159 57321.83510 159 57322.88659 159 57323.79641 57324.76798 159 Note. - aRadial velocities are quoted with respect to the mean. For the system (absolute) velocity see Table 1. 300 -42 -238 192 -109 -172 232 114 -85 -191 derived from our IGRINS spectra (see Section 2.5, and Section 3). In theory the relative RVs from our IGRINS data should be more precise than the absolute velocities, but this does not account for astrophysical jitter common to young stars. Empirical measurements of Hyades-age Sun-like stars in the optical suggest variations of 50m s−1 from activity (Paulson et al. 2004; Hillenbrand et al. 2015). Based on the amplitude of spot variations we see in the K2 light curve and the v sin i∗ measured from our IGRINS spectrum, the spot-induced RV jitter should be ∼120m s−1 in the Kepler bandpass, but observations of T Tauri stars suggest that this is smaller by a factor of two to three at the wavelength of our IGRINS data (2.2µm; Crockett et al. 2012). Another complication is the stability of IGRINS for precision RVs. The use of telluric lines to correct the wavelength scale should reduce instrumental variability to 5-20m s−1 (e.g., Figueira et al. 2010; Blake et al. 2010), but correlated errors may persist and radial velocity stability should be tested empirically. So instead of the expected errors of 50m s−1 derived from the scatter in RV variations between orders, we adopt the larger error (typically 150- 160m s−1) derived from scatter from using different RV templates. We report the resulting velocities and errors for each epoch in Table 3. We show the RVs in Figure 9 phased against both the planet's orbital and the star's rotation period. Although the velocity scatter is too large to measure the mass of K2-25b, we set a limit to the mass on the transiting body by assuming that it is the source of all variation in the RV. To this end we fit the RV data with a simple least-squares minimization (Markwardt 2009), fixing P to the value from the transit fit, e to 0, and limiting the eclipse time to a 3σ range given from the transit fits (but that are otherwise unconstrained) and the inclination to > 80◦. We found that the largest mass that is still consistent with the RV values at 5σ is 3 Jupiter masses, ruling out a grazing EB. Although this alone does not rule out the possibility of a grazing Saturn- or Jupiter-sized planet, such a solution is strongly disfavored by our transit fit (Section 5). Figure 6. Phase-folded light curve of K2-25's transit (black points). The red line shows the best-fit (highest likelihood) model from our MCMC fit (Section 5). The bottom panel shows the fit residuals. nitudes deeper than the faintest EB diluted that could possibly produce the transit signal after dilution by K2-25 (Kp = 14.53 mag). The false positive probability (FPP) was calculated by the method of Monte Carlo; model stars were placed at random locations in a circular field 16(cid:48)(cid:48) (4 Kepler pixels) in radius centered on K2-25. Stars were discarded or retained based on the contrast criterion ∆Kp < −2.5 log(δ/R), where δ = 0.012 is the transit depth and R is a pixel response function interpolated from the Kepler Instrument Handbook supplement values for the appropriate detector channel (13). We also dis- carded stars brighter than Kp = 19 mag based on the DSS POSS 1 image and stars with a contrast in the infrared K band brighter than the 7σ detection limit in our Keck 2-NIR2-AO imaging. We weighted each remaining star with the probability that a transit of an object with the observed orbital period (3.485 d) would have the observed duration ((cid:39)47 minutes) if placed around the background star, versus being placed around K2-25. This calculation requires a probability distribution for e, and we adopted a Rayleigh distribution with mean e = 0.1, appropriate for short-period binaries. We then summed up the number of stars in the 16" circle and divided by the ratio of the circular solid angle to 10 square degrees. We find a FPP of ≈ 4.5× 10−8, a consequence of the depth of the transit, brightness of K2-25, and the low background star counts at this moderate Galactic latitude. 6.2. Eclipsing Companion to K2-25 Because of the short orbital period, small stellar size, and long observing cadence, a V-shaped transit is expected for a low-eccentricity orbit, even if the eclipse/transit is partial. However, the V shape we see also leaves open the possibility that the observed transit is actually a grazing eclipse from a stellar companion with a 3.485-day orbital period. This is reflected in the tail in the posterior of the RP /R∗ distribution. Although the posterior never passes into values consistent with stars, the fits assume that the eclipsing body is not luminous, and hence this alone does not rule out the possibility that the transit is due to a star. Further, although no secondary eclipse is seen and the even and odd transits have consistent depths, there are regions of inclination, e, and ω space where there would be no secondary eclipse. Instead, we can rule out this possibility using the RVs A Transiting Planet in the Hyades 11 Figure 7. Posteriors from our MCMC fit. Median values for each parameter are marked with red dashed lines. Gray scaling corresponds to 1σ, 2σ, and 3σ (from light to dark). The left panel shows the fit with eccentricity and argument of periastron fixed at 0 and no prior on density, while the right panel shows a fit with ecos(ω) allowed to float and a prior on stellar density from our analysis in Section 4. For the latter fit we show the eccentricity posterior, but since eccentricity is fixed in the former, we instead show the transit duration posterior. esin(ω) and √ √ Figure 9. RVs from IGRINS phased to the planet's orbital period (top, 3.485 days) and star's rotation period (bottom, 1.88 days). Two phases are shown, but repeat measurements are shown in gray. In the top panel the expected signal from a Neptune-mass, Jupiter- mass, and 3×Jupiter mass planet (with e = 0 and P = 3.485 days) are shown as teal, blue, and red lines, respectively. In the bottom panel we show the predicted jitter from the ∼1.5% spot-induced variations in the K2 light curve and the v sin i∗ measurement of ∼8 km s−1, although this is expected to be smaller in the K band. 6.3. Eclipsing Binary Companion to K2-25 We considered the possibility that the transit signal is due to a EB bound to K2-25 (we consider background EBs in Sec. 6.1). To be missed by our AO images any compan- ion must be within (cid:39)10 AU or be too faint (∆Kp > 4 mag) to reproduce the transit depth (Figure 2). Kraus et al. (in review) show that planet formation is suppressed by > 85% inward of 60 AU. We tested the binary compan- ion hypothesis with our Doppler data because the stellar components of a 3.5 d-period system would exhibit signif- icant RV variation over the baseline of our observations, producing multiple, variable sets of lines in the IGRINS spectra. We see no second set of lines in any of our NIR spectra, nor is there a second peak in the cross-correlation function. We determine the significance of this nondetec- tion by simulating ternary (K2-25 plus EB companion) systems, adding companions to K2-25 drawn randomly from the observed binary mass ratio from Duchêne & Kraus (2013), but with the companion as a 3.485-day Figure 8. Archive images of K2-25 from DSS (top) and SDSS (bottom). Both images have the same scale, and a 20(cid:48)(cid:48) bar is shown in blue on the DSS image. The current location of the object is shown as a red circle in both images. Because the DSS image is from 1953, the target was ∼7(cid:48)(cid:48) from its current position, revealing potential unresolved background stars. Boxes corresponding to K2 apertures (5×5 and 3×3 K2 pixels) are shown in green and teal in the SDSS image. We use the smaller (9 pixel) aperture to cut out the faint background star. Because of the large K2 PSF, two background stars visible in SDSS still contaminate the smaller aperture, but both of these stars are too faint to reproduce the 1% transit depth. 12 Mann et al. EB with a mass ratio randomly drawn from a uniform distribution. We added BT-SETTL synthetic spectra to our stacked spectrum of K2-25 and then searched for a second set of lines or a second peak in the cross-correlation function. We found that 99.8% of our simulated systems either are insufficiently bright to produce the observed transit depth, would be seen in the AO image, or produce a second peak in the cross-correlation function of one or more of our IGRINS observations. This simulation is also likely an overestimate, as it only simulates triple systems. 7. SUMMARY AND DISCUSSION As members of the Hyades have common and well- established ages, metallicities, and distances, their other properties can be constrained more precisely, allowing more rigorous studies of how planets evolve structurally and dynamically with time. M dwarfs are especially inter- esting targets for exoplanet searches in clusters because their small size facilitates the discovery of smaller planets. To this end we have begun the KELP project to find transiting exoplanets around low-mass cluster members monitored by K2. K2-25 represents the first discovery in our search, as well as the first discovery of a transiting planet in the Hyades. We obtained moderate-resolution optical and NIR spec- tra of K2-25, used to measure its temperature and lumi- nosity. Multiple epochs of high-resolution, NIR spectra with the IGRINS spectrograph enabled us to confirm K2-25's membership in the Hyades cluster and rule out the possibility that the transit signal is due to a grazing eclipsing binary. Because the star is a member of the Hyades, we were able to derive a kinematic distance and correspondingly precise stellar parameters. We took ad- vantage of these tight constraints on stellar density to improve the transit fit by applying a prior on ρ∗, enabling weak constraints on e and better constraints on b. Using RVs from the IGRINS spectra, AO imaging, and the lack of background stars seen in the 1953 POSS data (during which K2-25 was in a different location), we are able to confirm the planetary nature of the transit signal. Owing to the excellent precision of Kepler, M dwarf KOIs usually have small errors on RP /R∗; instead, er- rors on planetary radii are usually dominated by errors in the stellar radius (e.g., Muirhead et al. 2012, 2015). This has motivated a plethora of follow-up programs and efforts to improve methods to constrain M dwarf radii (e.g., Rojas-Ayala et al. 2012; Zhou et al. 2014; Neves et al. 2014; Newton et al. 2015; Hartman et al. 2015). K2 M dwarfs are statistically closer than Kepler M dwarfs, but still have poorly constrained stellar radii (e.g., Cross- field et al. 2015; Montet et al. 2015). Like stars with known parallaxes (e.g., MEarth; Dittmann et al. 2014), because we know the (kinematic) distance of K2-25 its other parameters are more precisely established (e.g., a 6- 7% error in R∗). However, because the 30-minute cadence of the Kepler photometry is comparable to the transit duration ((cid:39)45 minutes), even K2's high precision light curves leave us with an RP /R∗ value that is only well constrained in one direction (1σ error is +26%, −6%). Thus, K2-25b represents a case where the limiting factor in planet parameters is the light curve, and not the stellar parameters. High-cadence photometry from the ground could sig- nificantly improve the planet parameters. While K2-25 is quite faint in the optical (V = 15.9) by moving to the near-infrared (z = 12.8, K = 10.4) a modest sized (≥ 2 m) telescope could achieve the requisite precision (mmag) with a cadence of < 5 minutes. This is sufficient to resolve out the transit duration and likely rule out or confirm the high impact parameter that is present in our transit fit posteriors (Section 5). The activity lifetime of an M dwarf is 1.2 ± 0.4 Gyr for a spectral type M2, rising to 4.5+0.5−1.0 Gyr at M4 and 7.0 ± 0.5 Gyr at M5 (West et al. 2008). The lack of Hα emission seen in the Kepler M dwarf planet hosts (Mann et al. 2013b) of similar spectral type suggests that they are all > 1 Gyr, significantly older than K2-25. A comparison to other transiting planets gives us some insight into how planets evolve beyond the Hyades' age (650-800 Myr). In Figure 10 we show the radius K2-25b as a function of host star mass and planet irradiance compared to transiting planets found by Kepler and MEarth with orbital periods < 100 days and host star masses M∗ < 0.5. Parameters for Kepler planets and stars are drawn from Gaidos et al. (2015), which are derived in a manner consistent with our own, and parameters for GJ 1214b and GJ 1132b are taken from Anglada-Escudé et al. (2013) and Berta-Thompson et al. (2015), respectively. It is clear that K2-25b is unusually large for its host star mass and orbital period. Only one Kepler planet orbiting an M∗ < 0.5 star has a larger radius than K2-25b (KOI 4928.01), but its host star is ∼50% more massive than K2-25. GJ 1214b (Charbonneau et al. 2009) is the closest match, as the planet is notably smaller, but it also orbits a significantly less massive star. GJ 581 has a similar mass ((cid:39) 0.3M(cid:12)) and harbors a nontransiting giant planet, so such large planets are possible. However, it is clear that large planets around very low mass stars are rare. Both Kepler and MEarth searched (cid:39) 1000 M∗ < 0.5M(cid:12) stars (Berta et al. 2013; Gaidos et al. 2015) and each found just ∼one large planet, while K2-25b was found after searching just ∼70 candidate members of the Hyades with M∗ < 0.5M(cid:12). Detection biases probably cannot explain the lack of such planets in the Kepler sample; close-in, Neptune-sized transiting planets should be obvious in the Kepler data, unless they were flagged as eclipsing binaries owing to a short transit duration. One possible explanation for the large size of K2-25b is that it is evolving under the influence of the environ- ment of its young host star. M dwarfs, like most stars, pass through a juvenile phase of elevated UV and X-ray emission, flares, and coronal mass ejections. Models pre- dict this activity to be capable of removing any weakly bound, primordial hydrogen/helium envelopes from rocky planets on close-in orbits (Lammer et al. 2014) and are supported by the ejection of a large cloud of neutral hy- drogen around the Neptune-mass GJ 436b (Ehrenreich et al. 2015). K2-25b could represent an early or inter- mittent phase of planetary evolution where the loss of a distended atmosphere has not yet reached completion. Detection and characterization of additional planets in young clusters are needed to test such scenarios. While there are only ∼ 100 M dwarfs in the Hyades observed in Campaign 4, many more will be observed in Campaign 13 and even more M dwarfs in Praesepe and Pleiades. Al- though Praesepe and Pleiades are more distant and hence A Transiting Planet in the Hyades 13 Figure 10. Planet size as a function of host star mass (left) and planet irradiance (right) for K2-25 (red) compared to transiting planets discovered by Kepler (black) and from the ground by MEarth (blue). Only planets orbiting stars with host masses < 0.5M(cid:12) and orbital periods < 100 days are included. their M dwarf members are much fainter, the Pleiades is significantly younger (∼110 Myr. Dahm 2015), and plan- ets as large as K2-25b should be quite obvious around even relatively faint M dwarfs. A.W.M. was supported through Hubble Fellowship grant 51364 awarded by the Space Telescope Science Institute, which is operated by the Association of Univer- sities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. This research was supported by NASA grant NNX11AC33G to E.G. E.G. was also sup- ported by a International Visitor grant from the Swiss National Science Foundation. We thank Bandit for his useful discussions and encouragement during the writing of this manuscript. This work used the Immersion Grating Infrared Spec- trograph (IGRINS), which was developed under a col- laboration between the University of Texas at Austin and the Korea Astronomy and Space Science Institute (KASI) with the financial support of the US National Science Foundation under grant ASTR1229522, of the University of Texas at Austin, and of the Korean GMT Project of KASI. The IGRINS pipeline package PLP was developed by Dr. Jae-Joon Lee at Korea Astronomy and Space Science Institute and Professor Soojong Pak's team at Kyung Hee University. SNIFS on the UH 2.2-m tele- scope is part of the Nearby Supernova Factory project, a scientific collaboration among the Centre de Recherche Astronomique de Lyon, Institut de Physique Nucléaire de Lyon, Laboratoire de Physique Nucléaire et des Hautes Energies, Lawrence Berkeley National Laboratory, Yale University, University of Bonn, Max Planck Institute for Astrophysics, Tsinghua Center for Astrophysics, and the Centre de Physique des Particules de Marseille. Some of the data presented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. Support for MAST for non-HST data is provided by the NASA Office of Space Science via grant NNX09AF08G and by other grants and contracts. This research was made possible through the use of the AAVSO Photometric The deep transit of K2-25b makes it a useful target for atmospheric characterization. Follow-up of similar-sized objects, such as GJ 1214b, has mostly suggested hazy, featureless atmospheres (e.g., Berta et al. 2012; Knutson et al. 2014; Kreidberg et al. 2014). However, no one has examined the atmosphere of such a young Neptune-sized planet, and it is not known wheather the atmosphere will show features that are no longer present in older counterparts. A high volatile content in the transmission spectrum also might explain K2-25b's unusual size. A measurement of K2-25b's mass would yield con- straints on the planet density, possibly providing in- sight on the unusual size of this planet. Assuming that the planet is Neptune-mass on a near-circular orbit, the Doppler amplitude is expected to be ∼ 15 m s−1. The scatter in our RV measurements is significantly larger than this (> 200 m s−1) even after accounting for the expected measurement error ((cid:39)50m s−1). We adopted a more conservative measurement of the error ((cid:39)150m s−1) to account for jitter common to young stars and the untested long-term stability of IGRINS at the 50m s−1 level. The RV scatter may also be due to a nontransiting planet. Additional RV measurements could resolve this question. Even if the source of the noise is astrophysical, it might be possible to remove signals that do not follow a Keplerian trend consistent with the orbital period of K2-25b. The faint optical magnitude and spot-induced jitter are limitations, but K2-25 would be an ideal target for monitoring by NIR spectrographs like CARMENES (Quirrenbach et al. 2012), the Infrared Doppler instrument (Kotani et al. 2014), SPIRou (Artigau et al. 2014), and the Habitable-zone Planet Finder (Mahadevan et al. 2010). 14 Mann et al. All-Sky Survey (APASS), funded by the Robert Martin Ayers Sciences Fund. The authors acknowledge the Texas Advanced Computing Center (TACC) at The University of Texas at Austin for providing grid resources that have contributed to the research results reported within this paper. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain. Facilities: IRTF (SpeX), UH:2.2m (SNIFS), Kepler, Keck:II (NIRC2), Smith (IGRINS) REFERENCES Abramov, O., & Mojzsis, S. J. 2009, Nature, 459, 419 Agüeros, M. A., Covey, K. R., Lemonias, J. J., et al. 2011, ApJ, PASP, 125, 306 Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2012, ApJS, ApJ, 806, 215 740, 110 203, 21 Berta-Thompson, Z. K., Irwin, J., Charbonneau, D., et al. 2015, Knutson, H. A., Benneke, B., Deming, D., & Homeier, D. 2014, Delfosse, X., Forveille, T., Ségransan, D., et al. 2000, A&A, 364, Delorme, P., Collier Cameron, A., Hebb, L., et al. 2011, MNRAS, Dittmann, J. A., Irwin, J. M., Charbonneau, D., & Berta-Thompson, Z. K. 2014, ApJ, 784, 156 Douglas, S. T., Agüeros, M. A., Covey, K. R., et al. 2014, ApJ, 217 413, 2218 795, 161 Dressing, C. D., & Charbonneau, D. 2013, ApJ, 767, 95 -. 2015, ApJ, 807, 45 Duchêne, G., & Kraus, A. 2013, ARA&A, 51, 269 Dutra-Ferreira, L., Pasquini, L., Smiljanic, R., Porto de Mello, G. F., & Steffen, M. 2016, A&A, 585, A75 Eastman, J., Gaudi, B. S., & Agol, E. 2013, PASP, 125, 83 Ehrenreich, D., Bourrier, V., Wheatley, P. J., et al. 2015, Nature, Feiden, G. A., & Chaboyer, B. 2012, ApJ, 757, 42 Figueira, P., Pepe, F., Lovis, C., & Mayor, M. 2010, A&A, 515, 522, 459 A106 Ford, E. B. 2006, ApJ, 642, 505 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, Foreman-Mackey, D., Montet, B. T., Hogg, D. W., et al. 2015, Fukugita, M., Ichikawa, T., Gunn, J. E., et al. 1996, AJ, 111, 1748 Gaidos, E. 2013, ApJ, 770, 90 Gaidos, E., Fischer, D. A., Mann, A. W., & Howard, A. W. 2013, ApJ, 771, 18 Gaidos, E., & Mann, A. W. 2013, ApJ, 762, 41 Gaidos, E., Mann, A. W., & Ansdell, M. 2016, ApJ, 817, 50 Gaidos, E., Mann, A. W., Kraus, A. L., & Ireland, M. 2015, ArXiv e-prints, arXiv:1512.04437 [astro-ph.EP] Gaidos, E., Mann, A. W., Lépine, S., et al. 2014, MNRAS, 443, Goldman, B., Röser, S., Schilbach, E., et al. 2013, A&A, 559, A43 Goldreich, P., & Soter, S. 1966, Icarus, 5, 375 Gray, D. F. 1992, The observation and analysis of stellar 2561 photospheres. Hartman, J. D., Bayliss, D., Brahm, R., et al. 2015, AJ, 149, 166 Hashimoto, G. L., Abe, Y., & Sugita, S. 2007, Journal of Geophysical Research (Planets), 112, 5010 Henden, A. A., Levine, S. E., Terrell, D., Smith, T. C., & Welch, D. 2012, Journal of the American Association of Variable Star Observers (JAAVSO), 40, 430 Hillenbrand, L., Isaacson, H., Marcy, G., et al. 2015, in Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, Vol. 18, 18th Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, ed. G. T. van Belle & H. C. Harris, 759 Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126, 398 Hubeny, I., & Lanz, T. 2011, Synspec: General Spectrum Synthesis Program, Astrophysics Source Code Library, ascl:1109.022 Kipping, D. M. 2010, MNRAS, 408, 1758 -. 2013, MNRAS, 435, 2152 Kipping, D. M., Dunn, W. R., Jasinski, J. M., & Manthri, V. P. 2012, MNRAS, 421, 1166 Nature, 505, 66 Kotani, T., Tamura, M., Suto, H., et al. 2014, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9147, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 14 Kovács, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369 Kraus, A., Ireland, M., Huber, D., Mann, A. W., & Dupuy, T. in Kraus, A. L., Shkolnik, E. L., Allers, K. N., & Liu, M. C. 2014, AJ, Kreidberg, L. 2015, PASP, 127, 1161 Kreidberg, L., Bean, J. L., Désert, J.-M., et al. 2014, Nature, 505, 69 3225 Lammer, H., Stökl, A., Erkaev, N. V., et al. 2014, MNRAS, 439, Lantz, B., Aldering, G., Antilogus, P., et al. 2004, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 5249, Optical Design and Engineering, ed. L. Mazuray, P. J. Rogers, & R. Wartmann, 146 Lee, J.-J. 2015, plp: Version 2.0 Lépine, S., Hilton, E. J., Mann, A. W., et al. 2013, AJ, 145, 102 Aigrain, S., Llama, J., Ceillier, T., et al. 2015, MNRAS, 450, 3211 Aldering, G., Adam, G., Antilogus, P., et al. 2002, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4836, Survey and Other Telescope Technologies and Discoveries, ed. J. A. Tyson & S. Wolff, 61 Aldering, G., Antilogus, P., Bailey, S., et al. 2006, ApJ, 650, 510 Allard, F., Homeier, D., & Freytag, B. 2011, in Astronomical Society of the Pacific Conference Series, Vol. 448, 16th Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, ed. C. Johns-Krull, M. K. Browning, & A. A. West, 91 Anderson, D. R., Smith, A. M. S., Lanotte, A. A., et al. 2011, MNRAS, 416, 2108 Anglada-Escudé, G., Rojas-Ayala, B., Boss, A. P., Weinberger, A. J., & Lloyd, J. P. 2013, A&A, 551, A48 Ansdell, M., Gaidos, E., Mann, A. W., et al. 2015, ApJ, 798, 41 Armstrong, D. J., Santerne, A., Veras, D., et al. 2015, A&A, 582, Artigau, É., Kouach, D., Donati, J.-F., et al. 2014, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9147, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Bacon, R., Copin, Y., Monnet, G., et al. 2001, MNRAS, 326, 23 Baraffe, I., Homeier, D., Allard, F., & Chabrier, G. 2015, A&A, 577, A42 Bayless, A. J., & Orosz, J. A. 2006, ApJ, 651, 1155 Becker, J. C., Vanderburg, A., Adams, F. C., Rappaport, S. A., & Schwengeler, H. M. 2015, ApJ, 812, L18 Berta, Z. K., Irwin, J., & Charbonneau, D. 2013, ApJ, 775, 91 Berta, Z. K., Charbonneau, D., Désert, J.-M., et al. 2012, ApJ, A33 747, 35 Nature, 527, 204 ApJS, 216, 7 757, 112 Blake, C. H., Charbonneau, D., & White, R. J. 2010, ApJ, 723, 684 Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977 Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Tamura, M. 2015, Boyajian, T. S., von Braun, K., van Belle, G., et al. 2012, ApJ, Claret, A., & Bloemen, S. 2011, A&A, 529, A75 Crockett, C. J., Mahmud, N. I., Prato, L., et al. 2012, ApJ, 761, Crossfield, I. J. M., Petigura, E., Schlieder, J. E., et al. 2015, ApJ, Cushing, M. C., Vacca, W. D., & Rayner, J. T. 2004, PASP, 116, 891 164 362 804, 10 Dahm, S. E. 2015, ApJ, 813, 108 Dawson, R. I., & Johnson, J. A. 2012, ApJ, 756, 122 Deen, C. P. 2013, AJ, 146, 51 Brandt, T. D., & Huang, C. X. 2015, ApJ, 807, 24 Buton, C., Copin, Y., Aldering, G., et al. 2013, A&A, 549, A8 Charbonneau, D., Berta, Z. K., Irwin, J., et al. 2009, Nature, 462, review, ApJ 147, 146 Lucy, L. B., & Sweeney, M. A. 1971, AJ, 76, 544 Luger, R., & Barnes, R. 2015, Astrobiology, 15, 119 Mahadevan, S., Ramsey, L., Wright, J., et al. 2010, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 7735, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Malo, L., Doyon, R., Lafrenière, D., et al. 2013, ApJ, 762, 88 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Mann, A. W., Brewer, J. M., Gaidos, E., Lépine, S., & Hilton, E. J. 2013a, AJ, 145, 52 Mann, A. W., Feiden, G. A., Gaidos, E., Boyajian, T., & von Braun, K. 2015, ApJ, 804, 64 Mann, A. W., Gaidos, E., & Aldering, G. 2011, PASP, 123, 1273 Mann, A. W., Gaidos, E., & Ansdell, M. 2013b, ApJ, 779, 188 Mann, A. W., Gaidos, E., Lépine, S., & Hilton, E. J. 2012, ApJ, 753, 90 Mann, A. W., & von Braun, K. 2015, PASP, 127, 102 Markwardt, C. B. 2009, in Astronomical Society of the Pacific Conference Series, Vol. 411, Astronomical Data Analysis Software and Systems XVIII, ed. D. A. Bohlender, D. Durand, & P. Dowler, 251 Meibom, S., Torres, G., Fressin, F., et al. 2013, Nature, 499, 55 Melis, C., Reid, M. J., Mioduszewski, A. J., Stauffer, J. R., & Bower, G. C. 2014, Science, 345, 1029 Montet, B. T., Morton, T. D., Foreman-Mackey, D., et al. 2015, ApJ, 809, 25 Morbidelli, A., Lunine, J. I., O'Brien, D. P., Raymond, S. N., & Walsh, K. J. 2012, Annual Review of Earth and Planetary Sciences, 40, 251 Morton, T. D., & Winn, J. N. 2014, ApJ, 796, 47 Muirhead, P. S., Johnson, J. A., Apps, K., et al. 2012, ApJ, 747, 144 801, 18 Muirhead, P. S., Becker, J., Feiden, G. A., et al. 2014, ApJS, 213, 5 Muirhead, P. S., Mann, A. W., Vanderburg, A., et al. 2015, ApJ, Neves, V., Bonfils, X., Santos, N. C., et al. 2014, A&A, 568, A121 Newton, E. R., Charbonneau, D., Irwin, J., et al. 2014, AJ, 147, 20 Newton, E. R., Charbonneau, D., Irwin, J., & Mann, A. W. 2015, Palmer, M., Arenou, F., Luri, X., & Masana, E. 2014, A&A, 564, ApJ, 800, 85 A49 Park, C., Jaffe, D. T., Yuk, I.-S., et al. 2014, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9147, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 1 Paulson, D. B., Cochran, W. D., & Hatzes, A. P. 2004, AJ, 127, Paulson, D. B., Sneden, C., & Cochran, W. D. 2003, AJ, 125, 3185 Petigura, E. A., Schlieder, J. E., Crossfield, I. J. M., et al. 2015, 3579 26 L38 389 785 315 Quinn, S. N., White, R. J., Latham, D. W., et al. 2012, ApJ, 756, 2831 ApJ, 811, 102 L33 -. 2014, ApJ, 787, 27 A Transiting Planet in the Hyades 15 Quirrenbach, A., Amado, P. J., Seifert, W., et al. 2012, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8446, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Rauer, H., Catala, C., Aerts, C., et al. 2014, Experimental Rayner, J. T., Toomey, D. W., Onaka, P. M., et al. 2003, PASP, Astronomy, 38, 249 115, 362 Reid, I. N., Hawley, S. L., & Gizis, J. E. 1995, AJ, 110, 1838 Ricker, G. R. 2014, Journal of the American Association of Variable Star Observers (JAAVSO), 42, 234 Rizzuto, A. C., Ireland, M. J., & Robertson, J. G. 2011, MNRAS, 416, 3108 Roeser, S., Demleitner, M., & Schilbach, E. 2010, AJ, 139, 2440 Rojas-Ayala, B., Covey, K. R., Muirhead, P. S., & Lloyd, J. P. 2012, ApJ, 748, 93 Röser, S., Schilbach, E., Piskunov, A. E., Kharchenko, N. V., & Scholz, R.-D. 2011, A&A, 531, A92 Sato, B., Izumiura, H., Toyota, E., et al. 2007, ApJ, 661, 527 Scargle, J. D. 1981, ApJS, 45, 1 Scholz, A., Irwin, J., Bouvier, J., et al. 2011, MNRAS, 413, 2595 Seager, S., & Mallén-Ornelas, G. 2003, ApJ, 585, 1038 Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163 124, 985 Stumpe, M. C., Smith, J. C., Van Cleve, J. E., et al. 2012, PASP, Swift, J. J., Montet, B. T., Vanderburg, A., et al. 2015, ApJS, 218, Terrien, R. C., Mahadevan, S., Bender, C. F., et al. 2012, ApJ, 747, Tukey, J. W. 1977, Exploratory data analysis Vacca, W. D., Cushing, M. C., & Rayner, J. T. 2003, PASP, 115, Van Eylen, V., & Albrecht, S. 2015, ApJ, 808, 126 van Leeuwen, F. 2009, A&A, 497, 209 Vanderburg, A., & Johnson, J. A. 2014, PASP, 126, 948 Vanderburg, A., Montet, B. T., Johnson, J. A., et al. 2015, ApJ, Vanhollebeke, E., Groenewegen, M. A. T., & Girardi, L. 2009, 800, 59 A&A, 498, 95 West, A. A., Hawley, S. L., Bochanski, J. J., et al. 2008, AJ, 135, West, A. A., Weisenburger, K. L., Irwin, J., et al. 2015, ApJ, 812, 3 Wizinowich, P., Acton, D. S., Shelton, C., et al. 2000, PASP, 112, Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, AJ, 140, 1868 Yelda, S., Lu, J. R., Ghez, A. M., et al. 2010, ApJ, 725, 331 Zacharias, N., Finch, C., Girard, T., et al. 2010, AJ, 139, 2184 Zhou, G., Bayliss, D., Hartman, J. D., et al. 2014, MNRAS, 437,
1906.01582
2
1906
2019-06-07T16:36:19
Dust Production of Comet 21P/Giacobini-Zinner Throughout its 2018 Apparition
[ "astro-ph.EP" ]
We present the results of a long term telescopic observation campaign of comet 21P/Giacobini-Zinner, parent body of the Draconid meteor shower, spanning $\sim 240$ days during its 2018 apparition. Determinations of comet 21P's dust production rate through the $Af\rho$ parameter derived from these images show that the comet had a highly asymmetric dust production rate that peaked $\sim 10-30$ days before perihelion, when the comet was at a heliocentric distance of $\sim 1.02 - 1.08 \thinspace \mathrm{AU}$. The single highest $Af\rho$ measurement occurred on 2018-Aug-14 (27 days before perihelion), and had a measured value of $Af\rho = 1594 \thinspace \mathrm{cm}$. The comet's $Af\rho$ profile is well described by a double-exponential model that rises rapidly during ingress and declines even more rapidly during its egress. These results are fully consistent with observations of comet 21P's dust and gas production rates during past apparitions, and suggest that the double-exponential model we have derived provides a reasonable and stable approximation for the comet's activity over the past thirty to forty years.
astro-ph.EP
astro-ph
DRAFT VERSION JUNE 10, 2019 Typeset using LATEX default style in AASTeX62 Dust Production of Comet 21P/Giacobini-Zinner Throughout its 2018 Apparition STEVEN EHLERT,1 NATALIE MOTICSKA,2 AND AURIANE EGAL3, 4, 5 1Qualis Corporation, Jacobs Space Exploration Group, NASA Meteoroid Environment Office Marshall Space Flight Center, Huntsville, AL 35812, USA 2Physical Sciences Department, Embry-Riddle Aeronautical University Daytona Beach, FL, USA 3Department of Physics and Astronomy, The University of Western Ontario London, Ontario N6A 3K7, Canada 4Centre for Planetary Science and Exploration, The University of Western Ontario London, Ontario N6A 5B8, Canada 5IMCCE, Observatoire de Paris, PSL Research University, CNRS Sorbonne Universit´es UPMC Univ. Paris 06, Univ. Lille, France (Received January 1, 2018; Revised January 7, 2018; Accepted June 10, 2019) Submitted to AJ ABSTRACT We present the results of a long term telescopic observation campaign of comet 21P/Giacobini-Zinner, parent body of the Draconid meteor shower, spanning ∼ 240 days during its 2018 apparition. Determinations of comet 21P's dust production rate through the Af ρ parameter derived from these images show that the comet had a highly asymmetric dust production rate that peaked ∼ 10 − 30 days before perihelion, when the comet was at a heliocentric distance of ∼ 1.02− 1.08 AU. The single highest Af ρ measurement occurred on 2018-Aug-14 (27 days before perihelion), and had a measured value of Af ρ = 1594 cm. The comet's Af ρ profile is well described by a double-exponential model that rises rapidly during ingress and declines even more rapidly during its egress. These results are fully consistent with observations of comet 21P's dust and gas production rates during past apparitions, and suggest that the double-exponential model we have derived provides a reasonable and stable approximation for the comet's activity over the past thirty to forty years. Keywords: comets: individual Comet 21P/Giacobini-Zinner -- meteoroids -- techniques: photometric 1. INTRODUCTION Of all of the major meteor showers, the October Draconids (009 DRA) remains one of the most difficult to model and forecast. The typical activity of the Draconids is usually comparable to the background sporadic rate for visual observers (with Zenithal Hourly Rates of ∼ 1−2 per hour, (Kronk 2014)), but strong outbursts have been observed in individual years. While some of these outbursts were anticipated, others were entirely unexpected by observers and were not associated with favorable configurations between Earth and the Draconid parent body. Further complicating matters, some of these outbursts were associated only with the lowest mass meteoroids. These outbursts were not observed visually1, but were only apparent in radio or radar instrument surveys2. Even for the Draconid outbursts that were anticipated in advance, models have frequently failed to predict the time of peak shower activity, the shower duration, and/or the peak shower flux (Kronk 2014). Many of the challenges associated with modeling the Draconid meteor shower are associated with the peculiar behavior of the shower's parent body. The parent body of the Draconids, comet 21P/Giacobini-Zinner (hereafter comet 21P) currently has an orbit with a period of 6.54 years, a perihelion distance of 1.0128 AU, and a semi-major axis of 3.50 AU (Laboratory 2019). Its Corresponding author: S. Ehlert [email protected] ∼ 10−5 kg. 1 By "visual" observations, we mean either individual eyewitnesses or optical instruments, which are generally sensitive to meteoroids more massive than 2 Radar instruments are typically sensitive to masses as low as ∼ 10−7 kg. 2 EHLERT, MOTICSKA, AND EGAL Tisserand parameter with respect to Jupiter is 2.466, classifying it as a Jupiter Family Comet (JFC). Comet 21P has therefore been subject to significant perturbations by Jupiter. Its orbit has been significantly altered over the past century. The most dramatic of these perturbations occurred during its 1959-1965 orbit, over which large changes to its non-gravitational forces were observed (Yeomans 1971). Photometric and spectroscopic measurements of the dust and gas production of comet 21P during previous apparitions have shown that this comet has an atypical composition (A'Hearn et al. 1995; Lara et al. 2003). In particular, comet 21P is considered the prototypical "carbon-depleted comet", as its C2 abundances have been measured to be significantly lower than the majority of other comets surveyed in A'Hearn et al. (1995). Many observers have also provided evidence that gas and dust production for comet 21P reaches its maximum value approximately 1 month prior to its arrival at perihelion. For example, the observations of Schleicher et al. (1987) suggested that gas and dust production dropped by a factor of ∼ 2 between pre-perihelion and post- perihelion, a result that was supported by both UV spectroscopy (McFadden et al. 1987) and broad-band photometry (Lara et al. 2003). For earlier apparitions, observations suggested that dust production may have peaked after perihelion (Sekanina 1985). The analysis of Pittichov´a et al. (2008) showed that during its 2005 apparition, the dust production rate at heliocentric distances of ∼ 1.7 − 3.0 AU (post-perihelion) decreased rapidly during egress, with a logarithmic slope of ∼ −2.0. When observing the comet at similar heliocentric distances pre-perihelion during its 2011 apparition (Blaauw et al. 2014) showed that comet 21P's dust production rate increased even more rapidly during ingress, with a logarithmic slope of ∼ −4.5. The majority of these results suggest that comet 21P's dust production rises rapidly during ingress, peaks approximately one month before perihelion, and decreases rapidly thereafter. However, discrepancies in coverage, instrumental setups, and measurement techniques between different authors limit this interpretation. In order to better model the activity of comet 21P and subsequently improve the forecasts for the Draconid meteor shower, NASA's Meteoroid Environment Office (MEO) has undertaken a long-term observation campaign of comet 21P. Unlike observa- tions taken during previous apparitions, we monitored the comet every few nights for nearly three months both before and after perihelion. We utilized telescopes located at multiple locations across the Earth, covering both the Northern and Southern Hemi- spheres, in order to minimize gaps in between observations. In this work, we present the results of this observation campaign. This is a companion paper to Egal et al. 2019 (submitted, hereafter Egal19) that utilized these observational data in conjunction with improved meteoroid stream models in order to forecast future Draconid meteor showers. Egal19 focuses almost exclusively on modeling the ejection of dust grains from the comet, the dynamics of the meteoroid stream, and comparing the modelled activity of the Draconid shower at Earth each year to historical observations. This paper instead will describe the observational campaign and measurements of comet 21P's dust production. This paper is structured as follows: Section 2 discusses the particulars of the observations including scheduling, telescope hardware, and filters. Section 3 describes the methods by which dust production rates of comet 21P were determined, and Section 4 presents the measurements themselves. We discuss the implications of these results in the context of meteor shower forecasting and observations of previous apparitions in Section 5. Unless otherwise noted, all dates and times correspond to UTC. 2. OBSERVATIONAL CONSIDERATIONS Observations of comet 21P commenced on 2018-05-06, and the final images were taken on 2018-12-30. Figure 1 illustrates the ephemeris of the comet during this time period. From this figure, two crucial aspects of the observation planning are noted. First, the comet was positioned in both the Northern and Southern hemispheres of the sky during this campaign. Telescopes in both hemispheres were necessarily utilized, and had to accommodate for the comet's non-sidereal motion. Second, the comet was usually located coincident with the Galactic Plane during these observations, meaning that it was usually positioned among crowded star fields. Images of comet 21P were taken using telescopes publicly available through the iTelescope network3. In particular, we imaged the comet using the following telescopes: T11 in Mayhill, New Mexico; T7 in Nerpio, Spain; and T30/T31 at Siding Spring Observatory in Australia. Table 1 describes the general properties of each of these telescopes. 2.1. Telescopes Utilized All images of comet 21P during this observation campaign were taken using the Johnson-Cousins RC-band filter (hereafter simply the R-band) in place, the standard filter for measuring Af ρ. This filter has an effective wavelength of 6407 A and an 2.2. Filters and Timing 3 https://www.itelescope.net DUST PRODUCTION OF COMET 21P DURING ITS 2018 APPARITION 3 Figure 1. The ephemeris of comet 21P during this observation campaign overlaid on an all-sky image of the Milky Way. The green curve denotes the Celestial Equator (0◦ declination) in Galactic coordinates, with the Northern Hemisphere above the green curve and the Southern Hemisphere below. The North and South celestial poles are denoted by white stars. The color of the ephemeris curve corresponds the Julian Date of the observations, which span from 2018-05-06 through 2018-12-30. The comet's state vector was generally centered on the Galactic Plane throughout the entirety of this observation campaign. The Galactic image is attributed to ESA/Gaia/DPAC. Telescope Location Diameter (m) Field of View ((cid:48)) Pixel Size ((cid:48)(cid:48)) T7 T11 T30 T31 Nerpio, Spain Mayhill, NM, USA Siding Spring, Australia Siding Spring, Australia 0.41 0.50 0.50 0.50 42.3 × 28.2 54.3 × 36.2 41.6 × 27.8 55.9 × 55.9 1.26 1.62 1.62 2.20 Table 1. A summary of the telescopes utilized during the course of this observation campaign. The columns denote: 1) the name of the telescope; 2) the location of the observatory where the telescope is located; 3) the diameter of the telescope's primary mirror, in m; 4) the field of view of the telescope, in arcminutes; and 5) the effective pixel size of the telescope camera after binning, in arcseconds. equivalent width of 1580 A (Bessell 2005). Binning was always set to two pixels. Exposure times ranged from fifteen seconds to one minute, in order to ensure that the comet never traveled more than one binned pixel during the exposure. Binning by two also increased the signal-to-noise ratio per pixel and more appropriately sampled the typical seeing at these three observatory sites. Because the comet was usually located within crowded star fields, the images were scheduled specifically to ensure separation between the comet and nearby stars. We did this by comparing the position of the comet as determined by JPL Horizons at a given time to the URAT-1 (Zacharias et al. 2015) and UCAC-4 (Zacharias et al. 2012) stellar catalogs, considering all stars with magnitudes brighter than 19. We only scheduled observations for times when the angular separation between the comet position and the nearest star was at least 20(cid:48)(cid:48) for fifteen minutes. Ten images were scheduled for each night, and all ten images were first inspected visually to ensure a minimum quality threshold. Images that were obviously unsuitable for further analysis due to tracking or weather issues were immediately rejected. Bias, dark, and flat-field corrections to each image were done in a standard method automatically by iTelescope. Summaries of the observations taken before and after perihelion can be found in Tables 2 and 3, respectively. We acquired high quality imaging data of the comet on a total of 70 nights - 32 of which were taken before perihelion and 38 afterwards. The average spacing between consecutive observations is ∼ 3 − 4 days, although the cadence is not uniform. Due to poor weather at the observing sites, there were several time periods when images were not taken for 10-14 days. 4 EHLERT, MOTICSKA, AND EGAL Table 2. Observations of Comet 21P taken prior to perihelion. The columns denote: 1) The start date and time of the first image taken for each observing session; 2) the telescope utilized for these images (see Table 1); 3) the exposure time of each image, in seconds; 4) the right ascension of the comet at the time of observation; 5) the declination of the comet at the time of observation; 6) the heliocentric distance of the comet, in AU; 7) the geocentric distance of the comet, in AU; and 8) the Sun - Comet - Earth phase angle at the time of observation, in degrees. Date Telescope Exposure (s) RA DEC rH ∆ Phase Angle (◦) 2018-05-06 09:26:36 UT 2018-05-14 09:24:13 UT 2018-05-17 08:36:47 UT 2018-05-24 10:17:02 UT 2018-05-25 10:17:23 UT 2018-05-28 07:37:01 UT 2018-06-05 08:22:15 UT 2018-06-09 08:12:32 UT 2018-06-11 09:56:31 UT 2018-06-12 10:07:35 UT 2018-06-17 08:44:00 UT 2018-06-20 05:21:49 UT 2018-06-30 02:15:03 UT 2018-07-01 09:17:14 UT 2018-07-09 02:07:53 UT 2018-07-12 22:22:46 UT 2018-07-15 22:22:34 UT 2018-07-18 05:31:39 UT 2018-07-19 01:38:31 UT 2018-07-21 09:27:03 UT 2018-07-22 22:37:43 UT 2018-07-24 03:05:18 UT 2018-07-24 22:32:43 UT 2018-07-26 01:57:41 UT 2018-08-05 08:21:41 UT 2018-08-06 10:43:20 UT 2018-08-08 06:46:37 UT 2018-08-14 08:26:57 UT 2018-08-15 09:02:05 UT 2018-08-27 10:22:12 UT 2018-08-31 09:27:18 UT 2018-09-01 08:21:49 UT T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T7 T11 T7 T7 T7 T11 T7 T11 T7 T7 T7 T7 T11 T11 T11 T11 T11 T11 T11 T11 300 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 30 30 15 15 15 19h49m47.s28 +20◦50(cid:48)08.9(cid:48)(cid:48) 20h01m55.s58 +24◦26(cid:48)29.2(cid:48)(cid:48) 20h06m27.s72 +25◦50(cid:48)48.5(cid:48)(cid:48) 20h17m24.s46 +29◦19(cid:48)58.6(cid:48)(cid:48) 20h18m58.s72 +29◦50(cid:48)27.9(cid:48)(cid:48) 20h23m33.s47 +31◦19(cid:48)41.2(cid:48)(cid:48) 20h36m43.s20 +35◦36(cid:48)17.2(cid:48)(cid:48) 20h43m35.s13 +37◦47(cid:48)57.9(cid:48)(cid:48) 20h47m15.s53 +38◦57(cid:48)11.6(cid:48)(cid:48) 20h49m04.s60 +39◦31(cid:48)03.1(cid:48)(cid:48) 42◦18(cid:48)39.0(cid:48)(cid:48) 20h58m20.s87 21h04m02.s44 +43◦56(cid:48)32.6(cid:48)(cid:48) 21h26m24.s35 +49◦36(cid:48)49.4(cid:48)(cid:48) 50◦21(cid:48)23.8(cid:48)(cid:48) 21h29m45.s00 21h52m33.s74 +54◦44(cid:48)17.8(cid:48)(cid:48) 22h06m21.s95 +56◦52(cid:48)12.0(cid:48)(cid:48) 22h18m38.s49 +58◦29(cid:48)17.0(cid:48)(cid:48) 22h29m07.s24 +59◦41(cid:48)23.7(cid:48)(cid:48) 22h33m12.s32 +60◦07(cid:48)05.7(cid:48)(cid:48) 22h45m21.s38 +61◦16(cid:48)22.9(cid:48)(cid:48) 22h54m10.s76 +62◦00(cid:48)36.2(cid:48)(cid:48) 23h01m21.s82 +62◦33(cid:48)13.0(cid:48)(cid:48) 23h06m30.s16 +62◦54(cid:48)49.0(cid:48)(cid:48) 23h14m04.s35 +63◦24(cid:48)11.3(cid:48)(cid:48) 00h41m25.s95 +66◦22(cid:48)56.5(cid:48)(cid:48) 00h52m49.s76 +66◦28(cid:48)59.2(cid:48)(cid:48) 01h12m34.s43 +66◦31(cid:48)20.4(cid:48)(cid:48) 02h21m10.s06 +65◦17(cid:48)54.3(cid:48)(cid:48) 02h32m43.s46 +64◦51(cid:48)44.8(cid:48)(cid:48) 04h29m01.s41 +54◦28(cid:48)57.1(cid:48)(cid:48) 04h57m08.s17 +49◦09(cid:48)39.9(cid:48)(cid:48) 05h03m12.s99 +47◦45(cid:48)55.2(cid:48)(cid:48) 1.9188 1.8450 1.8176 1.7520 1.7427 1.7159 1.6414 1.6045 1.5854 1.5761 1.5308 1.5048 1.4164 1.4050 1.3386 1.3065 1.2820 1.2637 1.2570 1.2390 1.2271 1.2182 1.2122 1.2038 1.1341 1.1274 1.1164 1.0834 1.0784 1.0322 1.0228 1.0209 1.4965 1.3821 1.3416 1.2493 1.2367 1.2011 1.1069 1.0627 1.0404 1.0297 0.9788 0.9503 0.8572 0.8455 0.7780 0.7454 0.7204 0.7015 0.6947 0.6759 0.6635 0.6540 0.6476 0.6386 0.5602 0.5521 0.5388 0.4969 0.4902 0.4238 0.4088 0.4059 31.37 32.68 33.21 34.57 34.78 35.40 37.32 38.39 38.97 39.26 40.78 41.72 45.32 45.84 49.11 50.87 52.31 53.44 53.86 55.05 55.86 56.48 56.92 57.53 63.30 63.94 65.00 68.51 69.09 75.15 76.58 76.86 3. IMAGE CALIBRATION AND COMET MEASUREMENTS Astrometric and photometric calibration was performed for every image using two independent pipelines. The first is nearly identical to that presented in Hosek et al. (2013) and Blaauw et al. (2014), and utilizes the software package Astrometrica (Raab 2018) to identify reference stars within the image field of view. The second utilizes SExtractor (Bertin & Arnouts 1996) to detect and identify reference stars within the field of view. Each pipeline provides the astrometric solution for each image, as well as an image-specific zero-point for photometric calibration. The US Naval Observatory A2 Catalog (Monet et al. 1998, hereafter USNO A2) was the reference catalog for both pipelines, using their R-band magnitudes. 3.1. Comet Photometry DUST PRODUCTION OF COMET 21P DURING ITS 2018 APPARITION 5 Table 3. Observations of Comet 21P taken after perihelion. The columns are identical to those in Table 2. Date Telescope Exposure (s) RA DEC rH ∆ Phase Angle (◦) 2018-09-10 09:33:04 UT 2018-09-12 09:57:17 UT 2018-09-13 09:02:16 UT 2018-09-14 11:01:23 UT 2018-09-18 11:06:48 UT 2018-09-19 10:26:18 UT 2018-09-21 10:57:54 UT 2018-09-25 10:56:50 UT 2018-09-28 09:47:28 UT 2018-09-29 10:02:27 UT 2018-09-30 10:02:09 UT 2018-10-05 10:22:13 UT 2018-10-08 11:47:13 UT 2018-10-10 10:57:19 UT 2018-10-14 17:02:37 UT 2018-10-15 17:37:43 UT 2018-10-21 15:53:22 UT 2018-10-24 16:43:04 UT 2018-10-26 16:41:50 UT 2018-11-04 14:22:59 UT 2018-11-10 17:42:55 UT 2018-11-11 17:37:57 UT 2018-11-25 15:12:21 UT 2018-11-29 13:22:10 UT 2018-12-01 14:32:27 UT 2018-12-04 14:22:00 UT 2018-12-05 14:39:02 UT 2018-12-06 14:10:02 UT 2018-12-07 16:41:41 UT 2018-12-18 15:06:26 UT 2018-12-23 13:42:10 UT 2018-12-24 12:16:50 UT 2018-12-25 16:52:33 UT 2018-12-26 15:02:55 UT 2018-12-27 15:01:35 UT 2018-12-28 16:03:39 UT 2018-12-29 14:03:17 UT 2018-12-30 14:32:47 UT T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T11 T31 T31 T31 T31 T31 T30 T30 T30 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 T31 15 15 15 15 15 15 15 15 15 15 15 15 15 15 30 30 30 30 30 30 30 30 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 60 +33◦00(cid:48)54.0(cid:48)(cid:48) 05h50m01.s97 +29◦31(cid:48)08.1(cid:48)(cid:48) 05h58m17.s73 +27◦50(cid:48)49.6(cid:48)(cid:48) 06h02m00.s79 +25◦58(cid:48)02.2(cid:48)(cid:48) 06h06m02.s29 +19◦06(cid:48)25.0(cid:48)(cid:48) 06h19m35.s76 +17◦28(cid:48)43.0(cid:48)(cid:48) 06h22m36.s17 +14◦09(cid:48)21.3(cid:48)(cid:48) 06h28m32.s40 +07◦53(cid:48)50.1(cid:48)(cid:48) 06h39m08.s32 +03◦35(cid:48)06.6(cid:48)(cid:48) 06h46m05.s59 +02◦10(cid:48)22.9(cid:48)(cid:48) 06h48m19.s12 +00◦48(cid:48)29.8(cid:48)(cid:48) 06h50m26.s80 07h00m03.s79 −05◦33(cid:48)00.0(cid:48)(cid:48) 07h05m06.s74 −09◦02(cid:48)37.2(cid:48)(cid:48) 07h08m02.s70 −11◦08(cid:48)33.2(cid:48)(cid:48) 07h13m35.s36 −15◦19(cid:48)15.7(cid:48)(cid:48) 07h14m45.s87 −16◦15(cid:48)26.0(cid:48)(cid:48) 07h20m22.s63 −21◦11(cid:48)30.7(cid:48)(cid:48) 07h22m28.s72 −23◦25(cid:48)39.1(cid:48)(cid:48) 07h23m34.s78 −24◦48(cid:48)11.5(cid:48)(cid:48) 07h25m44.s24 −30◦05(cid:48)24.2(cid:48)(cid:48) 07h24m36.s53 −33◦00(cid:48)43.8(cid:48)(cid:48) 07h24m13.s71 −33◦26(cid:48)03.6(cid:48)(cid:48) 07h13m41.s68 −37◦51(cid:48)56.4(cid:48)(cid:48) 07h09m11.s19 −38◦38(cid:48)02.2(cid:48)(cid:48) 07h06m37.s14 −38◦57(cid:48)03.9(cid:48)(cid:48) 07h02m38.s81 −39◦18(cid:48)37.8(cid:48)(cid:48) 07h01m15.s16 −39◦24(cid:48)14.8(cid:48)(cid:48) 06h59m52.s90 −39◦28(cid:48)52.9(cid:48)(cid:48) 06h58m18.s79 −39◦33(cid:48)09.8(cid:48)(cid:48) 06h42m13.s65 −39◦22(cid:48)51.9(cid:48)(cid:48) 06h35m07.s87 −38◦48(cid:48)47.7(cid:48)(cid:48) 06h33m49.s47 −38◦40(cid:48)25.3(cid:48)(cid:48) 06h32m11.s71 −38◦28(cid:48)59.4(cid:48)(cid:48) 06h30m57.s18 −38◦19(cid:48)30.4(cid:48)(cid:48) 06h29m37.s92 −38◦08(cid:48)39.3(cid:48)(cid:48) 06h28m16.s75 −37◦56(cid:48)40.8(cid:48)(cid:48) 06h27m06.s85 −37◦45(cid:48)37.9(cid:48)(cid:48) 06h25m50.s65 −37◦32(cid:48)45.7(cid:48)(cid:48) 1.0128 1.0133 1.0138 1.0146 1.0197 1.0214 1.0255 1.0360 1.0457 1.0494 1.0532 1.0748 1.0900 1.1004 1.1249 1.1312 1.1697 1.1909 1.2054 1.2738 1.3240 1.3324 1.4534 1.4886 1.5072 1.5344 1.5437 1.5527 1.5628 1.6637 1.7096 1.7183 1.7293 1.7379 1.7472 1.7569 1.7654 1.7749 0.3918 0.3923 0.3931 0.3943 0.4021 0.4047 0.4111 0.4269 0.4409 0.4461 0.4514 0.4804 0.4996 0.5124 0.5409 0.5479 0.5889 0.6101 0.6240 0.6853 0.7266 0.7332 0.8243 0.8499 0.8633 0.8831 0.8898 0.8963 0.9037 0.9802 1.0172 1.0244 1.0337 1.0410 1.0490 1.0575 1.0650 1.0734 77.99 77.83 77.70 77.52 76.49 76.16 75.40 73.60 72.08 71.53 70.98 68.06 66.22 65.03 62.44 61.82 58.29 56.54 55.40 50.59 47.48 47.00 40.67 39.04 38.22 37.05 36.67 36.30 35.89 32.18 30.72 30.46 30.14 29.90 29.64 29.38 29.15 28.91 With astrometric and photometric calibration models determined for each image, the magnitude of the comet was then deter- mined. Our dust production measurements were determined using the magnitude of the comet as measured in the 10(cid:48)(cid:48) × 10(cid:48)(cid:48) square aperture centered on the brightest pixel associated with the comet emission. This magnitude was determined by two independent programs: Fotometria con Astrometrica (FoCAs, Castellano-Roig 2018) was utilized to ingest the calibration model as derived by Astrometrica, whereas an independent Python-based program was utilized to ingest the SExtractor derived cal- ibration model. Both programs were confirmed to provide consistent magnitudes for the data presented in Hosek et al. (2013) 6 EHLERT, MOTICSKA, AND EGAL and Blaauw et al. (2014). The largest difference between the analyses of Hosek et al. (2013) and Blaauw et al. (2014) and this work is the choice of reference catalog. Both of these older works calibrated their images with respect to the Carlsberg Meridian Catalog, Release 14 (Copenhagen University et al. 2006, hereafter CMC-14). During this apparition, comet 21P was positioned outside of the CMC-14 survey area for the months of July and August4. The only other reference catalog compatible with FoCAs is USNO A2, and we therefore used this catalog throughout the observation campaign. The USNO A2 catalog provides R-band magnitudes for stars across the entire sky and also provides a useful means for cross- calibrating the measurements by using two separate pipelines. One limitation of this catalog, however, is that reference stars in different regions of the sky can be subject to rather large calibration discrepancies (of order ∼ 0.3 − 0.5 mag), introducing a systematic uncertainty associated with comparing the nightly variations of the comet's magnitude (and subsequently Af ρ measurements) as it moves across the sky5. In addition to the magnitudes used to measure dust production, we also present total coma magnitudes for each night. These magnitudes were determined in circular apertures. The radius of this total coma aperture was determined by visual inspection of the images from each night, and varied from 5-21 pixels. The aperture radius associated with the total coma was determined by visually identifying a sharp edge in the coma's surface brightness in the direction of Sun, and choosing the smallest integer radius that encloses this edge. The corresponding magnitudes were only determined using the SExtractor-based pipeline. 3.2. Determining Af ρ from Comet Photometry With photometric magnitudes of the comet measured for each image, we then convert these magnitudes into dust production rates using the Af ρ parameter described by A'hearn et al. (1984). As discussed in this paper, Af ρ is calculated from the comet's measured photometric flux F within a circular aperture of radius ρ (in cm) as where rH (in AU) and ∆ (in cm) correspond to the heliocentric and geocentric distances to the comet, respectively. The pho- tometric flux of the Sun is given as F(cid:74). For this work, we use the same calculations as Hosek et al. (2013) and Blaauw et al. (2014). The effective radius ρ of the square aperture is given as A(θ)f ρ = × ∆2 × r2 H × 4 ρ (cid:19) (cid:18) F F(cid:74) (cid:19) (cid:18) 1.12838r (cid:19) 206265(cid:48)(cid:48) −(R+27.15) 2.5 = 10 × ∆ ρ = tan (cid:18) F F(cid:74) where r = 5(cid:48)(cid:48). The comet's photometric flux ratio is calculated from its observed magnitude R as where we assume the apparent R-band magnitude of the Sun is -27.15 (Mann & von Braun 2015). As suggested in Equation 1, the initial value of Af ρ does not account for how the brightness of the comet varies with phase angle. The initial determination of A(θ)f ρ is then corrected for the phase angle using the Schleicher phase function (shown in Figure 2), which we will denote as A(0)f ρ. This phase function is a splice of measurements derived from observations of Comet Halley (Schleicher et al. 1998) at low phase angles and a Henyey-Greenstein model at high phase angles (Marcus 2007). A few example images of comet 21P as observed at the beginning and end of this observing campaign, as well as at perihelion are shown in Figure 3. The resulting Af ρ measurements for each night can be found in Tables 4 and 5. 4. RESULTS We plot the measured values of A(0)f ρ in two different sets of units. First, in Figure 4a, we show how the total apparent magnitude and corresponding A(0)f ρ measurement vary with time, using the date of perihelion as a reference date. Second, we show how A(0)f ρ varies with heliocentric distance in Figure 4b. This plot uses the same re-normalized, signed distances x described in Egal19, where x is defined as x = (rH − q) t − t(q) t − t(q) 4 Since these papers were completed, Carlsberg Meridian Catalog Release 15 was made public. Neither data release covers regions of the sky with declination of δ > 60◦, where comet 21P was located for these months. 5 This level of plate-to-plate variation was documented in the for the USNO A1 catalog (http://vizier.cfa.harvard.edu/vizier/VizieR/pmm/usno2.htx#pht). Although the USNO A2 analysis procedure made efforts to smooth out the variance between plates, the extent to which systematic uncertainties were reduced was not quantified. (1) (2) (3) (4) DUST PRODUCTION OF COMET 21P DURING ITS 2018 APPARITION 7 Figure 2. The Schleicher phase function assumed for accounting for variations in observing geometry between different nights. This phase function is normalized to have a value of unity at 0◦. where q is the perihelion distance of comet 21P, t corresponds to the date/time of the observations, and t(q) is the date/time at perihelion. We arbitrarily define pre-perihelion distances in these units to be negative and post-perihelion distances to be positive. Both of these figures, in particular Figure 4b, show that the comet reached its maximum value of Af ρ before perihelion. Our maximum single value of Af ρ (1594 cm) was measured on the night of 2018-08-14 (27 days before perihelion), although the interpretation of this single measurement is limited by the lack of observational data on neighboring nights and the overall uncertainty in absolute photometric calibration. Similarly high Af ρ values were measured on 2018-08-15 and 2018-08-31, corresponding to 26 and 10 days before perihelion, respectively. This range of dates corresponds to variations in heliocentric distance between 1.08 AU and 1.02 AU. Unfortunately, few observations of the comet were acquired in the latter half of August and first half of September due to poor weather at the two Northern Hemisphere observation sites. While these measurements provide strong evidence that the maximum value of Af ρ occurred before perihelion, we cannot state with certainty precisely where within this time window comet 21P reached its maximum value of Af ρ. These figures also make it clear that comet 21P's Af ρ values are highly asymmetric around the peak value, with a much steeper descending branch than ascending branch, regardless of whether Af ρ is considered a function of heliocentric distance or time. In Egal19, the Af ρ curve in Figure 4b was fit with a double-exponential model of the form (cid:40) Af ρ(x) = K1 + Af ρ(xmax) ∗ 10−γ1x−xmax K1 + Af ρ(xmax) ∗ 10−γ2x−xmax x ≤ 0 x ≥ 0 (5) where K1 is the asymptotic value of Af ρ, and γ1 and γ2 correspond to the logarithmic slopes of the ascending and descending branches, respectively. The best-fit double-exponential model is overlaid on the measurements in Figure 4b, and has K1 = 121, xmax = 0.04, Af ρ(xmax) = 1653. The best fit logarithmic slopes are γ1 = 2.10 and γ2 = 7.38, respectively. This model form better describes the data than other shapes such as a Gaussian, Moffat, or Lorentzian. It also offers more physically sensible model parameters than a broken power-law model fit, as in this latter case the two branches do not agree on a single prediction for the peak value of Af ρ. With imaging observations spanning more than 100 days on either side of perihelion and an average cadence of 3 − 4 days between measurements, the observation campaign described in this work provides a unique view into the short and medium term 5. DISCUSSION 8 EHLERT, MOTICSKA, AND EGAL Figure 3. Images of comet 21P as observed at different times during this observation campaign. In all three images the cyan circle denotes a fixed radius of 15(cid:48)(cid:48). Top Left: An image of comet 21P taken on the night of 2018-05-17. The comet is consistent with point-like emission. All images have been smoothed using a Gaussian kernel with a σ = 2 pixels. Top Right: An image of comet 21P taken on the night of 2018-12-30, the last night of the observation campaign. The comet emission is once again largely point-like. Bottom: An image of comet 21P taken on the night of 2018-09-10, when the comet was at perihelion. In this image, extended emission from the comet's coma is obvious. 111120149198265352458584729893107621223630842959881710821395176021702628798290103121144173206245289338 DUST PRODUCTION OF COMET 21P DURING ITS 2018 APPARITION 9 Table 4. Measurements of Af ρ for Comet 21P during its 2018 apparition pre-perihelion. The columns denote: 1) The date and time of the observations; 2) the physical size of the 10(cid:48)(cid:48) × 10(cid:48)(cid:48) square aperture, in units of 1000 km; 3) the extent of the coma used to measure the total magni- tude, in units of image pixels (with the corresponding extent in arcseconds given in parentheses); 4) The R-band magnitude of the comet as measured within the 10(cid:48)(cid:48)× 10(cid:48)(cid:48) aperture used for Af ρ measurements; 5) The R-band magnitude of the entire coma, as measured in a circular aperture with a radius denoted in column 3; 6) the resultant measurement of Af ρ before accounting for the phase angle correction, in units of cm; and 7) the final phase-angle corrected values of Af ρ, also in units of cm. Date ρ (×1000 km) 2018-05-06 2018-05-14 2018-05-17 2018-05-24 2018-05-25 2018-05-28 2018-06-05 2018-06-09 2018-06-11 2018-06-12 2018-06-17 2018-06-20 2018-06-30 2018-07-01 2018-07-09 2018-07-12 2018-07-15 2018-07-18 2018-07-19 2018-07-21 2018-07-22 2018-07-23 2018-07-24 2018-07-24 2018-07-26 2018-08-05 2018-08-06 2018-08-08 2018-08-14 2018-08-15 2018-08-27 2018-08-31 2018-09-01 6.12 5.65 5.49 5.12 5.06 4.91 4.53 4.20 4.26 4.22 4.01 3.89 3.50 3.46 3.19 3.05 2.95 2.87 2.84 2.77 2.71 2.71 2.68 2.65 2.62 2.29 2.26 2.20 2.03 2.00 1.74 1.68 1.66 Radius 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 5 (3.1) 7 (4.3) 7 (4.3) 7 (4.3) 7 (5.5) 7 (4.3) 7 (5.5) 9 (7.1) 9 (7.1) 11 (8.7) 11 (6.7) 11 (8.7) 11 (8.7) 11 (6.7) 13 (10.3) 15 (9.2) 15 (9.2) 15 (9.2) 17 (10.4) 17 (10.4) 19 (11.6) 19 (11.6) 19 (11.6) RAf ρ RComa A(θ)f ρ (cm) A(0)f ρ (cm) 17.00 15.84 16.01 15.71 15.53 15.44 14.88 14.60 14.52 14.80 14.77 14.34 13.86 13.71 13.66 13.01 12.75 12.72 12.89 12.93 12.79 12.87 12.72 12.71 12.32 11.77 11.66 11.66 11.25 11.25 11.26 10.79 11.03 61.63 156.24 126.64 147.80 171.39 176.56 255.85 300.90 322.27 245.06 230.42 325.67 424.40 478.74 432.02 732.79 877.97 859.17 716.49 658.89 725.33 663.15 747.47 744.69 1037.21 1300.70 1391.04 1319.57 1594.35 1539.27 1071.46 1510.27 1198.40 17.00 15.25 15.38 15.22 14.98 15.28 14.42 14.03 13.98 14.09 13.94 13.99 13.29 13.22 13.03 12.30 12.02 11.67 11.92 11.79 11.98 11.82 11.75 11.75 11.24 10.14 9.90 9.89 9.45 9.25 9.34 8.86 9.07 26.36 65.54 52.71 60.27 69.67 71.11 100.10 115.85 122.99 93.12 85.61 119.35 147.97 165.82 144.08 240.39 284.88 276.85 230.38 210.84 231.61 211.66 238.43 237.46 330.72 426.49 458.96 440.50 559.07 544.94 427.25 621.60 496.33 behavior of comet 21P. We will now discuss how these observations compare to observations of comet 21P taken in previous apparitions and what they imply for the physical properties of this comet. The steep and asymmetric logarithmic slopes in the dust production of this comet are both in agreement with observations of this comet as during previous apparitions. Significant differences between pre- and post-perihelion behavior of this comet was 10 EHLERT, MOTICSKA, AND EGAL Table 5. Measurements of Af ρ for Comet 21P during its 2018 apparition post-perihelion. The columns are identical to those of Table 4. Date ρ (×1000 km) 2018-09-10 2018-09-12 2018-09-13 2018-09-14 2018-09-18 2018-09-19 2018-09-21 2018-09-25 2018-09-28 2018-09-29 2018-09-30 2018-10-05 2018-10-08 2018-10-10 2018-10-14 2018-10-15 2018-10-21 2018-10-24 2018-10-26 2018-11-04 2018-11-10 2018-11-11 2018-11-25 2018-11-29 2018-12-01 2018-12-04 2018-12-05 2018-12-06 2018-12-07 2018-12-18 2018-12-23 2018-12-24 2018-12-25 2018-12-26 2018-12-27 2018-12-28 2018-12-29 2018-12-30 1.60 1.60 1.62 1.62 1.65 1.66 1.68 1.75 1.81 1.83 1.84 1.96 2.05 2.09 2.21 2.24 2.41 2.50 2.54 2.83 2.98 3.01 3.37 3.47 3.53 3.62 3.64 3.67 3.70 4.01 4.16 4.19 4.23 4.26 4.29 4.32 4.35 4.40 Radius 21 (12.8) 21 (12.8) 21 (12.8) 21 (12.8) 19 (11.6) 19 (11.6) 19 (11.6) 15 (9.2) 15 (9.2) 13 (7.9) 13 (7.9) 11 (6.7) 11 (6.7) 11 (6.7) 9 (4.1) 9 (4.1) 7 (3.2) 7 (3.2) 5 (2.3) 5 (3.0) 5 (3.0) 5 (3.0) 5 (3.0) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) 5 (2.3) RAf ρ RComa A(θ)f ρ (cm) A(0)f ρ (cm) 11.27 11.19 10.97 11.41 11.67 11.49 11.44 12.30 12.27 12.22 12.45 12.83 12.96 12.83 12.81 12.85 13.97 13.90 14.06 14.62 14.54 14.59 15.33 15.15 15.47 15.49 15.62 15.70 15.66 16.02 16.11 16.12 16.17 16.20 16.21 16.20 16.14 15.59 884.72 956.35 1182.39 793.27 661.23 790.42 862.93 430.47 480.68 518.36 431.19 354.89 344.92 412.58 475.56 471.69 196.77 225.87 203.03 146.42 173.23 167.95 103.97 130.01 99.52 101.59 91.55 85.99 91.04 75.91 74.80 75.16 72.78 72.05 72.56 74.43 79.42 134.04 8.95 9.20 9.16 9.27 9.53 9.44 9.52 10.65 10.61 10.65 10.83 11.52 11.68 11.38 11.43 11.38 13.07 12.88 13.17 14.11 14.32 14.09 15.08 14.45 14.90 14.85 14.92 14.97 14.84 15.27 16.03 16.06 15.94 15.71 15.33 15.31 15.24 14.96 376.07 405.04 499.31 333.56 271.60 322.30 345.95 166.04 179.78 191.81 157.89 123.56 116.91 137.76 154.78 152.81 62.79 72.05 64.90 48.15 58.81 57.35 38.69 49.57 38.41 39.91 36.17 34.17 36.39 32.08 32.31 32.59 31.70 31.50 31.85 32.79 35.11 59.47 observed as early as its 1985 apparition (Schleicher et al. 1987), who measured peak Af ρ values approximately one month before perihelion. Our peak Af ρ value is measured at a very similar time. During the 1998 apparition, Lara et al. (2003) measured a peak Af ρ value approximately 2 weeks before perihelion. The times of peak Af ρ value in both of these papers agree with the measurements presented here, especially given the limited observational data available for the comet during this same time period (∼ 14 − 30 days before perihelion) during all three apparitions. DUST PRODUCTION OF COMET 21P DURING ITS 2018 APPARITION 11 Figure 4. The measured values of Af ρ during the 2018 observation campaign. Left: The total apparent magnitude (solid curve) and A(0)f ρ (gray diamonds) of comet 21P as a function of time and heliocentric distance during its 2018 apparition. The date at which the comet reached perihelion, which we used as the reference date, is denoted by the vertical gray line. While the comet reached peak total brightness at perihelion, its dust production rate (as measured by Af ρ) peaked approximately ∼ 10−30 days prior to perihelion. Right: The measured values of A(0)f ρ of comet 21P as a function of heliocentric distance. The distances have been normalized to define post-perihelion distances as positive and pre-perihelion distances as negative. The solid red curve is the best-fit double-exponential model to these observations from Egal19. When considering similar ranges of heliocentric distance and fitting to a single power-law model, similar logarithmic slopes and normalizations to those measured in Blaauw et al. (2014) are also measured for this apparition. With the longer baseline of measurements presented here, our newly determined double-exponential model naturally accounts for discrepancies between the measured logarithmic slopes of Blaauw et al. (2014) (which fit pre-perihelion observations) and Pittichov´a et al. (2008) (which fit post-perihelion observations). Similar behavior was also observed in the comet's water production during its 1998 and 2005 apparitions (Combi et al. 2011). Utilizing observations of the Solar Heliospheric Observatory's (SOHO) Solar Wind ANisotropies (SWAN) camera, the water production rate was observed to scale as r−1.74 during egress. In fact, these indices are broadly consistent with the indices measured in this work. This broad overall consistency between water and dust production rates suggests that these two quantities are physically linked to one another. during ingress and r−11.9 H H Observational data are beginning to support the idea that the dust ejection behavior of comet 21P has not changed significantly over the past few apparitions. Although it is prohibitive to combine the data from every apparition into a single model for comet 21P's dust production rate, it is reassuring from the standpoint of modeling dust ejection from the surface of the comet that a single model is at least qualitatively consistent with all of the apparitions thus observed. We caution, however, that comet 21P has undergone rather dramatic and significant changes in its orbital elements since its discovery. The relative consistency of the past few apparitions does not necessarily imply stability in comet 21P's dust production on longer time scales, especially when considering apparitions earlier than its 1959-1965 orbit, when its non-gravitational forces underwent a sudden change that modified the comet's orbit. The results presented here demonstrate the constraining power of a long term monitoring program of a single comet for meteor shower forecasting. Although this observation campaign has measured Af ρ values for comet 21P at relatively high temporal resolution and over a long period of time on either side of perihelion, poor weather during July, August, and September resulted in relatively few observations in the weeks leading up to perihelion. Unfortunately, these were when Af ρ values for the comet were at their maximum. Subsequently, this particular time window is the most crucial to accurately model for forecasts of the Draconid meteor shower. We encourage future observation campaigns for comet 21P to place high priority on acquiring images of the comet during this time. Indeed, a modification of the activity slope induced by additional perihelion measurements can 12 EHLERT, MOTICSKA, AND EGAL modify the time of predicted shower maximum intensity by several minutes, and even hours when the corresponding activity profiles are derived from a small number of simulated meteoroids. The ultimate test for these observations will come when Earth is first predicted to encounter meteoroids ejected during this apparition. Simulations conducted over the period 1850-2030 in Egal19 point toward three potential Draconid outbursts over the next decade. In 2025, the Earth might approach a portion of the 2018 trail closer than 5 × 10−2 AU. However, most of the Draconid activity expected in 2025 will rather be produced by meteoroids ejected during the 2005 and 2012 apparitions. From preliminary estimates, an outburst caused by the 2018 trail should almost certainly occur in 2078. We anticipate a follow-up paper comparing the observed Draconid activity of this shower to predictions derived from these observations. As discussed in Egal19, these measurements of comet 21P's dust production can be immediately coupled to a dynamical stream model in order to produce forecasts for future Draconid shower outbursts. With these measurements in hand, these predictions can only improve. This work was supported by the NASA Meteoroid Environment Office under contract 80MSFC18C0011 (S.E.), NASA Mar- shall Space Flight Center's Internship Program (N.M.), and cooperative agreement 80NSSC18M0046 (A.E.). We thank Dr. William Cooke for his consistent support for this work and providing the resources required to carry out these long-term obser- vations. We thank Herbert Raab and Julio Castellano Roig for making the software programs Astrometrica and FoCAs publicly available. This work is based on observations taken using the iTelescope network, and we thank the staff of iTelescope for their efforts in developing the telescope network to enable such an observing campaign to be carried out. We finally thank the referee for their careful and throrough review of the initial manuscript. Their comments have certainly improved this paper. Software: Astrometrica (Raab 2018), FoCAs (Castellano-Roig 2018), SExtractor (Bertin & Arnouts 1996), Astropy (Astropy Collaboration et al. 2013, 2018) REFERENCES A'Hearn, M. F., Millis, R. C., Schleicher, D. O., Osip, D. J., & Laboratory, N. J. P. 2019, HORIZONS System, , . Birch, P. V. 1995, Icarus, 118, 223 https://ssd.jpl.nasa.gov/?horizons A'hearn, M. F., Schleicher, D. G., Millis, R. L., Feldman, P. D., & Lara, L. M., Licandro, J., Oscoz, A., & Motta, V. 2003, A&A, 399, Thompson, D. T. 1984, AJ, 89, 579 763 Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, Mann, A. W., & von Braun, K. 2015, Publications of the A&A, 558, A33 Astropy Collaboration, Price-Whelan, A. M., Sipocz, B. M., et al. 2018, AJ, 156, 123 Astronomical Society of the Pacific, 127, 102 Marcus, J. N. 2007, International Comet Quarterly, 29, 119 McFadden, L. A., A'Hearn, M. F., Feldman, P. D., et al. 1987, Bertin, E., & Arnouts, S. 1996, Astronomy and Astrophysics Icarus, 69, 329 Supplement Series, 117, 393 Bessell, M. S. 2005, Annual Review of Astronomy and Astrophysics, 43, 293 Blaauw, R. C., Suggs, R. M., & Cooke, W. J. 2014, Meteoritics and Planetary Science, 49, 45 Castellano-Roig, J. 2018, Fotometria Con Astrometrica, a Software Tool, v3.61, , . http://www.astrosurf.com/orodeno/focas/ Combi, M. R., Lee, Y., Patel, T. S., et al. 2011, AJ, 141, 128 Copenhagen University, O., Institute, A. O., Cambridge, Uk, & Real Instituto Y Observatorio de La Armada, F. E. S. 2006, VizieR Online Data Catalog, I/304 Hosek, Matthew W., J., Blaauw, R. C., Cooke, W. J., & Suggs, R. M. 2013, AJ, 145, 122 Kronk, G. W. 2014, Meteor Showers (Springer), doi:10.1007/978-1-4614-7897-3 Monet, D., Canzian, B., Harris, H., et al. 1998, VizieR Online Data Catalog, I/243 Pittichov´a, J., Woodward, C. E., Kelley, M. S., & Reach, W. T. 2008, AJ, 136, 1127 Raab, H. 2018, Astrometrica Software, Shareware for Research Grade CCD Photometry, v4.11.1.442, , . http://www.astrometrica.at/ Schleicher, D. G., Millis, R. L., & Birch, P. V. 1987, A&A, 187, 531 -- . 1998, Icarus, 132, 397 Sekanina, Z. 1985, AJ, 90, 827 Yeomans, D. K. 1971, The Astronomical Journal, 76, 83 Zacharias, N., Finch, C. T., Girard, T. M., et al. 2012, VizieR Online Data Catalog, I/322A Zacharias, N., Finch, C., Subasavage, J., et al. 2015, AJ, 150, 101
1707.07634
1
1707
2017-07-24T16:25:24
No large population of unbound or wide-orbit Jupiter-mass planets
[ "astro-ph.EP" ]
Gravitational microlensing is the only method capable of exploring the entire population of free-floating planets down to Mars-mass objects, because the microlensing signal does not depend on the brightness of the lensing object. A characteristic timescale of microlensing events depends on the mass of the lens: the less massive the lens, the shorter the microlensing event. A previous analysis of 474 microlensing events found an excess of very short events (1-2 days) - more than known stellar populations would suggest - indicating the existence of a large population of unbound or wide-orbit Jupiter-mass planets (reported to be almost twice as common as main-sequence stars). These results, however, do not match predictions of planet formation theories and are in conflict with surveys of young clusters. Here we report the analysis of a six times larger sample of microlensing events discovered during the years 2010-2015. Although our survey has very high sensitivity (detection efficiency) to short-timescale (1--2 days) microlensing events, we found no excess of events with timescales in this range, with a 95% upper limit on the frequency of Jupiter-mass free-floating or wide-orbit planets of 0.25 planet per main-sequence star. We detected a few possible ultrashort-timescale events (with timescales of less than 0.5 day), which may indicate the existence of Earth- and super-Earth-mass free-floating planets, as predicted by planet-formation theories. [abridged]
astro-ph.EP
astro-ph
No large population of unbound or wide-orbit Jupiter- mass planets Przemek Mr´oz1∗, Andrzej Udalski1, Jan Skowron1, Radosław Poleski2,1, Szymon Kozłowski1, Michał K. Szyma´nski1, Igor Soszy´nski1, Łukasz Wyrzykowski1, Paweł Pietrukowicz1, Krzysztof Ulaczyk3,1, Dorota Skowron1 & Michał Pawlak1 1Warsaw University Observatory, Aleje Ujazdowskie 4, 00-478 Warsaw, Poland 2Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA 3Department of Physics, University of Warwick, Coventry CV4 7AL, UK Planet formation theories predict that some planets may be ejected from their parent sys- tems as result of dynamical interactions and other processes1–3. Unbound planets can also be formed through gravitational collapse, in a similar way to that in which stars form4. A hand- ful of free-floating planetary-mass objects have been discovered by infrared surveys of young stellar clusters and star-forming regions5,6 as well as wide-field surveys7, but these surveys are incomplete8–10 for objects below 5 MJup. Gravitational microlensing is the only method capable of exploring the entire population of free-floating planets down to Mars-mass objects, because the microlensing signal does not depend on the brightness of the lensing object. A characteristic timescale of microlensing events depends on the mass of the lens: the less mas- sive the lens, the shorter the microlensing event. A previous analysis of 474 microlensing events found an excess of very short events11 (1–2 days) – more than known stellar popula- tions would suggest – indicating the existence of a large population of unbound or wide-orbit Jupiter-mass planets (reported to be almost twice as common as main-sequence stars). These results, however, do not match predictions of planet formation theories3,12 and are in conflict with surveys of young clusters8–10. Here we report the analysis of a six times larger sample of microlensing events discovered during the years 2010–2015. Although our survey has very ∗Corresponding author. 1 high sensitivity (detection efficiency) to short-timescale (1–2 days) microlensing events, we found no excess of events with timescales in this range, with a 95% upper limit on the fre- quency of Jupiter-mass free-floating or wide-orbit planets of 0.25 planet per main-sequence star. We detected a few possible ultrashort-timescale events (with timescales of less than 0.5 day), which may indicate the existence of Earth- and super-Earth-mass free-floating planets, as predicted by planet-formation theories. The sample of 2,617 microlensing events we analysed was selected from data collected dur- ing the fourth phase of the Optical Gravitational Lensing Experiment13 (OGLE-IV) during the years 2010–2015. The survey is monitoring dense fields towards the Galactic centre, nine of which (about 12.6 square degrees in total) were observed with a cadence of either 20 min or 60 min, al- lowing the detection of extremely short microlensing events. We analysed the light curves of almost 50 million stars identified on deep stacked images of each field; each light curve consisted of 4,500–12,000 data points. The selection of events was conducted in three steps, described in detail in the Methods. First, we searched for "bumps" in the light curves, which we define as at least three consecutive points 3σbase above the baseline level (σbase is the dispersion of points outside a 360-day window centred on the bump). To minimize contamination from moving objects (like asteroids) and photometry artifacts, we required that the centre of the additional flux coincided with the centre of the star. Events with a very low signal-to-noise ratio and those exhibiting a variability in the baseline were also rejected by these criteria. Next, we removed any remaining artefacts located mainly near the edges of the CCD camera, low-amplitude pulsating red giants, and other variable stars (dwarf novae, flaring stars) that have multiple bumps in their light curves. Finally, we fitted the microlensing point-source point-lens model to the data and required that the model describe the data appropriately. The lensing model has three parameters: the time t0 and the projected separation u0 (in Einstein radius units) between the lens and the source during 2 the closest approach, and the Einstein radius crossing time tE. (The angular Einstein radius θE of a lens depends on its mass M and relative lens-source parallax πrel = 1 au(D−1 S ), where DL and DS are the distances to the lens and the source, respectively, as follows: θE = √κM πrel, where κ = 8.14 mas/M(cid:12)). Two additional parameters describe the source flux Fs and blended unmagnified flux Fb from possible unresolved neighbours and/or the lens itself. To ensure that the shortest events were not mistaken for stellar flares, we required at least four data points on the L − D−1 rising branch of the light curve (two if the descending part of the light curve was also covered). Using our detection efficiency simulations (see below), we found that the event timescales cannot be reliably measured for faint, highly-blended events (with blending parameter fs = Fs/(Fs+ Fb) < 0.1, that is, less than 10% of the baseline flux comes from the source), which was predicted theoretically14. Therefore, to ensure that our final results were robust, highly-blended events were not included in our sample of high-quality events. Thus, regardless of the timescale, there can be no systematic shift between measured and real timescales for simulated data. The final distribution of the event timescales is shown in Fig. 1. To calculate the detection efficiency we conducted extensive image-level simulations, in which artificial microlensing events were injected into real OGLE images using the point spread function derived from the neighbouring stars. In total, 8.6 millions of artificial events were sim- ulated. Parameters u0, t0, and log tE were drawn from uniform distributions, but sources were randomly drawn from the luminosity function of each subfield. For simulated events we applied exactly the same selection criteria as those applied to the observed sample of events. Detection efficiency curves for all analyzed fields are shown in Extended Data Fig. 2. The detection-efficiency-corrected histogram of event timescales is presented in Fig. 2 and, 11. clearly, does not show the excess of events with timescales tE ≈ 1 − 2 d, claimed in ref. The difference (at a confidence level of 2.5 − 3σ) can be explained in part by the relatively small number of events found in the earlier analysis11. In addition to the 2,617 events analysed in this 3 work, we detected over twenty short-duration events that showed clear signatures of binarity15 and did not pass our strict selection criteria for the fit quality. Owing to lower photometric precision, such events may have been mistaken for single short-timescale events. It is also possible that event timescales measured in the previous work suffer from systematic effects (differential refraction, unphysical treatment of negative blending). Thanks to better image quality (smaller pixel scale, better seeing) and a narrower filter, our photometry is less prone to such systematic effects. function L = (cid:81) We modelled the observed distribution of event timescales by maximizing the likelihood i p(tE,i), where p(tE) = pmodel(tE)ε(tE) is the normalized predicted timescale distribution (corrected for the detection efficiency ε(tE))11, 16. We adopted a standard Galactic model17, 18 of the distribution and kinematics of stars and tested several mass functions. In our best-fitting model, the initial mass function (IMF) can be approximated as a broken power law with slopes −0.8 in the brown dwarf regime (0.01 < M < 0.08 M(cid:12)), −1.3 for low-mass stars (0.08 < M < 0.5 M(cid:12)), and −2.0 for M > 0.5 M(cid:12). We assumed that all stars more massive than 1 M(cid:12) evolved into white dwarfs, neutron stars and black holes, depending on their initial mass, and we assumed the binary fraction fbin = 0.4. The model is marked with a purple line in Figs 1 and 2. Our best-fitting model describes the observed timescale distribution well, but we found there remains a small possible excess of events with timescales 0.5 < tE < 1 d. If we assume that they can be attributed to Jupiter-mass lenses11 (Mlens = 10−3 M(cid:12)), the maximum-likelihood models predict their frequency of 0.05 per main-sequence star with a 68% confidence interval of [0, 0.12] planets per star. The 95% confidence limit is 0.25 Jupiter-mass planets per star. These results agree with upper limits on the frequency of Jupiter-mass planets inferred from direct imaging surveys19, 20, which suggests that almost the entire possible excess of events with timescales 0.5 < tE < 1 d can be attributed to planets on wide orbits21. The timescales of six events passing our criteria for high-quality events are shorter than 0.5 4 day and these events last less that one night (Fig. 3). We carefully checked CCD images by eye to ensure that these brightenings are real, which rules out problems such as photometry artefacts or asteroids. We also analysed historical light curves for these events; four of the six have been observed by the OGLE survey for 20 years and we did not find any evidence for other outbursts in archival data. Nevertheless, because these events were so short and the light curves were not fully covered, we cannot rule out the possibility that some of them might be flaring stars (especially BLG512.18.22725 and BLG500.10.140417). The best-fitting microlensing models of six short events constrain their Einstein timescales in the range 0.1 < tE < 0.4 d (Extended Data Table 1). Such short events should be caused by Earth- and super-Earth-mass objects, provided that they have kinematics that are similar to the brown dwarf, stellar and remnant lenses. They might be gravitationally unbound to any star or located at wide orbits (at least several astronomical units from the host star), given no signs of binarity in their light curves. Because the number of ultrashort events is very small and their nature is uncertain, we do not attempt to model their mass function. However, a mere detection of such ultra-hort events means that Earth-mass lenses must be very common. If we assume that 5 M⊕-mass planets are five times more common than main-sequence stars, the expected number of ultrashort microlensing events is 2.2. For a more realistic mass function in which Earth-mass planets12 are five times more common than main-sequence stars, the expected number of detections is 25% smaller. According to planet formation theories, most Earth- and super-Earth-mass planets should form at relatively small orbital separations (< 10 au)22. The most likely sources of wide-orbit and free-floating Earth-mass planets are dynamical interactions in young multi-planet systems12, 23, 24. Other mechanisms (including ejections from multiple-star systems, stellar fly-bys, interactions in stellar clusters, and post-main-sequence evolution of the host star(s)) have also been proposed3. Although these processes are unlikely to produce a sizable population of Jupiter-mass free-floating planets, Earth-mass planets can be scattered and ejected much more efficiently. 5 Thanks to the superb photometry quality and the possibility of continuous observations dur- ing approximately 100-day-long windows, future space-based missions, such as WFIRST25 and Euclid26, will have the potential to explore the population of free-floating Earth-mass planets in more detail. References. 1. Rasio, F. A. & Ford, E. B. Dynamical instabilities and the formation of extrasolar planetary systems. Science 274, 954-956 (1996). 2. Weidenschilling, S. J. & Marzari, F. Gravitational scattering as a possible origin for giant planets at small stellar distances. Nature 384, 619-621 (1996). 3. Veras, D. & Raymond, S. N. Planet-planet scattering alone cannot explain the free-floating planet population. Mon. Not. R. Astron. Soc. 421, L117-L121 (2012). 4. Luhman, K. L. The formation and early evolution of low-mass stars and brown dwarfs. Annu. Rev. Astron. Astrophys. 50, 65-106 (2012). 5. Zapatero Osorio, M. R. et al. Discovery of Young, Isolated Planetary Mass Objects in the σ Orionis Star Cluster. Science 290, 103-107 (2000). 6. Liu, M. C. et al. The Extremely Red, Young L Dwarf PSO J318.5338-22.8603: A Free-floating Planetary-mass Analog to Directly Imaged Young Gas-giant Planets. Astrophys. J. 777, L20 (2013). 7. Dupuy, T. J. & Kraus, A. L. Distances, Luminosities, and Temperatures of the Coldest Known Substellar Objects. Science 341, 1492-1495 (2013). 8. Scholz, A. et al. Substellar Objects in Nearby Young Clusters (SONYC). VI. The Planetary- mass Domain of NGC 1333. Astrophys. J. 756, 24 (2012). 6 9. Pena Ram´ırez, K., B´ejar, V. J. S., Zapatero Osorio, M. R., Petr-Gotzens, M. G. & Mart´ın, E. L. New Isolated Planetary-mass Objects and the Stellar and Substellar Mass Function of the σ Orionis Cluster. Astrophys. J. 754, 30 (2012). 10. Muzi´c, K., Scholz, A., Geers, V. C. & Jayawardhana, R. Substellar Objects in Nearby Young Clusters (SONYC) IX: The Planetary-Mass Domain of Chamaeleon-I and Updated Mass Func- tion in Lupus-3. Astrophys. J. 810, 159 (2015). 11. Sumi, T. et al. Unbound or distant planetary mass population detected by gravitational mi- crolensing. Nature 473, 349-352 (2011). 12. Ma, S., Mao, S., Ida, S., Zhu, W. & Lin, D. N. C. Free-floating planets from core accretion theory: microlensing predictions. Mon. Not. R. Astron. Soc. 461, L107-L111 (2016). 13. Udalski, A., Szyma´nski, M. K. & Szyma´nski, G. OGLE-IV: Fourth Phase of the Optical Grav- itational Lensing Experiment. Acta Astron. 65, 1-38 (2015). 14. Wo´zniak, P. & Paczy´nski, B. Microlensing of Blended Stellar Images. Astrophys. J. 487, 55-60 (1997). 15. Bennett, D. P., Sumi, T., Bond, I. A., et al. Planetary and Other Short Binary Microlensing Events from the MOA Short-event Analysis. Astrophys. J. 757, 119 (2012). 16. Calchi Novati, S., de Luca, F., Jetzer, P., Mancini, L. & Scarpetta, G. Microlensing constraints on the Galactic bulge initial mass function. Astron. Astrophys. 480, 723-733 (2008). 17. Han, C. & Gould, A. The Mass Spectrum of MACHOs from Parallax Measurements. Astro- phys. J. 447, 53 (1995). 18. Han, C. & Gould, A. Stellar Contribution to the Galactic Bulge Microlensing Optical Depth. Astrophys. J. 592, 172-175 (2003). 19. Lafreni`ere, D. et al. The Gemini Deep Planet Survey. Astrophys. J. 670, 1367-1390 (2007). 7 20. Bowler, B. P., Liu, M. C., Shkolnik, E. L. & Tamura, M. Planets around Low-mass Stars (PALMS). IV. The Outer Architecture of M Dwarf Planetary Systems. Astrophys. J. Suppl. Ser. 216, 7 (2015). 21. Clanton, C. & Gaudi, B. S. Constraining the Frequency of Free-floating Planets from a Syn- thesis of Microlensing, Radial Velocity, and Direct Imaging Survey Results. Astrophys. J. 834, 46 (2017). 22. Ida, S., Lin., D. N. C., & Nagasawa, M. Toward a deterministic model of planetary formation. VII. Eccentricity distribution of gas giants. Astrophys. J. 775, 42 (2013). 23. Pfyffer, S., Alibert, Y., Benz, W. & Swoboda, D. Theoretical models of planetary system formation. II. Post-formation evolution. Astron. Astrophys. 579, A37 (2015). 24. Barclay, T., Quintana, E. V., Raymond, S. N. & Penny, M. T. The demographics of rocky free-floating planets and their detectability by WFIRST. Astrophys. J. 841, 86 (2017). 25. Spergel, D. et al.Wide-Field InfrarRed Survey Telescope-Astrophysics Focused Telescope As- sets WFIRST-AFTA 2015 Report, arXiv:1503.03757 (2015). 26. Penny, M. T. et al. ExELS: an exoplanet legacy science proposal for the ESA Euclid mission - I. Cold exoplanets. Mon. Not. R. Astron. Soc. 434, 2-22 (2013). 27. Mao, S. & Paczynski, B. Mass Determination with Gravitational Microlensing. Astrophys. J. 473, 57 (1996). Acknowledgements We thank M. Kubiak and G. Pietrzy´nski, former members of the OGLE team, for their contribution to the collection of the OGLE photometric data over the past years. The OGLE project has received funding from the National Science Center, Poland through grant MAESTRO 2014/14/A/ST9/00121 to A.U. 8 Author Contributions P.M. analysed and interpreted the data, and prepared the manuscript. A.U. initiated the project, reduced the data, and conducted detection efficiency simulations. All authors collected the OGLE photometric observations, reviewed, discussed and commented on the present results and on the manuscript. Author Information Reprints and permissions information is available at www.nature.com/reprints. The authors declare no competing financial interests. Readers are welcome to comment on the online version of the paper. Correspondence and requests for materials should be addressed to P.M. ([email protected]). 9 Figure 1 Observed distribution of timescales of 2,617 high-quality microlensing events discovered by OGLE in 2010–2015. The purple line is the best-fitting model. The dotted line constrains the 95% confidence limit on the number of wide-orbit or unbound Jupiter-mass planets of 0.25 planets per star. The dashed red line is the best-fitting model from ref. 11 predicting almost two Jupiter-mass free-floating planets per star. According to that model we should find 64 events with 0.3 < tE < 1.8 d, but only 21 were observed (the discrepancy is even larger for events with 0.3 < tE < 1.3 d, where 6 events were found out of 42 expected). We detected six possible ultrashort-timescale events (tE < 0.5 d), which may be due to Earth-mass free-floating planets (grey histogram). Solid (dotted) green lines mark the expected microlensing signal assuming 5 M⊕ planets five (ten) times more frequent than stars. Error bars are the 1σ Poisson uncertainties on the counts of the 10 0.11101001,000tE(days)110100NumberofeventsperbinBest-fittingmodel95%conf.limitSumietal.(2011)10Earthsperstar5Earthsperstar−10123logtE number of events observed in a given tE bin. Figure 2 Distribution of event timescales corrected for the detection efficiency. This distribution, at short timescales, can be well approximated as a power-law with a slope of +3, consistent with theoretical expectations27. There remains a small possible excess of events with timescales 0.5 < tE < 1 d. If they were caused by the Jupiter-mass lenses, the best-fitting models predict their frequency of 0.05 Jupiter-mass planets per star with a 95% confidence limit of 0.25 planets per star (dotted purple line). All symbols are the same as in Fig. 1. Error bars are the 1σ Poisson uncertainties on the counts of the number of events observed in a given tE bin. 11 0.11101001,000tE(days)101102103NumberofeventsperbinBest-fittingmodel95%conf.limitSumietal.(2011)10Earthsperstar5Earthsperstar−10123logtE Figure 3 Light curves of ultrashort microlensing event candidates. The left panels show a close-up of the light curve at the event and the right panels show 5.5-year long light curves from OGLE-IV. Some of those events have been observed by OGLE for 20 years with no trace of other variability, but we nevertheless cannot exclude the possibility that some of them may be flaring stars. The shortest-timescale events are not well covered by observations and it is difficult, if not impossible, to either prove or disprove their nature 12 as free-floating planets. The detection efficiency at these timescales is very low, meaning that a very few detections imply the existence of a large population of Earth-mass free- floating or wide-orbit planets. Future space-based missions, like WFIRST and Euclid, will enable the exploration of these short events in more detail. Error bars represent 1σ uncertainties. HJD, Heliocentric Julian date. 13 Methods Data. All data presented in this paper were collected as part of the OGLE-IV sky survey13 during the years 2010-2015. The survey uses the 1.3-m Warsaw Telescope, located at Las Campanas Observatory, Chile. The observatory is operated by the Carnegie Institution for Science. The telescope is equipped with a mosaic, 32-chip CCD camera covering a field of view of 1.4 sqaure degrees with a pixel scale of 0.26(cid:48)(cid:48) per pixel. All objects analysed are located within nine OGLE fields, observed with a cadence of either 20 min (BLG501, BLG505, and BLG512) or 60 min (BLG500, BLG504, BLG506, BLG511, BLG534, and BLG611), covering in total 12.6 square degrees. We analyzed data collected between 2010 June 29 and 2015 November 8, that is, five and a half Galactic bulge observing seasons. Light curves consist of 4,500 – 12,000 data points, depending on the field, which gives a total of 380 billion photometric measurements. All analyzed data were taken through the I-band filter. Basic information about the fields analyzed is presented in Extended Data Table 2. OGLE photometric pipeline is based on the difference image analysis method (DIA)28, 29. For each field, a reference image is constructed by stacking several highest-quality and seeing frames. This reference image is then subtracted from incoming frames and the photometry is performed on subtracted images. Variable and transient objects that are detected on subtracted images are stored in two databases. The "standard" database consists of all stellar-like objects detected on the reference frame, whereas "new" objects (those that do not correlate with any identified stars) are stored separately; see the description of the OGLE photometric pipeline29, 30. Event selection. We analysed 50 million light curves, from all the objects from the "standard" database. We began our analysis by correcting photometric uncertainties31 and transforming mag- nitudes into flux. It is known that uncertainties returned by the DIA are underestimated and ref. 31 provides an algorithm for their correction, so that these uncertainties now reflect the real observa- tional scatter in the data. The selection criteria for high-quality microlensing events are summa- rized in Extended Data Table 3. 14 Cut 1. We placed a 360-day moving window on each light curve and measured the baseline flux Fbase and its dispersion σbase using data points outside the window (after rejecting 5σ outliers such as cosmic ray hits). We required χ2 out/d.o.f. ≤ 2.0, where d.o.f. are degrees of freedom, outside the window, so we could reject most of the variable stars. Some genuine microlensing the baseline flux. For each bump we calculated χ3+ = (cid:80) events with variable baseline or those longer than one year may have not passed this criterion. We defined a bump as a brightening with at least three consecutive points at least 3σbase above i(Fi − Fbase)/σi (i is the index within a bump) and nDIA, the number of detections on subtracted images. We required χ3+ ≥ 32 and nDIA ≥ 3 to pass this cut. (We note that with the current data we were able to set a lower threshold than in ref. 11, who used χ3+ ≥ 80). The introduction of the cut on nDIA allowed us to eliminate contamination from asteroids, photometry artefacts, and "ghost" microlensing events, which are stars affected by real variability of neighbouring stars32. Cut 2. Cut 1 criteria were insufficient to remove all artefacts. For example, reflections within the telescope might cause spurious, short brightenings of neighboring stars correlated in time. Reflections were especially troublesome near the edges of CCD detectors #1, #7, #8, #16, #17, #25, #26, and #32 of the OGLE-IV mosaic camera, located at the edges of the telescope field of view13. To quantify the concurrence of bumps, we defined the similarity of two bumps as s = N1/N2, where N1 is the number of individual frames when both bumps were detected on subtracted images and N2 is the number of frames when at least one bump was detected. We calculated similarities for all possible pairs of bumps shorter than five days and then rejected objects with s ≥ 0.4. This threshold value was chosen after visual inspection of light curves and images of possible short events. It allowed us to reject over 95% of artefacts, while removing none of the genuine microlensing events from the sample. A number of stars that passed cut 1 criteria were OGLE small amplitude red giants (OSARGs)33 which are red-giant variable stars showing low-amplitude (< 0.13 mag in the I band) pulsations with (frequently multiple) periods in the range 10 < P < 100 d. Some pulsation cycles in OS- 15 ARGs might have slightly higher amplitudes so they were detected by our algorithm as potential microlensing events. We therefore rejected all objects with a bump amplitude A ≤ 0.1 mag, so only a few genuine microlensing events were discarded in this step. The remaining OSARGs were easily rejected in the next step, because the microlensing light curve fit yielded nonphysical parameters. Finally, we rejected all objects with more than one bump in the light curve. These were mostly dwarf novae and some remaining photometry artefacts. Twenty-nine genuine microlensing events were also rejected, most of them binary source or binary lens events, and some microlensing events with variable baseline. Cut 3. For the remaining 11,989 event candidates, we fitted the microlensing point-source point-lens model. The lensing model has three parameters: the time t0 and projected separation u0 (in Einstein radius units) between the lens and the source during the closest approach, and the Einstein radius crossing time tE. The source flux Fs and the blend flux Fb were found analytically using the least-squares method. We also calculated the four-parameter fits, where the blend flux was set to zero, Fb = 0. We performed the initial fit using the simplex algorithm using the data from a 360-day window centred on the event and later refined the parameters using all available data. fit,2tE fit for the entire dataset, χ2 fit,tE for t − t0 < 2tE, and χ2 We calculated a number of goodness-of-fit statistics. χ2 for fit,k for t − t0 < k (where k = 1 or k = 5 days). We t − t0 < tE, χ2 removed 4σ outliers provided that adjacent datapoints are within 1σ from the best-fitting model and ti±1 − ti < 1 day. We required χ2/d.o.f. ≤ 2.0, which removes the majority of non-standard microlensing events (finite source, parallax, binary) in addition to non-microlensing events. We allowed for some amount of negative blending, that is, the blend flux Fb > −F0 was allowed, where F0 = 0.251 is the flux corresponding to an 19.5-mag star (here F = 1 corresponds to an 18-mag star). If Fb < −F0 and the four-parameter model was marginally worse (∆χ2 < 4) than the five-parameter model, we chose the four-parameter model. Usually, a high negative blending 16 indicates that the single lensing model has been fitted to a non-microlensing event (like a dwarf nova outburst, OSARG, or stellar flare). However, a small amount of negative blending does not necessarily mean that the model is unphysical. The background (mainly unresolved main-sequence stars) in crowded fields of the Galactic bulge is not uniform and if the source happens to be located in a lower-density region, the blend flux might be negative. The issue of negative blending is discussed by refs 34–36. We checked that our prior on the negative blending has no impact on the final event timescale distribution (which remains the same after choosing F0 = 0.1, that is, the flux corresponding to a 20.5-mag star). We also required at least nr ≥ 2 datapoints on the rising part of the light curve (t0 − tE < t < t0) and at least nd ≥ 2 datapoints on the descending branch (t0 < t < t0 + tE). If nd < 2, we required nr ≥ 4. These cuts allowed us to eliminate contamination from flaring stars, which can rise very steeply37 (within minutes), but fade slowly (on a timescale of hours). If the rising part of the light curve is not sufficiently sampled, a flare might be mistaken for a very short microlensing event. Our image-level simulations (see below) showed that we were unable to robustly measure the true timescale of an event if the event is faint and the blending is high (fs < 0.1, that is, less than 10% of baseline flux comes from the source). Therefore, to ensure that the final results are sound we did not include events with blending parameter fs < 0.1. The inclusion of highly- blended events had little effect on the final results, although we found an increased number of long-timescale events (tE > 100 d). The purity of our sample is almost 100%. Over 90% of microlensing events detected in the real-time by the OGLE Early Warning System30 passed our "cut 2" criteria. We detected additional 20–30% events (depending on the field) compared to Early Warning System detections. The final distribution of timescales of detected microlensing events is shown in Fig. 1. Extended Data Table 4 presents the number of events detected in individual fields and timescale bins. 17 Detection Efficiency. To calculate the event detection efficiency, we carried out extensive image- level simulations in which we injected artificial microlensing events into real OGLE frames using the PSF derived from neighboring stars. In each iteration we simulated 5,000 events per CCD detector, so the star density did not increase much (by 5–10%). We carried out six iterations for each field, so in total 8.6 million of events were simulated in all fields. Parameters t0 and u0 were drawn from uniform distributions: 0.0 ≤ u0 < 1.5 and 2455377 ≤ t0 < 2457388. Einstein timescales were drawn from a log-uniform distribution −1.0 ≤ log tE < 2.5. Sources were taken from the range 14 ≤ Is < 22 mag from the luminosity function of each subfield, which was created as follows. We constructed a very deep luminosity function for the subfield BLG513.12, which was observed both by the OGLE-IV survey and the Hubble Space Telescope38. The OGLE-IV luminosity function and the Hubble Space Telescope luminosity func- tion overlap in the range 16 < I < 18 mag (Extended Data Fig. 1). This deep luminosity function was used as a template to generate artificial microlensing events in other fields, after shifting it so that the centroid of the red clump matched the observed centroid. We therefore took into account variable bulge geometry and reddening. If there was evidence for differential reddening, we di- vided subfields into smaller parts. There were a few subfields (7% of the total analysed area) where we were not able to detect the red clump owing to extremely high extinction; these were omitted from the final calculations (we detected only a negligible number of 48 microlensing events in these fields). For the simulated events we applied exactly the same selection criteria as for the real events (Extended Data Table 3). The detection efficiency curves for all analysed fields are shown in Extended Data Fig. 2 and listed in Extended Data Tab. 5. We note that detection efficiency for events with tE = 2 d is very high, up to 53% of the maximum efficiency for field BLG512. Efficiencies for fields observed with 20-min and 60-min cadence are very similar, except for the shortest events with tE < 0.5 day. In general, we found that detection efficiencies are most sensitive to crowding and interstellar reddening toward the given field (fields with higher reddening and 18 higher crowding have lower efficiencies). We note that events were simulated using a standard point-lens point-source model. Higher-order effects, like the parallax (causing deviations in the light curve induced by the Earth's motion39), were not included and so detection efficiencies for long events (tE ≥ 100 d) may be slightly overestimated. Similarly, we did not include the finite source effect, which may reduce our detection efficiency for the shortest events (tE ∼ 0.1 d), when the Einstein ring size becomes similar to the source star radius40. Parameter recovery. We also used our simulations to ensure that there is no systematic difference between measured and real timescales. In Extended Data Fig. 3 we plot timescales for simulated events passing all criteria from Extended Data Table 3. We found there is no systematic bias in measured timescales, unless events were faint and highly blended. This effect was predicted by ref. 14, where it was found theoretically that in such cases the event timescale, impact parameter and blending parameter may be severely correlated, because information on the event timescale comes mostly from wings of the light curve that can be more easily affected by the photometric noise. In Extended Data Fig. 4a we show the ratio between measured and "real" (simulated) timescale tE,out/tE,in versus the blending parameter fs = Fs/(Fs + Fb). It is clear that timescales of highly blended and faint events are not well measured and systematically overestimated. A similar effect was also noticed in the earlier work11, where it was found that tE,in was systematically about 5% smaller than tE,out regardless of tE. Strong correlations between blending, impact parameter, and event timescale may also lead to the incorrect determination of parameters. For example, one of short events reported by ref. 11, MOA-ip-1, has incorrectly measured timescale. The best-fitting model with tE = 8.2+8.1−3.6 d is better by ∆χ2 = 9 than the model presented in the original paper (tE = 0.73 ± 0.08 d). To be conservative, we decided not to include highly-blended events (fs < 0.1) in our final sample of high-quality events. Thanks to this selection cut, there is almost no bias in the measured timescales (see Extended Data Figs. 3 and 4b). 19 Modeling Timescale Distribution. The actual timescale distribution depends on the distribution and kinematics of lenses and sources as well as the underlying mass function27, 41, 42. The timescale distribution can be computed from a multi-dimensional integral42, 43: (cid:90) f (tE) ∝ ρ(DS)ρ(DL)RE(DL, DS, M )Φ(M ) × vrelf (vrel)δ(tE − RE vrel )dDLdDSdvreldM, where ρ(D) is the distribution of lenses and sources along the line-of-sight, RE = θEDL the Einstein radius, vrel is the lens-source relative velocity projected onto the plane of the sky, and Φ(M ) is the mass function. We expect the timescale distribution to have power-law tails with slopes of +3 and −3 at short and long timescales, respectively27, 44. To compare the measured distribution of Einstein timescales with models, we maximized the following log-likelihood function: lnL = (cid:88) i ln p(tE,i), where p(tE) = pmodel(tE)ε(tE) is the normalized predicted timescale distribution, which serves as our likelihood function. Here pmodel(tE) is the timescale distribution from the Galactic model and ε(tE) is the detection efficiency in a given field. The summation was performed over all events. We adopted a standard Galactic model17, 18, which incorporates the boxy-shaped bulge model45 and the double exponential model of the Galactic disk46. Mass function. A detailed modeling of the initial mass function (IMF) would require population synthesis calculations, in addition to more sophisticated Galactic models, which is beyond the scope of this work. However, we can obtain useful constrains on slopes of the IMF using a simple model. Here we followed the approach of ref. 47 and we assumed that all stars with initial masses 1 < M/M(cid:12) ≤ 8 evolved into white dwarfs following the empirical initial-final mass relation for white dwarfs48 Mfinal = 0.339 + 0.129 Minit. Masses of neutron stars (with initial masses in the range 8 < M/M(cid:12) ≤ 20) peak around 1.33 M(cid:12) with a 68% confidence interval of (1.21, 1.43) M(cid:12) (ref. 49), while for black holes we assumed a Gaussian distribution at 7.8 ± 1.2 M(cid:12) (ref. 50). 20 We fitted the following initial mass function:  Φ(M ) = . 0.01M(cid:12) ≤ M < 0.08M(cid:12) 0.08M(cid:12) ≤ M < Mbreak a1M−αbd a2M−αms a3M−2.0 M ≥ Mbreak We allowed αbd and αms to vary, but we assumed a fixed IMF slope of −2.0 above M > Mbreak = 0.5M(cid:12) (ref. 51), because our experiment was designed to analyze the low-mass end of the IMF. We also considered models with Mbreak = 0.7M(cid:12) and models with binary fraction fbin (cid:54)= 0, where we assumed a flat mass ratio distribution52 f (q) = 1 in a range 0 ≤ q ≤ 1. We conducted modelling using events with tE > 0.5 and tE > 2.0 days and in both cases we obtained virtually identical results. Constraints on slopes of the IMF are shown in Extended Data Fig. 5. In general, we found that models with non-zero binary fraction describe the event timescale distribution better than models with fbin = 0. The standard IMF53 with fbin = 0 does not describe the entire timescale distribution well, especially at long timescales tE > 50 d, which has already been noted54. This may indicate that the current Galactic model underpredicts the number of long- timescale events, or the mass function of remnants (especially black holes) is underestimated, or remnants have distinct kinematics from brown dwarf and stellar lenses. The discrepancy can be also explained, if we assume that some fraction of lenses (fbin) are binary systems. Our models with fbin = 0.4 are substantially better than with fbin = 0.0 (with improvement in log-likelihood ∆χ2 = 2.0(lnLmax,1 − lnLmax,2) = 18.6). For the best-fitting models αbd ≈ 0.8 and αms ≈ 1.3 with 3σ confidence intervals: 0.2 < αbd < 1.3 and 1.1 < αms < 1.5. This corresponds to 0.90± 0.05 (1σ) brown dwarfs per main-sequence star. Ref. 11 obtained a slightly lower IMF slope in the brown dwarf regime of αbd = 0.49+0.24−0.27, but they used fixed αms = 1.3 and fbin = 0 (their slope αbd is in fact consistent with our models from Extended Data Fig. 5a for fixed αms). The IMF slope derived in the stellar regime is consistent with the "canonical"53 value of −1.3. Observations of brown dwarfs in open clusters and star-forming regions indicate αbd ≈ 0.6 − 0.7 (ref. 55 and references therein) and our models are consistent with those values. On the 21 other hand, censuses of nearby field brown dwarfs tend to prefer lower slopes. Ref. 56 found a 60% confidence interval of αbd ≈ 0.3 ± 0.6 and other studies support αbd ∼ 0 (ref. 55). However, mass function measurements for isolated field brown dwarfs are affected by difficulties in measuring their ages, distances, and masses. Planetary mass function. To explain the excess of short events, ref. 11 modelled their event timescale distribution using a stellar IMF with αbd = 0.5, αms = 1.3, and Mbreak = 0.7M(cid:12) with additional planetary component, approximated as a delta function at M = 10−3 M(cid:12). That model is shown in Figs 1 and 2 as dashed red line. According to that model we should find 64 events with 0.3 < tE < 1.8 d, but only 21 were observed (the discrepancy is even larger for events with 0.3 < tE < 1.3 d, where 6 events were found out of 42 expected). Moreover, model of ref. 11 systematically underpredicts the number of long-timescale events (because of its very low sensitivity to long events, tE > 100 d, they found only five events in this range). Our best-fitting model describes the observed timescale distribution well, but there remains a small possible excess of events with timescales 0.5 < tE < 1 d (Figs 1 and 2). If we assume, following ref. 11, they are due to Jupiter-mass lenses (Mlens = 10−3 M(cid:12)), the best-fitting models predict their frequency of 0.05 Jupiter-mass planet per star with 68% confidence interval of [0, 0.12] planets per star. The 95% confidence limit is 0.25 Jupiter-mass planet per star. Our results agree with upper limits on the frequency of Jovian-mass planets inferred from direct imaging surveys 19, 57. For example, a high-contrast adaptive imaging search20 for giant planets around nearby M- dwarf stars did not find any planets, providing very strong upper limits (at the 95% confidence limit) of 10-16% (depending on the model) for planets of between 1 and 13 Jupiter masses, at a distance of approximately 10 − 100 AU. This suggests that almost the entire possible excess of events with timescales 0.5 < tE < 1 d can be attributed to planets on wide orbits. Code availability. We have opted not to make the event detection and simulation codes publicly available, because they were designed to work with internal photometric databases. The code for the modelling of the timescale distribution is available from the corresponding author upon 22 reasonable request. Data availability. The data that support the findings of this study are available from the corre- sponding author upon reasonable request. 28. Alard, C. & Lupton, R. H. A Method for Optimal Image Subtraction. Astrophys. J. 503, 325- 331 (1998). 29. Wo´zniak, P. R. Difference Image Analysis of the OGLE-II Bulge Data. I. The Method. Acta Astron. 50, 421-450 (2000). 30. Udalski, A. The Optical Gravitational Lensing Experiment. Real Time Data Analysis Systems in the OGLE-III Survey. Acta Astron. 53, 291-305 (2003). 31. Skowron, J. et al. Analysis of Photometric Uncertainties in the OGLE-IV Galactic Bulge Mi- crolensing Survey Data. Acta Astron. 66, 1-14 (2016). 32. Wyrzykowski, Ł. et al. OGLE-III Microlensing Events and the Structure of the Galactic Bulge. Astrophys. J. Suppl. Ser. 216, 12 (2015). 33. Wray, J. J., Eyer, L. & Paczy´nski, B. OGLE small-amplitude variables in the Galactic bar. Mon. Not. R. Astron. Soc. 349, 1059-1068 (2004). 34. Park, B.-G. et al. MOA-2003-BLG-37: A bulge jerk-parallax microlens degeneracy. Astro- phys. J. 609, 166-172 (2004). 35. Jiang, G. et al. OGLE-2003-BLG-238: Microlensing mass estimate for an isolated star. Astro- phys. J. 617, 1307-1315 (2004). 36. Smith, M. C., Wo´zniak, P., Mao, S. & Sumi, T. Blending in gravitational microlensing ex- periments: source confusion and related systematics. Mon. Not. R. Astron. Soc. 380, 805-818 (2007). 23 37. Hawley, S. L. et al. Kepler Flares. I. Active and Inactive M Dwarfs. Astrophys. J. 797, 121 (2014). 38. Holtzman, J. A. et al. The Luminosity Function and Initial Mass Function in the Galactic Bulge. Astron. J. 115, 1946-1957 (1998). 39. Gould, A. Extending the MACHO search to about 106 solar masses. Astrophys. J. 392, 442 (1992). 40. Bennett, D. P. & Rhie, S. H. Detecting Earth-mass planets with gravitational microlensing. Astrophys. J. 472, 660-664 (1996). 41. Kiraga, M. & Paczynski, B. Gravitational microlensing of the Galactic bulge stars. Astrophys. J. 430, L101-L104 (1994). 42. Han, C. & Gould, A. Statistical Determination of the MACHO Mass Spectrum. Astrophys. J. 467, 540 (1996). 43. Bissantz, N., Debattista, V. P. & Gerhard, O. Large-Scale Model of the Milky Way: Stellar Kinematics and the Microlensing Event Timescale Distribution in the Galactic Bulge. Astro- phys. J. 601, L155-L158 (2004). 44. Wood, A. & Mao, S. Optical depths and time-scale distributions in Galactic microlensing. Mon. Not. R. Astron. Soc. 362, 945-951 (2005). 45. Dwek, E. et al. Morphology, near-infrared luminosity, and mass of the Galactic bulge from COBE DIRBE observations. Astrophys. J. 443, 716-730 (1995). 46. Zheng, Z., Flynn, C., Gould, A., Bahcall, J. N. & Salim, S. M Dwarfs from Hubble Space Telescope Star Counts. Astrophys. J. 555, 393-404 (2001). 47. Gould, A. Measuring the Remnant Mass Function of the Galactic Bulge. Astrophys. J. 535, 928-931 (2000). 24 48. Williams, K. A., Bolte, M. & Koester, D. Probing the Lower Mass Limit for Supernova Pro- genitors and the High-Mass End of the Initial-Final Mass Relation from White Dwarfs in the Open Cluster M35 (NGC 2168). Astrophys. J. 693, 355-369 (2009). 49. Kiziltan, B., Kottas, A., De Yoreo, M. & Thorsett, S. E. The Neutron Star Mass Distribution. Astrophys. J. 778, 66 (2013). 50. Ozel, F., Psaltis, D., Narayan, R. & McClintock, J. E. The Black Hole Mass Distribution in the Galaxy. Astrophys. J. 725, 1918-1927 (2010). 51. Zoccali, M. et al. The Initial Mass Function of the Galactic Bulge down to 0.15 Msolar. Astrophys. J. 530, 418-428 (2000). 52. Belczynski, K. et al. Compact Object Modeling with the StarTrack Population Synthesis Code. Astrophys. J. Suppl. Ser. 174, 223-260 (2008). 53. Kroupa, P. On the variation of the initial mass function. Mon. Not. R. Astron. Soc. 322, 231-246 (2001). 54. Wegg, C., Gerhard, O. & Portail, M. MOA-II Galactic microlensing constraints: the inner Milky Way has a low dark matter fraction and a near maximal disc. Mon. Not. R. Astron. Soc. 463, 557-570 (2016). 55. Alves de Oliveira, C. The low mass end of the IMF. Mem. Soc. Astron. Ital. 84, 905 (2013). 56. Allen, P. R., Koerner, D. W., Reid, I. N. & Trilling, D. E. The Substellar Mass Function: A Bayesian Approach. Astrophys. J. 625, 385-397 (2005). 57. Quanz, S. P., Lafreni`ere, D., Meyer, M. R., Reggiani, M. M. & Buenzli, E. Direct imag- ing constraints on planet populations detected by microlensing. Astron. Astrophys. 541, A133 (2012). 25 Extended Data Figure 1 Galactic bulge luminosity function used for simulations. a, Deep luminosity function (LF) for subfield BLG513.12, which was observed both by the OGLE-IV survey and the Hubble Space Telescope (HST)38. Both LFs overlap in the range 16 < I < 18 mag. This deep LF was used as a template to generate artificial microlensing events in analysed fields, after shifting to match the red clump's centroid in a given field. b, Comparison between the observed LF for subfield BLG512.32 and the simulated LF. 26 12141618202224 magnitude101100101102103104stars / arcmin / magHST LF (Holtzman+ 1998)OGLE-IV LFa121416182022 magnitude101100101102103104stars / arcmin / magOGLE-IV LF for field BLG512.32Simulations (LF shifted by 0.50 mag)b Extended Data Figure 2 Detection efficiency curves. Detection efficiencies as a function of the Einstein timescale tE for all analysed fields (averages for all subfields in the given field). Fields BLG501, BLG505, and BLG512 were observed with a 20-min ca- dence, and the remaining fields with a 60-min cadence. Error bars are the 1σ Poisson uncertainties on the counts of the number of simulated events in a given tE bin. 27 1.00.50.00.51.01.52.02.5(/)103102101Detection efficiencyBLG500BLG501BLG504BLG505BLG506BLG511BLG512BLG534BLG611 Extended Data Figure 3 Comparison between measured Einstein timescales tE,out and "real" (simulated) timescales tE,in for simulated events. Only events passing se- lection criteria from Extended Data Table 3 (including the cut on the blending parameter fs > 0.1) are shown. Note that the colour scale is logarithmic. There is no systematic offset between measured and real timescales. 28 0.1110100, (days)0.1110100, (days)110100Number density Extended Data Figure 4 Comparison between measured and "real" (simulated) parameters. a, Ratio between the measured Einstein timescale tE,out and "real" (simu- lated) timescale tE,in for simulated events versus the blending parameter fs = Fs/(Fs +Fb). Timescales of faint and highly-blended (fs < 0.1) events are not well measured and are biased by a strong degeneration between Einstein timescale, blending and impact pa- rameters. Timescales of events showing a high negative blending (fs > 1.5) are sys- tematically underestimated, but the bias is relatively small and such events comprise a negligible fraction of all events. b, Distributions of tE,out/tE,in for simulated events passing selection criteria from Extended Data Table 3 (including the cut on the blending parame- ter fs > 0.1). Regardless of the timescale, there is no systematic bias between measured and real timescales within 1%. For 90% of simulated events 0.63 < tE,out/tE,in < 1.65. The 29 0.00.51.01.52.02.53.03.54.0,/,0.00.51.01.52.0aAll0.00.51.01.52.02.53.03.54.0,/,,< 0.00.51.01.52.02.53.03.54.0,/,<,< 0.00.51.01.52.02.53.03.54.0,/,,>0.00.51.01.52.0,/,0.00.20.40.60.81.0Fraction of eventsmedian=1.00MAD=0.12bAll0.00.51.01.52.0,/,median=1.01MAD=0.15,< 0.00.51.01.52.0,/,median=1.01MAD=0.13<,< 0.00.51.01.52.0,/,median=1.00MAD=0.12,> MAD is the median absolute deviation from the data's median. Extended Data Figure 5 Constraints on IMF slopes: a, Assuming that all lenses are single; b, assuming binary fraction fbin = 0.4. 30 0.00.51.01.52.00.60.81.01.21.41.61.82.0a01020304050607080901000.00.51.01.52.00.60.81.01.21.41.61.82.0b0102030405060708090100 Star BLG501.31.5900 BLG501.02.127000 BLG500.10.140417 BLG501.26.33361 BLG505.27.114211 BLG512.18.22725 RA 17:50:42.45 17:53:13.44 17:53:16.89 17:54:17.54 17:59:04.18 18:05:25.00 Decl. -29:24:49.7 -30:18:59.6 -28:40:51.4 -29:18:17.0 -28:36:51.7 -28:28:23.9 t0 2456175.648 2457172.692 2456116.554 2455671.124 2457157.780 2456064.921 tE 0.241 0.146 0.246 0.320 0.158 0.128 tE 1σ conf.int. [0.21,0.78] [0.12,0.26] [0.23,0.37] [0.29,0.79] [0.15,0.21] [0.08,0.19] u0 0.772 0.517 0.377 0.471 0.597 0.138 Is 18.20 19.13 19.08 18.04 19.14 20.95 fs 0.97 0.77 1.24 1.11 1.38 0.16 Extended Data Table 1: Best-fitting parameters for ultrashort microlensing event candidates. Is is the source brightness and fs = Fs/(Fs + Fb) is the blending parameter. The inclusion of the finite source effect does not improve χ2 much (typically ∆χ2 = 0.0 − 3.3). Equatorial coordinates are given for the epoch J2000. We also show 1σ confidence intervals for tE. RA, right ascension; Decl., declination. Field BLG500 BLG501 BLG504 BLG505 BLG506 BLG511 BLG512 BLG534 BLG611 RA 17:51:60 17:51:56 17:57:33 17:57:34 17:57:31 18:03:02 18:03:04 17:51:51 17:35:33 Decl. -28:36:35 -29:50:00 -27:59:40 -29:13:15 -30:27:23 -27:22:49 -28:36:39 -31:04:15 -27:09:41 l 0.9999 359.9392 2.1491 1.0870 0.0103 3.2835 2.2154 358.8644 0.3282 b Nstars Nepochs 4708 12117 6435 12083 4712 4595 10268 4652 4526 4.0 5.2 5.8 6.9 5.3 5.5 6.9 4.2 5.0 -1.0293 -1.6400 -1.7747 -2.3890 -2.9974 -2.5219 -3.1355 -2.2547 2.8242 Extended Data Table 2: Basic information about analysed fields. Equatorial coordi- nates are given for the epoch J2000. Nstars is the number of stars in millions and Nepochs is the number of observed frames during 2010–2015. 31 Criteria out/dof ≤ 2.0 χ2 nDIA ≥ 3 χ3+ =(cid:80) i(Fi − Fbase)/σi ≥ 32 s < 0.4 A > 0.1 mag nbump = 1 fit,tE fit,2tE fit/dof ≤ 2.0 χ2 χ2 /dof ≤ 2.0 /dof ≤ 2.0 χ2 fit,1/dof ≤ 2.0 χ2 fit,5/dof ≤ 2.0 χ2 2455377 ≤ t0 ≤ 2457388 u0 ≤ 1 Is ≤ 22.0 nr ≥ 2 if nd ≥ 2 nr ≥ 4 if nd < 2 Fb > −0.251 fs > 0.1 Number 43,158 11,989 Remarks No variability outside the 360-day window centered on the event Centroid of the additional flux coincides with the source star centroid Significance of the bump Rejecting photometry artifacts Rejecting low-amplitude variables Rejecting objects with multiple bumps Fit quality: χ2 for all data χ2 for t − t0 < tE χ2 for t − t0 < 2tE χ2 for t − t0 < 1 day χ2 for t − t0 < 5 days Event peaked between 2010 June 29 and 2015 December 31 The minimum impact parameter The minimum I-band source magnitude Rising and descending parts of the light curve should be sufficiently sampled The maximum negative blend flux, corresponding to I = 19.5 mag star Rejecting highly-blended events 2617 Extended Data Table 3: Selection criteria for high-quality microlensing events. 32 Bin 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 log tE -0.93 -0.79 -0.65 -0.51 -0.37 -0.23 -0.09 0.05 0.19 0.33 0.47 0.61 0.75 0.89 1.03 1.17 1.31 1.45 1.59 1.73 1.87 2.01 2.15 2.29 2.43 BLG500 0 0 1 0 0 0 1 1 0 3 4 10 17 22 25 26 29 23 15 12 7 3 5 0 0 BLG501 1 0 1 1 0 1 0 1 4 5 9 19 40 32 39 35 62 42 39 25 13 9 2 0 0 BLG504 0 0 0 0 0 0 0 1 0 4 7 13 17 24 30 46 38 39 27 20 11 6 3 1 0 BLG505 0 1 0 0 0 0 0 0 2 4 8 28 39 55 78 57 62 53 40 39 20 11 2 1 1 BLG506 0 0 0 0 0 0 1 0 0 1 5 10 19 26 34 44 39 32 32 19 10 6 7 3 2 BLG511 0 0 0 0 0 0 0 0 4 0 8 10 11 19 22 33 30 32 24 21 10 7 1 2 0 BLG512 0 1 0 0 0 0 0 0 1 3 3 13 13 28 40 46 40 41 25 31 12 3 4 3 0 BLG534 0 0 0 0 0 0 0 0 1 5 8 6 17 20 22 24 24 33 20 18 6 2 2 0 0 BLG611 0 0 0 0 0 0 0 0 3 2 5 3 9 20 25 23 38 36 21 11 8 4 2 1 0 Extended Data Table 4: Number of events detected in individual timescale bins. There are 25 bins equally spaced in log tE between −1.0 and 2.5. 33 Bin 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 log tE -0.93 -0.79 -0.65 -0.51 -0.37 -0.23 -0.09 0.05 0.19 0.33 0.47 0.61 0.75 0.89 1.03 1.17 1.31 1.45 1.59 1.73 1.87 2.01 2.15 2.29 2.43 BLG500 0.0016 0.0030 0.0041 0.0061 0.0096 0.0130 0.0194 0.0278 0.0371 0.0447 0.0508 0.0608 0.0658 0.0737 0.0760 0.0858 0.0872 0.0949 0.0964 0.1024 0.1000 0.1029 0.0989 0.0853 0.0618 BLG501 0.0033 0.0071 0.0086 0.0118 0.0144 0.0209 0.0279 0.0365 0.0423 0.0506 0.0557 0.0630 0.0669 0.0746 0.0769 0.0826 0.0831 0.0898 0.0940 0.0973 0.1004 0.0965 0.0928 0.0788 0.0539 BLG504 0.0021 0.0046 0.0061 0.0089 0.0126 0.0176 0.0255 0.0368 0.0461 0.0559 0.0630 0.0701 0.0750 0.0855 0.0910 0.0939 0.1026 0.1099 0.1145 0.1192 0.1207 0.1182 0.1122 0.0979 0.0638 BLG505 0.0045 0.0078 0.0110 0.0139 0.0180 0.0248 0.0343 0.0423 0.0506 0.0571 0.0692 0.0753 0.0816 0.0876 0.0940 0.0950 0.1014 0.1055 0.1108 0.1134 0.1174 0.1124 0.1072 0.0890 0.0538 BLG506 0.0015 0.0043 0.0057 0.0086 0.0120 0.0189 0.0299 0.0396 0.0495 0.0593 0.0675 0.0784 0.0866 0.0937 0.1011 0.1035 0.1079 0.1184 0.1191 0.1264 0.1288 0.1253 0.1148 0.0998 0.0596 BLG511 0.0016 0.0038 0.0057 0.0077 0.0119 0.0181 0.0278 0.0388 0.0486 0.0596 0.0680 0.0758 0.0832 0.0907 0.0949 0.1035 0.1067 0.1151 0.1212 0.1249 0.1254 0.1218 0.1146 0.0914 0.0560 BLG512 0.0039 0.0085 0.0126 0.0144 0.0186 0.0297 0.0381 0.0503 0.0603 0.0705 0.0790 0.0876 0.0940 0.0990 0.1056 0.1113 0.1131 0.1206 0.1286 0.1302 0.1336 0.1331 0.1160 0.0906 0.0548 BLG534 0.0013 0.0033 0.0047 0.0068 0.0095 0.0160 0.0226 0.0335 0.0395 0.0484 0.0592 0.0641 0.0746 0.0772 0.0838 0.0899 0.0913 0.1012 0.1048 0.1105 0.1111 0.1085 0.1029 0.0888 0.0578 BLG611 0.0013 0.0041 0.0053 0.0084 0.0121 0.0180 0.0290 0.0390 0.0525 0.0631 0.0755 0.0863 0.0874 0.1025 0.1107 0.1204 0.1252 0.1361 0.1389 0.1470 0.1525 0.1500 0.1458 0.1295 0.0891 Extended Data Table 5: Detection efficiencies for the analysed fields. There are 25 bins equally spaced in log tE between −1.0 and 2.5. 34
1902.01841
1
1902
2019-02-05T18:21:16
Seasonal evolution of temperatures in Titan's lower stratosphere
[ "astro-ph.EP" ]
The Cassini mission offered us the opportunity to monitor the seasonal evolution of Titan's atmosphere from 2004 to 2017, i.e. half a Titan year. The lower part of the stratosphere (pressures greater than 10 mbar) is a region of particular interest as there are few available temperature measurements, and because its thermal response to the seasonal and meridional insolation variations undergone by Titan remains poorly known. In this study, we measure temperatures in Titan's lower stratosphere between 6 mbar and 25 mbar using Cassini/CIRS spectra covering the whole duration of the mission (from 2004 to 2017) and the whole latitude range. We can thus characterize the meridional distribution of temperatures in Titan's lower stratosphere, and how it evolves from northern winter (2004) to summer solstice (2017). Our measurements show that Titan's lower stratosphere undergoes significant seasonal changes, especially at the South pole, where temperature decreases by 19 K at 15 mbar in 4 years.
astro-ph.EP
astro-ph
Seasonal evolution of temperatures in Titan's lower stratosphere M. Sylvestrea,∗, N. A. Teanbya, J. Vatant d'Olloneb, S. Vinatierc, B. B´ezardc, S. Lebonnoisb, P. G. J. Irwind aSchool of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road, Bristol BS8 1 RJ, UK bLaboratoire de M´et´eorologie Dynamique (LMD/IPSL), Sorbonne Universit´e, ENS, PSL Research University, Ecole Polytechnique, cLESIA, Observatoire de Paris, Universit´e PSL, CNRS, Sorbonne Universit´e, Univ. Paris Diderot, Sorbonne Paris Cit´e, 5 place Jules Universit´e Paris Saclay, CNRS, 4 Place Jussieu, F 75252 Paris Cedex 05, France dAtmospheric, Oceanic, & Planetary Physics, Department of Physics, University of Oxford, Clarendon Laboratory, Parks Road, Oxford Janssen, 92195 Meudon, France OX1 3PU, UK 9 1 0 2 b e F 5 . ] P E h p - o r t s a [ 1 v 1 4 8 1 0 . 2 0 9 1 : v i X r a Abstract The Cassini mission offered us the opportunity to monitor the seasonal evolution of Titan's atmosphere from 2004 to 2017, i.e. half a Titan year. The lower part of the stratosphere (pressures greater than 10 mbar) is a region of particular interest as there are few available temperature measurements, and because its thermal response to the seasonal and meridional insolation variations undergone by Titan remain poorly known. In this study, we measure temperatures in Titan's lower stratosphere between 6 mbar and 25 mbar using Cassini/CIRS spectra covering the whole duration of the mission (from 2004 to 2017) and the whole latitude range. We can thus characterize the meridional distribution of temperatures in Titan's lower stratosphere, and how it evolves from northern winter (2004) to summer solstice (2017). Our measurements show that Titan's lower stratosphere undergoes significant seasonal changes, especially at the South pole, where temperature decreases by 19 K at 15 mbar in 4 years. 1. Introduction Titan has a dense atmosphere, composed of N2 and CH4, and many trace gases such as hydrocarbons (e.g. C2H6, C2H2) and nitriles (e.g. HCN, HC3N) produced by its rich photochemistry. Like Earth, Titan has a stratosphere, located between 50 km (∼ 100 mbar) and 400 km (∼ 0.01 mbar), characterized by the increase of its temperature with altitude because of the absorption of incoming sunlight by methane and hazes. Titan's atmosphere undergoes strong variations of insolation, due to its obliquity (26.7◦) and to the eccentricity of Saturn's orbit around the Sun (0.0565). The Cassini spacecraft monitored Titan's atmosphere during 13 years (from 2004 to 2017), from northern winter Its data are a unique opportunity to summer solstice. to study the seasonal evolution of its stratosphere, especially with mid-IR observations from Cassini/CIRS (Composite InfraRed Spectrometer, Flasar et al. (2004)). They showed that at pressures lower than 5 mbar, the stratosphere exhibits strong seasonal variations of temper- ature and composition related to changes in atmospheric dynamics and radiative processes. For instance, during northern winter (2004-2008), high northern latitudes were enriched in photochemical products such as HCN ∗Corresponding author Preprint submitted to Elsevier or C4H2, while there was a "hot spot" in the upper stratosphere and mesosphere (0.1 - 0.001 mbar, Achter- berg et al. (2008); Coustenis et al. (2007); Teanby et al. (2007); Vinatier et al. (2007)). These observations were interpreted as evidence of subsidence above the North pole during winter, which is a part of the pole-to-pole atmospheric circulation cell predicted for solstices by Titan GCMs (Global Climate Models, Lora et al. (2015); Lebonnois et al. (2012); Newman et al. (2011)). These models also predict that the circulation pattern should reverse around equinoxes, via a transitional state with two equator-to-pole cells. These changes began to affect the South pole in 2010, when measurements showed that pressures inferior to 0.03 mbar exhibited an enrichment in gases such as HCN or C2H2, which propagated downward during autumn, consistent with the apparition of a new circulation cell with subsidence above the South pole (Teanby et al., 2017; Vinatier et al., 2015). Some uncertainties remain about the seasonal evolu- tion of the lower part of the stratosphere, i.e. at pressures from 5 mbar (120 km) to 100 mbar (tropopause, 50 km). Different estimates of radiative timescales have been calculated for this region. In Strobel et al. (2010), the radiative timescales in this region vary from 0.2 Titan years at 5 mbar to 2.5 Titan years at 100 mbar. This means that the lower stratosphere should be the transition zone from parts of the atmosphere which are sensitive to February 6, 2019 seasonal insolation variations, to parts of the atmosphere which are not. In contrast, in the radiative-dynamical model of B´ezard et al. (2018), radiative timescales are between 0.02 Titan year at 5 mbar and 0.26 Titan year at 100 mbar, implying that this whole region should exhibit a response to the seasonal cycle. From northern winter to equinox, CIRS mid-IR observations showed that temperature variations were lower than 5 K between 5 mbar and 10 mbar (Bampasidis et al., 2012; Achterberg et al., 2011). Temporal variations intensified after spring equinox, as Coustenis et al. (2016) measured a cooling by 16 K and an increase in gases abundances at 70◦S from 2010 to 2014, at 10 mbar, associated with the autumn subsidence above the South pole. Sylvestre et al. (2018) showed that this subsidence affects pressure levels as low as 15 mbar as they measured strong enrichments in C2N2, C3H4, and C4H2 at high southern latitudes from 2012 to 2016 with CIRS far-IR observations. However, we have little information on temperatures and their seasonal evolution for pressures greater than 10 mbar. Temperatures from the surface to 0.1 mbar can be measured by Cassini radio-occultations, but the published profiles were measured mainly in 2006 and 2007 (Schinder et al., 2011, 2012), so they provide little information on seasonal variations of temperature. In this study, we analyse all the available far-IR Cassini/CIRS observations to probe temperatures from 6 mbar to 25 mbar, and measure the seasonal variations of lower stratospheric temperatures. As these data were acquired throughout the Cassini mission from 2004 to 2017, and cover the whole latitude range, they provide a unique overview of the thermal evolution of the lower stratosphere from northern winter to summer solstice, and a better understanding of the radiative and dynamical processes at play in this part of Titan's atmosphere. 2. Data analysis 2.1. Observations We measure lower stratospheric temperatures using Cassini/CIRS (Flasar et al., 2004) spectra. CIRS is a thermal infrared spectrometer with three focal planes op- erating in three different spectral domains: 10 - 600 cm−1 (17 - 1000 µm) for FP1, 600 - 1100 cm−1 (9 - 17 µm) for FP3, and 1100 - 1400 cm−1 (7 - 9 µm) for FP4. FP1 has a single circular detector with an angular field of view of 3.9 mrad, which has an approximately Gaussian spatial response with a FWHM of 2.5 mrad. FP3 and FP4 are each composed of a linear array of ten detectors. Each of these detectors has an angular field of view of 0.273 mrad. In this study, we use FP1 far-IR observations, where nadir spectra are measured at a resolution of 0.5 cm−1, in "sit-and-stare" geometry (i.e the FP1 detector probes the same latitude and longitude during the whole duration of the acquisition). In this type of observation, the average spatial field of view is 20◦ in latitude. An acquisition lasts between 1h30 and 4h30, allowing the recording of 100 to 330 spectra. The spectra from the same acquisition are √ averaged together, which increases the S/N by a factor N (where N is the number of spectra). As a result, we lines obtain an average spectrum where the rotational of CH4 (between 70 cm−1 and 170 cm−1) are resolved and can be used to retrieve Titan's lower stratospheric temperature. An example averaged spectrum is shown in Fig. 1. We analysed all the available observations with the characteristics mentioned above. As shown in table 1, this type of nadir far-IR observation has been performed throughout the Cassini mission (from 2004 to 2017), at all latitudes. Hence, the analysis of this dataset enables us to get an overview of Titan's lower stratosphere and its seasonal evolution. Figure 1: Example of average spectrum measured with the FP1 de- tector of Cassini/CIRS (in black) and its fit by NEMESIS (in red). The measured spectrum was obtained after averaging 106 spectra observed at 89◦N in March 2007. The rotational lines of CH4 are used to retrieve stratospheric temperature. The "haystack" feature is visible only at high latitudes during autumn and winter. 2.2. Retrieval method We follow the same method as Sylvestre et al. (2018). We use the portion of the spectrum between 70 cm−1 and 400 cm−1, where the main spectral features are: the ten rotational lines of CH4 (between 70 cm−1 and 170 cm−1), 2 100150200250300350400Wavenumber (cm−1)10152025Radiance (10−8W/cm2/sr/cm−1)C4H2C2N2C3H4CH4"Haystack"FP1 spectrum at 89◦N, March 2007FitMeasured spectra the C4H2 band at 220 cm−1, the C2N2 band at 234 cm−1, and the C3H4 band at 327 cm−1 (see Fig. 1). The continuum emission comes from the collisions between the three main components of Titan's atmosphere (N2, CH4, and H2), and from the spectral contributions of the hazes. We retrieve the temperature profile using the con- strained non-linear inversion code NEMESIS (Irwin et al., 2008). We define a reference atmosphere, which takes into account the abundances of the main constituents of Titan's atmosphere measured by Cassini/CIRS (Coustenis et al., 2016; Nixon et al., 2012; Cottini et al., 2012; Teanby et al., 2009), Cassini/VIMS (Maltagliati et al., 2015), ALMA (Molter et al., 2016) and Huy- gens/GCMS(Niemann et al., 2010). We also consider the haze distribution and properties measured in previous studies with Cassini/CIRS (de Kok et al., 2007, 2010; Vinatier et al., 2012), and Huygens/GCMS (Tomasko et al., 2008). We consider four types of hazes, following de Kok et al. (2007): hazes 0 (70 cm−1 to 400 cm−1), A (centred at 140 cm−1), B (centred at 220 cm−1) and C (centred at 190 cm−1). For the spectra measured at high northern and southern latitudes during autumn and winter, we add an offset from 1 to 3 cm−1 to the nominal haze B cross-sections between 190 cm−1 and 240 cm−1, as in Sylvestre et al. (2018). This modification improves the fit of the continuum in the "haystack" which is a strong emission feature between 190 cm−1 and 240 cm−1 (see Fig. 1) seen at high latitudes during autumn and winter (e.g. in Coustenis et al. (1999); de Kok et al. (2007); Anderson et al. (2012); Jennings et al. (2012, 2015)). The variation of the offset allows us to take into account the evolution of the shape of this feature throughout autumn and winter. The composition of our reference atmosphere and the spectroscopic parameters adopted for its constituents are fully detailed in Sylvestre et al. (2018). We retrieve the temperature profile and scale factors applied to the a priori profiles of C2N2, C4H2, C3H4, and hazes 0, A, B and C, from the spectra using the constrained non-linear inversion code NEMESIS (Irwin et al., 2008). This code generates synthetic spectra from the reference atmosphere. At each iteration, the difference between the synthetic and the measured spectra is used to modify the profile of the retrieved variables, and minimise a cost function, in order to find the best fit for the measured spectrum. The sensitivity of the spectra to the temperature can be measured with the inversion kernels for the temperature (defined as Kij = ∂Ii , where Ii is the radiance measured ∂Tj at wavenumber wi, and Tj the temperature at pressure level pj) for several wavenumbers. The contribution of the methane lines to the temperature measurement can be isolated by defining their own inversion kernels K CH4 ij as follows: ij = Kij − K cont K CH4 ij (1) ij ij ij ij where K cont is the inversion kernel of the continuum for the same wavenumber. Figure 2 shows K CH4 for three of the rotational methane lines in the left panel, and the comparison between the sum of the 10 K CH4 (for the 10 rotational CH4 lines) and inversion kernels for the continuum (K cont at the wavenumbers of the CH4 lines and Kij outside of the CH4 lines) in the right panel. The CH4 lines allow us to measure lower stratospheric temperatures generally between 6 mbar and 25 mbar, with a maximal sensitivity at 15 mbar. The continuum emission mainly probes temperatures at higher pressures, around the tropopause and in the troposphere. The continuum emission mostly originates from the N2-N2 and N2-CH4 collisions induced absorption with some contribution from the hazes, for which we have limited constraints. However, Fig. 2 shows that the continuum emission comes from pressure levels located several scale heights below the region probed by the CH4 lines, so the lack of constraints on the hazes and tropospheric temperatures does not affect the lower stratospheric temperatures which are the main focus of this study. ij Figure 2: Sensitivity of temperature measurements at 72◦N in April 2007. Left panel: Normalised inversion kernels KCH4 in three of the CH4 rotational lines. Right panel: Comparison between the inversion kernels in the continuum (Kcont for three of the CH4 lines in dot-dashed lines, and Kij for other wavenumbers in the continuum in dashed lines) and the sum of the inversion kernels KCH4 of the CH4 rotational lines. CH4 rotational lines dominate the temperature retrievals in the lower stratosphere, generally from 6 to 25 mbar (and up to 35 mbar, depending on the datasets). The continuum emission probes temperatures at pressures higher than 50 mbar, mainly in the troposphere. ij ij 2.3. Error sources The main error sources in our temperature retrievals are the measurement noise and the uncertainties related 3 to the retrieval process such as forward modelling errors or the smoothing of the temperature profile. The total error on the temperature retrieval is estimated by NEMESIS and is in the order of 2 K from 6 mbar to 25 mbar. The other possible error source is the uncertainty on CH4 abundance, as Lellouch et al. (2014) showed that it can vary from 1% to 1.5% at 15 mbar. We performed additional temperature retrievals on several datasets, in order to assess the effects of these variations on the temperature retrievals. First, we selected datasets for which CH4 abundance was measured by Lellouch et al. (2014). In Figure 3, we show examples of these tests for two of these datasets: 52◦N in May 2007 and 15◦S in October 2006, for which Lellouch et al. (2014) measured respective CH4 abundances of qCH4 = 1.20 ± 0.15% and qCH4 = 0.95 ± 0.08% (the nominal value for our retrievals is qCH4 = 1.48 ± 0.09% from Niemann et al. (2010)). At 52◦N, the temperature profile obtained with the methane abundance from Lellouch et al. (2014) does not differ by more than 4 K from the nominal temperature profile. At 15 mbar (where the sensitivity to temperature is maximal in our retrievals), the difference of temperature between these two profiles is 2 K. Even a CH4 volume mixing ratio as low as 1% yields a temperature only 4 K warmer than the nominal temperature at 15 mbar. At 15◦S, the difference of temperature between the nominal retrieval and the retrieval with the methane abundance retrieved by Lellouch et al. (2014) (qCH4 = 0.95%), is approximately 9 K on the whole pressure range. We performed additional temperature retrievals using CIRS FP4 nadir spectra measured at the same times and latitudes as the two datasets shown in Figure 3. In FP4 nadir spectra, the methane band ν4 is visible between 1200 cm−1 and 1360 cm−1. This spectral feature allows us to probe temperature between 0.1 mbar and 10 mbar, whereas methane rotational lines in the CIRS FP1 nadir spectra generally probe temperature between 6 mbar and 25 mbar. Temperature can thus be measured with both types of retrievals from 6 mbar to 10 mbar. We performed FP4 temperature retrievals with the nominal methane abundance and the abundances measured by Lellouch et al. (2014), as shown in Figure 3. FP4 temperature retrievals seem less sensitive to changes in the methane volume mixing ratio, as they yield a maximal temperature difference of 3 K at 52◦N , and 4 K at 15◦S between 6 mbar and 10 mbar. In both cases, FP1 and FP4 temperature retrievals are in better agreement in their common pressure range when the nominal methane abundance (qCH4 = 1.48%) is used for both retrievals. This suggests that qCH4 = 1.48% is the best choice, at least in the pressure range covered by both types of temperature retrievals (from 6 mbar to 10 mbar). Changing the abundance of CH4 in the whole stratosphere seems to induce an error on the temperature measurements between 6 mbar and 10 mbar (up to 9 K at 15◦S), which probably affects the temperature at 15 mbar in the FP1 retrievals, because of the vertical resolution of nadir retrievals (represented by the width of the inversion kernels in Fig. 2). Consequently, assessing the effects of CH4 abundance variations on temperature at 15 mbar by changing qCH4 in the whole stratosphere seems to be a very unfavourable test, and the uncertainties on temperature determined by this method are probably overestimated for the FP1 temperature retrievals. Over- all, when retrieving temperature from CIRS FP1 nadir spectra with qCH4 = 1% for datasets spanning different times and latitudes, we found temperatures warmer than our nominal temperatures by 2 K to 10 K at 15 mbar, with an average of 5 K. In Lellouch et al. (2014), authors found that temperature changes by 4-5 K on the whole pressure range when varying qCH4 at 15◦S, but they determined temperatures using FP4 nadir and limb data, which do not probe the 15 mbar pressure level. 3. Results i.e. Figures 4 and 5 show the temperatures measured with Cassini/CIRS far-IR nadir data at 6 mbar (minimal pressure probed by the CIRS far-IR nadir observations) and 15 mbar (pressure level where these observations are the most sensitive). Figure 4 maps the seasonal evolution of temperatures throughout the Cassini mission (from 2004 to 2017, from mid-northern winter to early summer), while Figure 5 is focused on the evolution of the meridional gradient of temperature from one season to another. In both figures, both pressure levels exhibit significant seasonal variations of temperature and follow similar trends. Maximal temperatures are reached near the equator in 2005 (152 K at 6 mbar, 130 K at 15 mbar, at 18◦S, at LS = 300◦), while the minimal temperatures are reached at high southern latitudes in autumn (123 K at 6 mbar, 106 K at 15 mbar at 70◦S in 2016, at LS = 79◦). The maximal seasonal variations of temperature are located at the poles for both pressure levels. At high northern latitudes (60◦N - 90◦N), at 15 mbar, the temperature increased overall from winter to summer solstice. For instance at 70◦N, temperature increased by 10 K from January 2007 to September 2017. At 6 mbar, temperatures at 60◦N stayed approximately constant from winter to spring, whereas latitudes poleward from 70◦N warmed up. At 85◦N, the temperature increased continuously from 125 K in March 2007 to 142 K in September 2017. In the meantime, at high southern latitudes (60◦S - 90◦S), at 6 mbar and 15 mbar, temperatures strongly decreased from southern summer (2007) to late autumn (2016). It is the largest seasonal temperature change we measured in the lower stratosphere. At 70◦S, temperature decreased by 24 K at 6 mbar and by 19 K at 15 mbar 4 Figure 3: Temperature profiles from CIRS FP1 and FP4 nadir ob- servations at 52◦N in May 2007 (top panel) and 15◦S in October 2006 (bottom panel), retrieved with the methane abundances mea- sured by Niemann et al. (2010) (nominal value in this study) and Lellouch et al. (2014). In both cases, the nominal value from Nie- mann et al. (2010) yields a better agreement between the two types of observations. Figure 4: Evolution of temperatures at 6 mbar (120 km) and 15 mbar (85 km) from northern winter (2004) to summer (2017). The length of the markers shows the average size of the field of view of the CIRS FP1 detector. Temperatures exhibit similar strong seasonal changes at both pressure levels, especially at the poles. 5 distribution occur because of the temperature variations poleward from 60◦. At 6 mbar, temperature variations occur mostly in the southern hemisphere at latitudes higher than 40◦S, and near the North pole at latitudes higher than 70◦N. Figure 5: Meridional distribution of temperatures at 6 mbar (120 km) and 15 mbar (85 km), for three different seasons: late northern win- ter (2007, blue triangles), mid-spring (2013, green circles), and near summer solstice (from July 2016 to September 2017, red diamonds). The plain lines are the meridional distributions given by GCM sim- ulations at comparable seasons (see section 4). In both observations and model the meridional gradient of temperatures evolves from one season to another at both pressure levels. between January 2007 and June 2016. This decrease seems to be followed by a temperature increase toward winter solstice. At 70◦S, temperatures varied by +8 K at 6 mbar from June 2016 to April 2017. Temperatures at high southern latitudes began to evolve in November 2010 at 6 mbar, and 2 years later (in August 2012) at 15 mbar. Other latitudes experience moderate seasonal tem- perature variations. At low latitudes (between 30◦N and 30◦S), temperature decreased overall from 2004 to 2017 at both pressure levels. For instance, at the equator, at 6 mbar temperature decreased by 6 K from 2006 to 2016. At mid-southern latitudes, temperatures stayed constant from summer (2005) to mid-autumn (June 2012 at 6 mbar, and May 2013 at 15 mbar), then they decreased by approximately 10 K from 2012-2013 to 2016. At mid-northern latitudes temperatures increased overall from winter to spring. At 50◦N, temperature increased from 139 K to 144 K from 2005 to 2014. In Figure 5, at 6 mbar and 15 mbar, the meridional temperature gradient evolves from one season to another. During late northern winter, temperatures were approximately constant from 70◦S to 30◦N, and then decreased toward the North pole. In mid-spring, temperatures were decreasing from equator to poles. Near the summer solstice, at 15 mbar, the meridional temperature gradient reversed compared to winter (summer temperatures constant in northern and low southern latitudes then decreasing toward the South Pole), while at 6 mbar, temperatures globally decrease from the equator to the South pole and 70◦N, then increase slightly between 70◦N and 90◦N. At 15 mbar, most of these changes in the shape of the temperature Figure 6: Temperature variations in the lower stratosphere during the Cassini mission for different latitudes. The blue profiles were measured during northern winter (in 2007). The red profiles were measured in late northern spring (in 2017 for 85◦N, in 2016 for the other latitudes). The seasonal temperature variations are observed at most latitudes, and on the whole probed pressure range. Figure 6 shows the first and the last temperature profiles measured with CIRS nadir far-IR data, for several latitudes. As in Fig. 4, the maximal temperature variations are measured at high southern latitudes for all pressure levels. At 70◦S, the temperature decreased by 25 K at 10 mbar. Below 10 mbar the seasonal temperature difference decreases rapidly with increasing pressure until it reaches 10 K at 25 mbar, whereas it is nearly constant 85◦N also exhibits a between 5 mbar and 10 mbar. decrease of the seasonal temperature gradient below the 10 mbar pressure level, although it is less pronounced than near the South pole. At 45◦S, the temperature decreased by approximately 10 K from 2007 to 2016, over the whole probed pressure range. At the equator, the temperature varies by -5 K from 2005 to 2016 at 6 mbar and the amplitude of this variation seems to decrease slightly with increasing pressure until it becomes negligible at 25 mbar. However the amplitude of these variations is in the same range as the uncertainty on temperature due to potential CH4 variations. 4. Discussion 4.1. Comparison with previous results Figure 7 shows a comparison between our results and previous studies where temperatures have been measured 6 1201251301351401451501556 mbar200720132016-20179080706050403020100102030405060708090Latitude10511011512012513013514015 mbar200720132016-2017Temperature (K) 300 km). CIRS FP1 and FP4 temperature profiles are in good averall agreement. The profile we measured at 28◦S in February 2006 and the corresponding radio-occultation profile are within error bars for pressures lower than 13 mbar, then the difference between them increases up to 8 K at 25 mbar. The bottom left panel of Fig. 7 shows the radio-occultation temperatures in 2006 and 2007 compared to CIRS nadir FP1 temperatures at 15 mbar, where their sensitivity to the temperature is maximal. Although, the radio-occultations temperatures are systematically higher than the CIRS temperatures by 2 K to 6 K, they follow the same meridional trend. CIRS FP1 temperatures at the equator are also lower than the temperature measured by the HASI instrument at 15 mbar during Huygens descent in Titan's atmosphere in 2005. If we take into account the effect of the spatial variations of CH4 at 15 mbar observed by Lellouch et al. (2014) by decreasing the CH4 abundance to 1% (the lower limit in Lellouch et al. (2014)) in the CIRS FP1 temperature measurements (dashed line in the middle panel of Fig. 7), the agreement between the three types of observations is good in the southern hemisphere. The differences between radio-occultations, HASI and CIRS temperatures might also be explained by the difference of vertical resolution. Indeed nadir observations have a vertical resolution in the order of 50 km while radio-occultations and HASI observations have respective vertical resolutions of 1 km and 200 m around 15 mbar. 4.2. Effects of Saturn's eccentricity Figure 8: Temporal evolution of Titan's lower stratospheric temper- atures at the equator (5◦N - 5◦S) at 6 mbar (left panel) and 15 mbar (right panel), compared with a simple model of the evolution of the temperature as a function of the distance between Titan and the Sun (green line). The reduced χ2 between this model and the observa- tions is 0.95 at 6 mbar and 1.07 at 15 mbar. The amplitude of the temperature variations at Titan's equator throughout the Cassini mission can be explained by the effect of Saturn's eccentricity. Because of Saturn's orbital eccentricity of 0.0565, the 7 Figure 7: Comparison of nadir FP1 temperatures with previous stud- ies. Top left panel: Comparison between CIRS nadir FP1 (trian- gles) and CIRS nadir FP4 temperatures at 6 mbar (circles, Bampa- sidis et al. (2012)[1], and Coustenis et al. (2016)[2]) in 2010 (cyan) and 2014 (purple). Right panel: Comparison between tempera- ture profiles from CIRS nadir FP1 observations (thick solid lines), CIRS nadir FP4 observations (thin dot-dashed lines, Coustenis et al. (2016)[2]), and Cassini radio-occultation (thin dashed line, Schin- der et al. (2011)[3]). Our results are in good agreement with CIRS FP4 temperatures, but diverge somewhat from radio-occultation pro- files with increasing pressure. Bottom left panel: Comparison be- tween temperatures at 15 mbar from our CIRS FP1 nadir measure- ments (magenta triangles), Cassini radio-occultations in 2006 and 2007 (cyan circles, Schinder et al. (2011, 2012), [3], [4]), and the Huygens/HASI measurement in 2005 (yellow diamond, Fulchignoni et al. (2005), [5]).The dashed magenta line shows the potential ef- fect of the CH4 variations observed by Lellouch et al. (2014). If we take into account this effect, the agreement between our data, the radio-occultations and the HASI measurements is good. in the lower stratosphere at similar epochs, latitudes and pressure levels. In the top left and right panels, our temperature measurements are compared to results from CIRS FP4 nadir observations (Bampasidis et al., 2012; Coustenis et al., 2016) which probe mainly the 0.1-10 mbar pressure range. In the top left panel, the temperatures measured at 6 mbar by these two types of observations are in good agreement for the two considered epochs (2009- 2010 and 2014). We obtain similar meridional gradients with both types of observations, even if FP4 temperatures are obtained from averages of spectra over bins of 10◦ of latitudes (except at 70◦N and 70◦S where the bins are 20◦ wide in latitude), whereas the average size in latitude of the field of view of the FP1 detector is 20◦. It thus seems than the wider latitudinal size of the FP1 field of view has little effect on our temperature measurements. In the right panel, our temperature profiles are compared to two profiles measured by Coustenis et al. (2016) using CIRS FP4 nadir observations (at 50◦S in April 2010, and at 70◦S in June 2012), and with Cassini radio-occultations measurements from Schinder et al. (2011, 2012), which probe the atmosphere from the surface to 0.1 mbar (0 - 4681012141618Year+2e3142144146148150152154Temperature (K)6 mbar4681012141618Year+2e312012212412612813013215 mbar distance between Titan and the Sun varies enough to affect significantly the insolation. For instance, throughout the Cassini mission, the solar flux received at the equator has decreased by 19% because of the eccentricity. We make a simple model of the evolution of the temperature T at the equator as a function of the distance between Titan and the Sun. In this model we assume that the temperature T at the considered pressure level and at a given time depends only on the absorbed solar flux F and we neglect the radiative exchanges between atmospheric layers: σT 4 = F (2) where  is the emissivity of the atmosphere at this pressure level, and σ the Stefan-Boltzmann constant. T can thus be defined as a function of the distance d between Titan and the Sun: T 4 = αL(cid:12) 16σπd2 (3) where L(cid:12) is the solar power, and α the absorptivity of the atmosphere. If we choose a reference temperature T0 where Titan is at a distance d0 from the Sun, a relation similar to (3) can be written for T0. If we assume  and α to be constant, T can then be written as: (cid:114) T = T0 d0 d (4) Figure 8 shows a comparison between this model and the temperatures measured between 5◦N and 5◦S from 2006 to 2016, at 6 mbar and 15 mbar. We choose T0 as the temperature at the beginning of the observations (December 2005/January 2006) which provides the best fit between our model and the observations while being consistent with the observations at the same epoch (T0 = 151.7 K at 6 mbar, and T0 = 129 K at 15 mbar). At 6 mbar, we measure a temperature decrease from 2006 to 2016. This is similar to what has been measured at 4 mbar by B´ezard et al. (2018) with CIRS mid-IR observations, whereas their radiative-dynamical model predicts a small temperature maximum around the northern spring equinox (2009). At 15 mbar, equatorial temperatures are mostly constant from 2005 to 2016, with a marginal decrease in 2016. Our model predicts tempera- ture variations of 8 K at 6 mbar and 7 K at 15 mbar from 2006 to 2016. Both predictions are consistent with the measurements and with radiative timescales shorter than one Titan year at 6 mbar and 15 mbar, as in B´ezard et al. (2018) where they are respectively equal to 0.024 Titan year and 0.06 Titan year. At both pressure levels, the model captures the magnitude of the temperature change, but does not fully match its timing or shape (especially in 2012-2014), implying that a more sophisticated model is needed. The remaining differences between our model and the temperature measurements could be decreased by adding a temporal lag to our model (2-3 years at 6 mbar and 3-4 years at 15 mbar), but the error bars on 8 the temperature measurements are too large to constrain the lag to a value statistically distinct from zero. Even with this potential lag, the agreement between the model and the temperatures measured at 6 mbar shows that the amplitude of the temporal evolution throughout the Cassini mission may be explained by the effects of Saturn's eccentricity. At 15 mbar, given the error bars and the lack of further far-IR temperature measurements at the equator in 2016 and 2017, it remains difficult to draw a definitive conclusion about the influence of Saturn's eccentricity at this pressure level. 4.3. Implication for radiative and dynamical processes of the lower stratosphere In Section 3, we showed that in the lower stratosphere, the seasonal evolution of the temperature is maximal at high latitudes, especially at the South Pole. At 15 mbar, the strong cooling of high southern latitudes started in 2012, simultaneously with the increase in C2N2, C4H2, and C3H4 abundances measured at the same latitudes and pressure-level in Sylvestre et al. (2018). We also show that this cooling affects the atmosphere at least down to the 25 mbar pressure level (altitude of 70 km). The enrichment of the gases and cooling are consistent with the onset of a subsidence above the South Pole during autumn, as predicted by GCMs (Newman et al., 2011; Lebonnois et al., 2012), and inferred from previous CIRS observations at higher altitudes (Teanby et al., 2012; Vinatier et al., 2015; Coustenis et al., 2016). As Titan's atmospheric circulation transitions from two equator-to-poles cells (with upwelling above the equator and subsidence above the poles) to a single pole-to-pole cell (with a descending branch above South Pole), this subsidence drags downward photochemical species created at higher altitudes toward the lower stratosphere. Teanby et al. (2017) showed that enrichment in trace gases may be so strong that their cooling effect combined with the insolation decrease may exceed the adiabatic heating between 0.3 mbar and 10 mbar (100 - 250 km). Our observations show that this phenomena may be at play down as deep as 25 mbar. We compare retrieved temperature fields with results of simulations from IPSL 3D-GCM (Lebonnois et al., 2012) with an updated radiative transfer scheme (Vatant d'Ollone et al., 2017) now based on a flexible correlated-k method and up-to-date gas spectroscopic data (Rothman et al., 2013). It does not take into account the radiative feedback of the enrichment in hazes and trace gases in the polar regions, but it nevertheless appears that there is a good agreement in terms of seasonal cycle between the model and the observations. As shown in Figure 5, at 6 mbar meridional distributions and values of tempera- tures in the model match well the observations. It can be pointed out that in both model and observations there is a noticeable asymmetry between high southern latitudes where the temperature decreases rapidly from the equinox to winter, and high northern latitudes which evolve more slowly from winter to summer. For instance, in both CIRS data and model, between 2007 and 2013 at 6 mbar and 70◦N the atmosphere has warmed by only about 2 K, while in the meantime at 70◦S it has cooled by about 10-15 K. This is consistent with an increase of radiative timescales at high northern latitudes (due to lower temperatures, Achterberg et al. (2011)) which would remain cold for ap- proximately one season even after the return of sunlight. Figure 9 shows the temporal evolution of the temperature at 70◦N over one Titan year in the lower stratosphere in the GCM simulations and also emphasizes this asymmetry between the ingress and egress of winter at high latitudes. In Figure 5, at 15 mbar modeled temperatures underes- timate the observations by roughly 5-10 K, certainly due to a lack of infrared coolers such as clouds condensates (Jennings et al., 2015). However, observations and sim- ulations exhibit similar meridional temperature gradients for the three studied epochs, and similar seasonal temper- ature evolution. For instance, in 2016-2017 we measured a temperature gradient of -11 K between the North and South Pole, whereas GCM simulations predict a tempera- ture gradient of -12 K. At 70◦S, temperature decreases by 10 K between 2007 and 2016-2017 in the GCM and in our observations. Besides, at 15 mbar, the seasonal behaviour remains the same as at 6 mbar, although more damped. Indeed comparison with GCM results also supports the idea that the seasonal effects due to the variations of in- solation are damped with increasing depth in the lower stratosphere and ultimately muted below 25 mbar, as dis- played in Figure 9. At lower altitudes the seasonal cy- cle of temperature at high latitudes is even inverted with temperatures increasing in the winter and decreasing in summer. Indeed at these altitudes, due to the radiative timescales exceeding one Titan year, temperature is no more sensitive to the seasonal variations of solar forcing, but to the interplay of ascending and descending large scale vertical motions of the pole-to-pole cell, inducing re- spectively adiabatic heating above winter pole and cooling above summer pole, as previously discussed in Lebonnois et al. (2012). Further analysis of simulations -- not presented here - also show that after 2016, temperatures at high southern latitudes began to slightly increase again at 6 mbar, which is consistent with observations, whereas at 15 mbar no change in the trend is observed, certainly due to a phase shift of the seasonal cycle between the two altitudes induced by the difference of radiative timescales, which is also illustrated in Figure 9. We also show in Figure 6 that at high southern lati- tudes, from 6 to 10 mbar seasonal temperature variations are approximately constant with pressure and can be larger than 10 K, whereas they decrease with increasing pressure below 10 mbar. This transition at 10 mbar may be caused by the increase of radiative timescales in the Figure 9: Seasonal evolution of Titan's lower stratospheric tempera- tures modeled by the IPSL 3D-GCM at 70◦N - between 5 mbar and 50 mbar, starting at northern spring equinox. In the pressure range probed by the CIRS far-IR observations (from 6 mbar to 25 mbar), there is a strong asymmetry between the rapid temperature changes after autumn equinox (LS = 180◦) and the slow evolution of the thermal structure after spring equinox (LS = 0◦). lower stratosphere. Strobel et al. (2010) estimated that the radiative timescale increases from one Titan season at 6 mbar to half a Titan year at 12 mbar. It can thus be expected that this region should be a transition zone between regions of the atmosphere where the atmospheric response to the seasonal insolation variations is significant and comes with little lag, to regions of the atmosphere where they are negligible. However, this transition should be observable at other latitudes such as 45◦S, whereas Figure 6 shows a seasonal gradient constant with pressure at this latitude. Furthermore, in B´ezard et al. (2018), the authors show that the method used to estimate radiative timescales in Strobel et al. (2010) tends to overestimate them, and that in their model radiative timescales are less than a Titan season down to the 35 mbar pressure level, which is more consistent with the seasonal variations measured at 45◦S. The 10 mbar transition can also be caused by the interplay between photochemical, radiative and dynamical processes at high latitudes. Indeed, as photochemical species are transported downward by the subsidence above the autumn/winter pole, build up and cool strongly the lower atmosphere, the condensation level of species such as HCN, HC3N, C4H2 or C6H6 may be shifted upward, toward the 10 mbar level. Hence, below this pressure level, the volume mixing ratios of these gases 9 04590135180225270315360Solar longitude ()567891020304050Pressure (mbar)GCM Temperatures at 70N708090100110120130140150160Temperature (K) ST/MOO7715/1) and the Cassini project. JVO and SL acknowledge support from the Centre National d'Etudes Spatiales (CNES). GCM simulations have been performed thanks to computation facilities provided by the Grand ´Equipement National de Calcul Intensif (GENCI) on the Occigen/CINES cluster (allocation A0040110391). This research made use of Astropy, a community-developed core Python package for Astronomy (Astropy Collaboration et al., 2013), and matplotlib, a Python library for pub- lication quality graphics (Hunter, 2007) Appendix. Cassini/CIRS Datasets analysed in this study would rapidly decrease, along with their cooling effect. Many observations, especially during the Cassini mission showed that during winter and autumn, polar regions host clouds composed of ices of photochemical species. For instance, the "haystack" feature showed in Fig. 1 has been studied at both poles in Coustenis et al. (1999); Jennings et al. (2012, 2015), and is attributed to a mixture of condensates, possibly of nitrile origin. Moreover, HCN ice has been measured in the southern polar cloud observed by de Kok et al. (2014) with Cassini/VIMS observations. C6H6 ice has also been detected by Vinatier et al. (2018) in CIRS observations of the South Pole. The condensation curve for C4H2 in Barth (2017) is also consistent with the formation of C4H2 ice around 10 mbar with the temper- atures we measured at 70◦S in 2016. These organic ices may also have a cooling effect themselves as B´ezard et al. (2018) showed that at 9 mbar, the nitrile haze measured by Anderson and Samuelson (2011) contributes to the cooling with an intensity comparable to the contribution of gases such as C2H2 and C2H6. 5. Conclusion In this paper, we analysed all the available nadir far-IR CIRS observations to measure Titan's lower stratospheric temperatures (6 mbar - 25 mbar) throughout the 13 years of the Cassini mission, from northern winter to summer solstice. In this pressure range, significant temperature changes occur from one season to another. Temperatures evolve moderately at low and mid-latitudes (less than 10 K between 6 and 15 mbar). At the equator, at 6 mbar we measure a temperature decrease mostly due to Saturn's eccentricity. Seasonal temperature changes are maximal at high latitudes, especially in the southern hemisphere where they reach up to -19 K at 70◦S between summer (2007) and late autumn (2016) at 15 mbar. The strong seasonal evolution of high southern latitudes is due to a complex interplay between photochemistry, atmospheric dynamics with the downwelling above the autumn/winter poles, radiative processes with a large contribution of the gases transported toward the lower stratosphere, and possibly condensation due to the cold autumn polar temperatures and strong enrichments in trace gases. Recent GCM simulations show a good agreement with the observed seasonal variations in this pressure range, even though these simulations do not include coupling with variations of opacity sources. In particular at high latitudes, the fast decrease of temperatures when entering winter and slower increase when getting into summer is well reproduced in these simulations. Acknowledgements This research was funded by the UK Sciences and Technology Facilities Research council (grant number 10 Table 1: Far-IR CIRS datasets presented in this study. N stands for the number of spectra measured during the acquisition. FOV is the field of view. The asterisk denotes datasets where two different latitudes were observed. Date Observations CIRS 00BTI FIRNADCMP001 PRIME 12 Dec. 2004 15 Feb. 2005 CIRS 003TI FIRNADCMP002 PRIME CIRS 005TI FIRNADCMP002 PRIME 31 Mar. 2005 01 Apr. 2005 CIRS 005TI FIRNADCMP003 PRIME 16 Apr. 2005 CIRS 006TI FIRNADCMP002 PRIME 06 Jun. 2005 CIRS 009TI COMPMAP002 PRIME CIRS 013TI FIRNADCMP003 PRIME 21 Aug. 2005 22 Aug. 2005 CIRS 013TI FIRNADCMP004 PRIME 28 Oct. 2005 CIRS 017TI FIRNADCMP003 PRIME 26 Dec. 2005 CIRS 019TI FIRNADCMP002 PRIME CIRS 020TI FIRNADCMP002 PRIME 14 Jan. 2006 27 Feb. 2006 CIRS 021TI FIRNADCMP002 PRIME 18 Mar. 2006 CIRS 022TI FIRNADCMP003 PRIME 19 Mar. 2006 CIRS 022TI FIRNADCMP008 PRIME CIRS 023TI FIRNADCMP002 PRIME 01 May 2006 CIRS 024TI FIRNADCMP003 PRIME 19 May 2006 02 Jul. 2006 CIRS 025TI FIRNADCMP002 PRIME CIRS 025TI FIRNADCMP003 PRIME 01 Jul. 2006 07 Sep. 2006 CIRS 028TI FIRNADCMP003 PRIME CIRS 029TI FIRNADCMP003 PRIME 23 Sep. 2006 10 Oct. 2006 CIRS 030TI FIRNADCMP002 PRIME 09 Oct. 2006 CIRS 030TI FIRNADCMP003 PRIME 25 Oct. 2006 CIRS 031TI COMPMAP001 VIMS CIRS 036TI FIRNADCMP002 PRIME 28 Dec. 2006 27 Dec. 2006 CIRS 036TI FIRNADCMP003 PRIME 12 Jan. 2007 CIRS 037TI FIRNADCMP001 PRIME 13 Jan. 2007 CIRS 037TI FIRNADCMP002 PRIME CIRS 038TI FIRNADCMP001 PRIME 28 Jan. 2007 CIRS 038TI FIRNADCMP002 PRIME 29 Jan. 2007 22 Feb. 2007 CIRS 039TI FIRNADCMP002 PRIME 09 Mar. 2007 CIRS 040TI FIRNADCMP001 PRIME CIRS 040TI FIRNADCMP002 PRIME 10 Mar. 2007 26 Mar. 2007 CIRS 041TI FIRNADCMP002 PRIME 10 Apr. 2007 CIRS 042TI FIRNADCMP001 PRIME CIRS 042TI FIRNADCMP002 PRIME 11 Apr. 2007 CIRS 043TI FIRNADCMP001 PRIME 26 Apr. 2007 27 Apr. 2007 CIRS 043TI FIRNADCMP002 PRIME 13 May 2007 CIRS 044TI FIRNADCMP002 PRIME 28 May 2007 CIRS 045TI FIRNADCMP001 PRIME CIRS 045TI FIRNADCMP002 PRIME 29 May 2007 13 Jun. 2007 CIRS 046TI FIRNADCMP001 PRIME 14 Jun. 2007 CIRS 046TI FIRNADCMP002 PRIME 29 Jun. 2007 CIRS 047TI FIRNADCMP001 PRIME 30 Jun. 2007 CIRS 047TI FIRNADCMP002 PRIME CIRS 048TI FIRNADCMP001 PRIME 18 Jul. 2007 19 Jul. 2007 CIRS 048TI FIRNADCMP002 PRIME 01 Oct. 2007 CIRS 050TI FIRNADCMP001 PRIME 02 Oct. 2007 CIRS 050TI FIRNADCMP002 PRIME CIRS 052TI FIRNADCMP002 PRIME 19 Nov. 2007 CIRS 053TI FIRNADCMP001 PRIME 04 Dec. 2007 CIRS 053TI FIRNADCMP002 PRIME 05 Dec. 2007 11 N 224 180 241 240 178 184 192 248 119 124 107 213 401 83 215 350 307 190 350 312 340 286 160 136 321 161 107 254 254 23 159 109 102 103 272 263 104 104 231 346 60 102 204 238 96 260 144 106 272 223 102 Latitude (◦N) FOV (◦) 16.4 -18.7 -41.1 47.8 54.7 -89.7 30.1 -53.7 20.1 -0.0 19.5 -30.2 -0.4 25.3 -35.0 -15.5 25.1 39.7 29.7 9.5 -59.1 33.9 -14.5 -89.1 78.6 75.2 -70.3 86.3 -39.7 69.9 -49.2 88.8 61.2 -60.8 71.5 -51.4 77.1 -0.5 -22.3 52.4 17.6 -20.8 9.8 20.1 -34.8 49.5 -10.1 29.9 40.3 -40.2 59.4 20.3 18.5 25.7 28.5 29.9 21.1 15.5 25.0 19.8 17.6 19.7 22.5 18.4 24.1 27.8 21.6 21.7 25.6 19.7 19.4 23.4 19.9 16.3 12.6 21.0 19.1 20.6 16.7 22.0 21.2 21.1 13.3 19.3 26.0 22.6 24.7 20.0 18.8 22.6 29.5 28.6 19.0 23.2 23.7 31.4 35.8 23.8 19.7 26.5 25.8 28.3 CIRS 054TI FIRNADCMP002 PRIME CIRS 055TI FIRNADCMP001 PRIME CIRS 055TI FIRNADCMP002 PRIME CIRS 059TI FIRNADCMP001 PRIME CIRS 059TI FIRNADCMP002 PRIME CIRS 062TI FIRNADCMP002 PRIME CIRS 067TI FIRNADCMP002 PRIME CIRS 069TI FIRNADCMP001 PRIME CIRS 069TI FIRNADCMP002 PRIME CIRS 093TI FIRNADCMP002 PRIME CIRS 095TI FIRNADCMP001 PRIME CIRS 097TI FIRNADCMP001 PRIME CIRS 106TI FIRNADCMP001 PRIME CIRS 107TI FIRNADCMP002 PRIME CIRS 110TI FIRNADCMP001 PRIME CIRS 111TI FIRNADCMP002 PRIME CIRS 112TI FIRNADCMP001 PRIME CIRS 112TI FIRNADCMP002 PRIME CIRS 114TI FIRNADCMP001 PRIME CIRS 115TI FIRNADCMP001 PRIME CIRS 119TI FIRNADCMP002 PRIME CIRS 122TI FIRNADCMP001 PRIME CIRS 123TI FIRNADCMP002 PRIME CIRS 124TI FIRNADCMP002 PRIME CIRS 125TI FIRNADCMP001 PRIME CIRS 125TI FIRNADCMP002 PRIME CIRS 129TI FIRNADCMP001 PRIME CIRS 131TI FIRNADCMP001 PRIME CIRS 131TI FIRNADCMP002 PRIME CIRS 132TI FIRNADCMP002 PRIME CIRS 133TI FIRNADCMP001 PRIME CIRS 134TI FIRNADCMP001 PRIME CIRS 138TI FIRNADCMP001 PRIME CIRS 139TI COMPMAP001 PRIME* CIRS 139TI COMPMAP001 PRIME* CIRS 148TI FIRNADCMP001 PRIME CIRS 153TI FIRNADCMP001 PRIME CIRS 158TI FIRNADCMP501 PRIME CIRS 159TI FIRNADCMP001 PRIME CIRS 160TI FIRNADCMP001 PRIME CIRS 160TI FIRNADCMP002 PRIME CIRS 161TI FIRNADCMP001 PRIME CIRS 161TI FIRNADCMP002 PRIME CIRS 166TI FIRNADCMP001 PRIME CIRS 167TI FIRNADCMP002 PRIME CIRS 169TI FIRNADCMP001 PRIME CIRS 172TI FIRNADCMP001 PRIME CIRS 172TI FIRNADCMP002 PRIME CIRS 174TI FIRNADCMP002 PRIME CIRS 175TI FIRNADCMP002 PRIME CIRS 185TI FIRNADCMP001 PRIME CIRS 185TI FIRNADCMP002 PRIME CIRS 190TI FIRNADCMP001 PRIME CIRS 190TI FIRNADCMP002 PRIME CIRS 194TI FIRNADCMP001 PRIME CIRS 195TI FIRNADCMP001 PRIME CIRS 197TI FIRNADCMP001 PRIME 21 Dec. 2007 05 Jan. 2008 06 Jan. 2008 22 Feb. 2008 23 Feb. 2008 25 Mar. 2008 12 May 2008 27 May 2008 28 May 2008 20 Nov. 2008 05 Dec. 2008 20 Dec. 2008 26 Mar. 2009 27 Mar. 2009 06 May 2009 22 May 2009 06 Jun. 2009 07 Jun. 2009 09 Jul. 2009 24 Jul. 2009 12 Oct. 2009 11 Dec. 2009 28 Dec. 2009 13 Jan. 2010 28 Jan. 2010 29 Jan. 2010 05 Apr. 2010 19 May 2010 20 May 2010 05 Jun. 2010 20 Jun. 2010 06 Jul. 2010 24 Sep. 2010 14 Oct. 2010 14 Oct. 2010 08 May 2011 11 Sep. 2011 13 Dec. 2011 02 Jan. 2012 29 Jan. 2012 30 Jan. 2012 18 Feb. 2012 19 Feb. 2012 22 May 2012 07 Jun. 2012 24 Jul. 2012 26 Sep. 2012 26 Sep. 2012 13 Nov. 2012 29 Nov. 2012 05 Apr. 2013 06 Apr. 2013 23 May 2013 24 May 2013 10 Jul. 2013 25 Jul. 2013 11 Sep. 2013 12 107 190 284 172 98 115 286 112 112 161 213 231 165 164 282 168 218 274 164 146 166 212 186 272 156 280 119 188 229 167 187 251 190 132 108 200 227 369 275 322 280 121 89 318 293 258 282 270 298 299 244 303 224 298 186 186 330 60.4 18.7 44.6 -24.9 17.1 59.3 29.5 -44.6 9.5 43.7 -14.0 -10.9 -60.3 33.5 -68.1 -27.1 48.7 -58.9 -71.4 50.7 0.4 39.8 -46.1 -1.2 39.9 -44.9 -45.1 -30.0 -19.8 49.4 -49.7 -10.0 -30.1 -70.9 -53.8 -10.0 9.9 -29.9 -42.2 -40.0 -0.2 9.9 -15.0 -19.9 -45.4 -9.7 44.9 -70.4 -71.8 -59.9 15.0 -88.9 -0.2 -45.0 30.0 19.6 60.5 21.1 30.5 22.2 20.7 20.0 17.1 21.0 27.3 19.3 21.1 20.7 23.7 19.2 30.4 25.7 23.1 21.0 20.2 25.4 20.1 18.3 24.7 22.3 19.0 27.5 27.3 28.2 22.1 21.5 27.4 36.1 20.0 21.2 20.6 16.7 18.3 19.0 24.7 23.7 21.7 18.3 18.4 17.3 19.9 21.7 20.7 18.5 23.2 21.8 19.3 20.1 16.8 25.6 20.0 19.7 24.5 19.4 CIRS 198TI FIRNADCMP001 PRIME CIRS 198TI FIRNADCMP002 PRIME CIRS 199TI FIRNADCMP001 PRIME CIRS 200TI FIRNADCMP001 PRIME CIRS 200TI FIRNADCMP002 PRIME CIRS 201TI FIRNADCMP001 PRIME CIRS 201TI FIRNADCMP002 PRIME CIRS 203TI FIRNADCMP001 PRIME CIRS 203TI FIRNADCMP002 PRIME CIRS 204TI FIRNADCMP002 PRIME CIRS 205TI FIRNADCMP001 PRIME CIRS 205TI FIRNADCMP002 PRIME CIRS 206TI FIRNADCMP001 PRIME CIRS 206TI FIRNADCMP002 PRIME CIRS 207TI FIRNADCMP001 PRIME CIRS 207TI FIRNADCMP002 PRIME CIRS 208TI FIRNADCMP001 PRIME CIRS 208TI FIRNADCMP002 PRIME CIRS 209TI FIRNADCMP001 PRIME CIRS 209TI FIRNADCMP002 PRIME CIRS 210TI FIRNADCMP001 PRIME CIRS 210TI FIRNADCMP002 PRIME CIRS 211TI FIRNADCMP001 PRIME CIRS 211TI FIRNADCMP002 PRIME CIRS 212TI FIRNADCMP002 PRIME CIRS 213TI FIRNADCMP001 PRIME CIRS 213TI FIRNADCMP002 PRIME CIRS 215TI FIRNADCMP001 PRIME CIRS 215TI FIRNADCMP002 PRIME CIRS 218TI FIRNADCMP001 PRIME CIRS 218TI FIRNADCMP002 PRIME CIRS 222TI FIRNADCMP001 PRIME CIRS 222TI FIRNADCMP002 PRIME CIRS 230TI FIRNADCMP001 PRIME CIRS 231TI FIRNADCMP001 PRIME CIRS 231TI FIRNADCMP002 PRIME CIRS 232TI FIRNADCMP001 PRIME CIRS 232TI FIRNADCMP002 PRIME CIRS 234TI FIRNADCMP001 PRIME CIRS 235TI FIRNADCMP001 PRIME CIRS 235TI FIRNADCMP002 PRIME CIRS 236TI FIRNADCMP001 PRIME CIRS 236TI FIRNADCMP002 PRIME CIRS 238TI FIRNADCMP002 PRIME CIRS 248TI FIRNADCMP001 PRIME CIRS 248TI FIRNADCMP002 PRIME CIRS 250TI FIRNADCMP002 PRIME CIRS 259TI COMPMAP001 PIE CIRS 270TI FIRNADCMP001 PRIME CIRS 283TI COMPMAP001 PRIME* CIRS 283TI COMPMAP001 PRIME* CIRS 287TI COMPMAP001 PIE CIRS 288TI COMPMAP002 PIE CIRS 292TI COMPMAP001 PRIME 187 306 329 187 210 329 234 187 239 199 144 161 181 161 179 163 329 175 181 233 329 237 225 258 257 187 258 250 232 249 232 125 233 282 254 236 249 92 328 163 221 88 238 220 185 186 219 302 166 114 134 305 269 192 13 Oct. 2013 14 Oct. 2013 30 Nov. 2013 01 Jan. 2014 02 Jan. 2014 02 Feb. 2014 03 Feb. 2014 07 Apr. 2014 07 Apr. 2014 18 May 2014 18 Jun. 2014 18 Jun. 2014 19 Jul. 2014 20 Jul. 2014 20 Aug. 2014 21 Aug. 2014 21 Sep. 2014 22 Sep. 2014 23 Oct. 2014 24 Oct. 2014 10 Dec. 2014 11 Dec. 2014 11 Jan. 2015 12 Jan. 2015 13 Feb. 2015 16 Mar. 2015 16 Mar. 2015 07 May 2015 08 May 2015 06 Jul. 2015 07 Jul. 2015 28 Sep. 2015 29 Sep. 2015 15 Jan. 2016 31 Jan. 2016 01 Feb. 2016 16 Feb. 2016 17 Feb. 2016 04 Apr. 2016 06 May 2016 07 May 2016 07 Jun. 2016 07 Jun. 2016 25 Jul. 2016 13 Nov. 2016 14 Nov. 2016 30 Nov. 2016 01 Feb. 2017 21 Apr. 2017 10 Jul. 2017 10 Jul. 2017 11 Aug. 2017 11 Aug. 2017 12 Sep. 2017 13 88.9 -69.8 68.4 49.9 -59.8 19.9 -39.6 75.0 0.5 0.4 -45.1 30.3 -50.3 30.6 -70.0 79.7 -80.0 60.5 -35.2 50.5 -70.3 -19.6 19.6 40.0 -40.0 -31.6 23.4 -50.0 -30.0 -20.0 -40.0 30.0 -0.1 -15.0 15.0 0.4 -50.2 -19.8 19.8 -60.0 15.7 -70.5 60.8 15.4 -88.9 30.3 -19.8 -69.0 -74.7 60.0 67.5 88.9 66.7 70.4 8.7 24.0 23.9 19.6 21.3 26.8 20.9 18.0 27.5 27.0 20.5 19.1 17.8 18.4 17.8 17.6 15.6 17.8 17.7 18.5 25.2 27.6 25.0 19.3 30.1 19.6 20.5 31.0 21.7 19.9 25.2 21.7 18.6 19.5 19.6 18.9 24.5 21.5 24.7 19.7 20.1 20.5 20.0 20.5 18.3 17.4 28.4 20.6 25.4 26.5 24.7 9.3 23.7 19.2 References References Achterberg, R. K., Conrath, B. J., Gierasch, P. J., Flasar, F. M., Nixon, C. A., Mar. 2008. Titan's middle-atmospheric tempera- tures and dynamics observed by the Cassini Composite Infrared Spectrometer. Icarus194, 263 -- 277. Achterberg, R. K., Gierasch, P. J., Conrath, B. J., Michael Flasar, F., Nixon, C. A., Jan. 2011. Temporal variations of Titan's middle-atmospheric temperatures from 2004 to 2009 observed by Cassini/CIRS. Icarus211, 686 -- 698. Anderson, C., Samuelson, R., Achterberg, R., Apr. 2012. Ti- tan's stratospheric condensibles at high northern latitudes during northern winter. In: Cottini, V., Nixon, C., Lorenz, R. (Eds.), Ti- tan Through Time; Unlocking Titan's Past, Present and Future. p. 59. Anderson, C. M., Samuelson, R. E., Apr. 2011. Titan's aerosol and stratospheric ice opacities between 18 and 500 µm: Vertical and spectral characteristics from Cassini CIRS. Icarus212, 762 -- 778. Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., Greenfield, P., Droettboom, M., Bray, E., Aldcroft, T., Davis, M., Gins- burg, A., Price-Whelan, A. M., Kerzendorf, W. E., Conley, A., Crighton, N., Barbary, K., Muna, D., Ferguson, H., Grollier, F., Parikh, M. M., Nair, P. H., Unther, H. M., Deil, C., Woillez, J., Conseil, S., Kramer, R., Turner, J. E. H., Singer, L., Fox, R., Weaver, B. A., Zabalza, V., Edwards, Z. I., Azalee Bostroem, K., Burke, D. J., Casey, A. R., Crawford, S. M., Dencheva, N., Ely, J., Jenness, T., Labrie, K., Lim, P. L., Pierfederici, F., Pontzen, A., Ptak, A., Refsdal, B., Servillat, M., Streicher, O., Oct. 2013. Astropy: A community Python package for astronomy. A&A558, A33. Bampasidis, G., Coustenis, A., Achterberg, R. K., Vinatier, S., Lav- vas, P., Nixon, C. A., Jennings, D. E., Teanby, N. A., Flasar, F. M., Carlson, R. C., Moussas, X., Preka-Papadema, P., Romani, P. N., Guandique, E. A., Stamogiorgos, S., Dec. 2012. Thermal and Chemical Structure Variations in Titan's Stratosphere during the Cassini Mission. ApJ760, 144. Barth, E. L., Mar. 2017. Modeling survey of ices in Titan's strato- sphere. Planet. Space Sci.137, 20 -- 31. B´ezard, B., Vinatier, S., Achterberg, R. K., Mar. 2018. Seasonal radiative modeling of Titan's stratospheric temperatures at low latitudes. Icarus302, 437 -- 450. Cottini, V., Nixon, C. A., Jennings, D. E., Anderson, C. M., Gorius, N., Bjoraker, G. L., Coustenis, A., Teanby, N. A., Achterberg, R. K., B´ezard, B., de Kok, R., Lellouch, E., Irwin, P. G. J., Flasar, F. M., Bampasidis, G., Aug. 2012. Water vapor in Titan's stratosphere from Cassini CIRS far-infrared spectra. Icarus220, 855 -- 862. Coustenis, A., Achterberg, R. K., Conrath, B. J., Jennings, D. E., Marten, A., Gautier, D., Nixon, C. A., Flasar, F. M., Teanby, N. A., B´ezard, B., Samuelson, R. E., Carlson, R. C., Lellouch, E., Bjoraker, G. L., Romani, P. N., Taylor, F. W., Irwin, P. G. J., Fouchet, T., Hubert, A., Orton, G. S., Kunde, V. G., Vinatier, S., Mondellini, J., Abbas, M. M., Courtin, R., Jul. 2007. The com- position of Titan's stratosphere from Cassini/CIRS mid-infrared spectra. Icarus189, 35 -- 62. Coustenis, A., Jennings, D. E., Achterberg, R. K., Bampasidis, G., Lavvas, P., Nixon, C. A., Teanby, N. A., Anderson, C. M., Cot- tini, V., Flasar, F. M., May 2016. Titan's temporal evolution in stratospheric trace gases near the poles. Icarus270, 409 -- 420. Coustenis, A., Schmitt, B., Khanna, R. K., Trotta, F., Oct. 1999. Plausible condensates in Titan's stratosphere from Voyager in- frared spectra. Planet. Space Sci.47, 1305 -- 1329. de Kok, R., Irwin, P. G. J., Teanby, N. A., Nixon, C. A., Jennings, D. E., Fletcher, L., Howett, C., Calcutt, S. B., Bowles, N. E., Flasar, F. M., Taylor, F. W., Nov. 2007. Characteristics of Titan's stratospheric aerosols and condensate clouds from Cassini CIRS far-infrared spectra. Icarus191, 223 -- 235. de Kok, R., Irwin, P. G. J., Teanby, N. A., Vinatier, S., Tosi, F., Negrao, A., Osprey, S., Adriani, A., Moriconi, M. L., Coradini, A., May 2010. A tropical haze band in Titan's stratosphere. Icarus207, 485 -- 490. de Kok, R. J., Teanby, N. A., Maltagliati, L., Irwin, P. G. J., Vinatier, S., Oct. 2014. HCN ice in Titan's high-altitude southern polar cloud. Nature514, 65 -- 67. Flasar, F. M., Kunde, V. G., Abbas, M. M., Achterberg, R. K., Ade, P., Barucci, A., B´ezard, B., Bjoraker, G. L., Brasunas, J. C., Cal- cutt, S., Carlson, R., C´esarsky, C. J., Conrath, B. J., Coradini, A., Courtin, R., Coustenis, A., Edberg, S., Edgington, S., Ferrari, C., Fouchet, T., Gautier, D., Gierasch, P. J., Grossman, K., Irwin, P., Jennings, D. E., Lellouch, E., Mamoutkine, A. A., Marten, A., Meyer, J. P., Nixon, C. A., Orton, G. S., Owen, T. C., Pearl, J. C., Prang´e, R., Raulin, F., Read, P. L., Romani, P. N., Samuel- son, R. E., Segura, M. E., Showalter, M. R., Simon-Miller, A. A., Smith, M. D., Spencer, J. R., Spilker, L. J., Taylor, F. W., Dec. 2004. Exploring The Saturn System In The Thermal Infrared: The Composite Infrared Spectrometer. Space Sci. Rev.115, 169 -- 297. Fulchignoni, M., Ferri, F., Angrilli, F., Ball, A. J., Bar-Nun, A., Barucci, M. A., Bettanini, C., Bianchini, G., Borucki, W., Colom- batti, G., Coradini, M., Coustenis, A., Debei, S., Falkner, P., Fanti, G., Flamini, E., Gaborit, V., Grard, R., Hamelin, M., Harri, A. M., Hathi, B., Jernej, I., Leese, M. R., Lehto, A., Lion Stoppato, P. F., L´opez-Moreno, J. J., Makinen, T., McDonnell, J. A. M., McKay, C. P., Molina-Cuberos, G., Neubauer, F. M., Pirronello, V., Rodrigo, R., Saggin, B., Schwingenschuh, K., Seiff, A., Simoes, F., Svedhem, H., Tokano, T., Towner, M. C., Traut- ner, R., Withers, P., Zarnecki, J. C., Dec. 2005. In situ mea- surements of the physical characteristics of Titan's environment. Nature438, 785 -- 791. Hunter, J. D., 2007. Matplotlib: A 2d graphics environment. Com- puting In Science & Engineering 9 (3), 90 -- 95. Irwin, P. G. J., Teanby, N. A., de Kok, R., Fletcher, L. N., Howett, C. J. A., Tsang, C. C. C., Wilson, C. F., Calcutt, S. B., Nixon, C. A., Parrish, P. D., Apr. 2008. The NEMESIS planetary atmo- sphere radiative transfer and retrieval tool. J. Quant. Spec. Ra- diat. Transf.109, 1136 -- 1150. Jennings, D. E., Achterberg, R. K., Cottini, V., Anderson, C. M., Flasar, F. M., Nixon, C. A., Bjoraker, G. L., Kunde, V. G., Carl- son, R. C., Guandique, E., Kaelberer, M. S., Tingley, J. S., Al- bright, S. A., Segura, M. E., de Kok, R., Coustenis, A., Vinatier, S., Bampasidis, G., Teanby, N. A., Calcutt, S., May 2015. Evo- lution of the Far-infrared Cloud at Titan's South Pole. ApJ804, L34. Jennings, D. E., Anderson, C. M., Samuelson, R. E., Flasar, F. M., Nixon, C. A., Kunde, V. G., Achterberg, R. K., Cottini, V., de Kok, R., Coustenis, A., Vinatier, S., Calcutt, S. B., Jul. 2012. Sea- sonal Disappearance of Far-infrared Haze in Titan's Stratosphere. ApJ754, L3. Lebonnois, S., Burgalat, J., Rannou, P., Charnay, B., Mar. 2012. Titan global climate model: A new 3-dimensional version of the IPSL Titan GCM. Icarus218, 707 -- 722. Lellouch, E., B´ezard, B., Flasar, F. M., Vinatier, S., Achterberg, R., Nixon, C. A., Bjoraker, G. L., Gorius, N., Mar. 2014. The dis- tribution of methane in Titan's stratosphere from Cassini/CIRS observations. Icarus231, 323 -- 337. Lora, J. M., Lunine, J. I., Russell, J. L., Apr. 2015. GCM simula- tions of Titan's middle and lower atmosphere and comparison to observations. Icarus250, 516 -- 528. Maltagliati, L., B´ezard, B., Vinatier, S., Hedman, M. M., Lel- louch, E., Nicholson, P. D., Sotin, C., de Kok, R. J., Sicardy, B., Mar. 2015. Titan's atmosphere as observed by Cassini/VIMS solar occultations: CH4, CO and evidence for C2H6 absorption. Icarus248, 1 -- 24. Molter, E. M., Nixon, C. A., Cordiner, M. A., Serigano, J., Irwin, P. G. J., Teanby, N. A., Charnley, S. B., Lindberg, J. E., Aug. 2016. ALMA Observations of HCN and Its Isotopologues on Titan. AJ152, 42. Newman, C. E., Lee, C., Lian, Y., Richardson, M. I., Toigo, A. D., Jun. 2011. Stratospheric superrotation in the TitanWRF model. Icarus213, 636 -- 654. Niemann, H. B., Atreya, S. K., Demick, J. E., Gautier, D., Haber- 14 in Titan's middle atmosphere during the northern spring derived from Cassini/CIRS observations. Icarus250, 95 -- 115. Vinatier, S., Rannou, P., Anderson, C. M., B´ezard, B., de Kok, R., Samuelson, R. E., May 2012. Optical constants of Titan's strato- spheric aerosols in the 70-1500 cm−1 spectral range constrained by Cassini/CIRS observations. Icarus219, 5 -- 12. Vinatier, S., Schmitt, B., B´ezard, B., Rannou, P., Dauphin, C., de Kok, R., Jennings, D. E., Flasar, F. M., Aug. 2018. Study of Titan's fall southern stratospheric polar cloud composition with Cassini/CIRS: Detection of benzene ice. Icarus310, 89 -- 104. man, J. A., Harpold, D. N., Kasprzak, W. T., Lunine, J. I., Owen, T. C., Raulin, F., Dec. 2010. Composition of Titan's lower atmo- sphere and simple surface volatiles as measured by the Cassini- Huygens probe gas chromatograph mass spectrometer experiment. Journal of Geophysical Research (Planets) 115, E12006. Nixon, C. A., Temelso, B., Vinatier, S., Teanby, N. A., B´ezard, B., Achterberg, R. K., Mandt, K. E., Sherrill, C. D., Irwin, P. G. J., Jennings, D. E., Romani, P. N., Coustenis, A., Flasar, F. M., Apr. 2012. Isotopic Ratios in Titan's Methane: Measurements and Modeling. ApJ749, 159. Rothman, L. S., Gordon, I. E., Babikov, Y., Barbe, A., Chris Benner, D., Bernath, P. F., Birk, M., Bizzocchi, L., Boudon, V., Brown, L. R., Campargue, A., Chance, K., Cohen, E. A., Coudert, L. H., Devi, V. M., Drouin, B. J., Fayt, A., Flaud, J.-M., Gamache, R. R., Harrison, J. J., Hartmann, J.-M., Hill, C., Hodges, J. T., Jacquemart, D., Jolly, A., Lamouroux, J., Le Roy, R. J., Li, G., Long, D. A., Lyulin, O. M., Mackie, C. J., Massie, S. T., Mikhailenko, S., Muller, H. S. P., Naumenko, O. V., Nikitin, A. V., Orphal, J., Perevalov, V., Perrin, A., Polovtseva, E. R., Richard, C., Smith, M. A. H., Starikova, E., Sung, K., Tashkun, S., Tennyson, J., Toon, G. C., Tyuterev, V. G., Wagner, G., Nov. 2013. The HITRAN2012 molecular spectroscopic database. J. Quant. Spec. Radiat. Transf.130, 4 -- 50. Schinder, P. J., Flasar, F. M., Marouf, E. A., French, R. G., McGhee, C. A., Kliore, A. J., Rappaport, N. J., Barbinis, E., Fleischman, D., Anabtawi, A., Oct. 2011. The structure of Titan's atmosphere from Cassini radio occultations. Icarus215, 460 -- 474. Schinder, P. J., Flasar, F. M., Marouf, E. A., French, R. G., McGhee, C. A., Kliore, A. J., Rappaport, N. J., Barbinis, E., Fleischman, D., Anabtawi, A., Nov. 2012. The structure of Titan's atmosphere from Cassini radio occultations: Occultations from the Prime and Equinox missions. Icarus221, 1020 -- 1031. Strobel, D. F., Atreya, S. K., B´ezard, B., Ferri, F., Flasar, F. M., Fulchignoni, M., Lellouch, E., Muller-Wodarg, I., 2010. Atmo- spheric Structure and Composition. p. 235. Sylvestre, M., Teanby, N. A., Vinatier, S., Lebonnois, S., Irwin, P. G. J., Jan. 2018. Seasonal evolution of C2N2, C3H4, and C4H2 abundances in Titan's lower stratosphere. A&A609, A64. Teanby, N. A., B´ezard, B., Vinatier, S., Sylvestre, M., Nixon, C. A., Irwin, P. G. J., de Kok, R. J., Calcutt, S. B., Flasar, F. M., Nov. 2017. The formation and evolution of Titan's winter polar vortex. Nature Communications 8, 1586. Teanby, N. A., Irwin, P. G. J., de Kok, R., Jolly, A., B´ezard, B., Nixon, C. A., Calcutt, S. B., Aug. 2009. Titan's stratospheric C 2N 2, C 3H 4, and C 4H 2 abundances from Cassini/CIRS far- infrared spectra. Icarus202, 620 -- 631. Teanby, N. A., Irwin, P. G. J., de Kok, R., Vinatier, S., B´ezard, B., Nixon, C. A., Flasar, F. M., Calcutt, S. B., Bowles, N. E., Fletcher, L., Howett, C., Taylor, F. W., Feb. 2007. Vertical profiles of HCN, HC ¡SUB¿3¡/SUB¿N, and C ¡SUB¿2¡/SUB¿H ¡SUB¿2¡/SUB¿ in Titan's atmosphere derived from Cassini/CIRS data. Icarus186, 364 -- 384. Teanby, N. A., Irwin, P. G. J., Nixon, C. A., de Kok, R., Vinatier, S., Coustenis, A., Sefton-Nash, E., Calcutt, S. B., Flasar, F. M., Nov. 2012. Active upper-atmosphere chemistry and dynamics from po- lar circulation reversal on Titan. Nature491, 732 -- 735. Tomasko, M. G., Doose, L., Engel, S., Dafoe, L. E., West, R., Lem- mon, M., Karkoschka, E., See, C., Apr. 2008. A model of Ti- tan's aerosols based on measurements made inside the atmosphere. Planet. Space Sci.56, 669 -- 707. Vatant d'Ollone, J., Lebonnois, S., Guerlet, S., Apr. 2017. Modelling of Titan's middle atmosphere with the IPSL climate model. In: EGU General Assembly Conference Abstracts. Vol. 19. p. 10169. Vinatier, S., B´ezard, B., Fouchet, T., Teanby, N. A., de Kok, R., Irwin, P. G. J., Conrath, B. J., Nixon, C. A., Romani, P. N., Flasar, F. M., Coustenis, A., May 2007. Vertical abundance pro- files of hydrocarbons in Titan's atmosphere at 15◦S and 80◦ N retrieved from Cassini/CIRS spectra. Icarus188, 120 -- 138. Vinatier, S., B´ezard, B., Lebonnois, S., Teanby, N. A., Achterberg, R. K., Gorius, N., Mamoutkine, A., Guandique, E., Jolly, A., Jennings, D. E., Flasar, F. M., Apr. 2015. Seasonal variations 15
1712.07655
2
1712
2018-04-26T18:04:25
Effects of Disk Warping on the Inclination Evolution of Star-Disk-Binary Systems
[ "astro-ph.EP" ]
Several recent studies have suggested that circumstellar disks in young stellar binaries may be driven into misalignement with their host stars due to secular gravitational interactions between the star, disk and the binary companion. The disk in such systems is twisted/warped due to the gravitational torques from the oblate central star and the external companion. We calculate the disk warp profile, taking into account of bending wave propagation and viscosity in the disk. We show that for typical protostellar disk parameters, the disk warp is small, thereby justifying the "flat-disk" approximation adopted in previous theoretical studies. However, the viscous dissipation associated with the small disk warp/twist tends to drive the disk toward alignment with the binary or the central star. We calculate the relevant timescales for the alignment. We find the alignment is effective for sufficiently cold disks with strong external torques, especially for systems with rapidly rotating stars, but is ineffective for the majority of star-disk-binary systems. Viscous warp driven alignment may be necessary to account for the observed spin-orbit alignment in multi-planet systems if these systems are accompanied by an inclined binary companion.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 8 October 2018 (MN LATEX style file v2.2) Effects of Disk Warping on the Inclination Evolution of Star-Disk-Binary Systems J. J. Zanazzi1(cid:63), and Dong Lai1 1Cornell Center for Astrophysics and Planetary Science, Department of Astronomy, Cornell University, Ithaca, NY 14853, USA 8 October 2018 ABSTRACT Several recent studies have suggested that circumstellar disks in young stellar binaries may be driven into misalignement with their host stars due to secular gravitational interactions between the star, disk and the binary companion. The disk in such systems is twisted/warped due to the gravitational torques from the oblate central star and the external companion. We calculate the disk warp profile, taking into account of bending wave propagation and viscosity in the disk. We show that for typical protostellar disk parameters, the disk warp is small, thereby justifying the "flat-disk" approximation adopted in previous theoretical studies. However, the viscous dissipation associated with the small disk warp/twist tends to drive the disk toward alignment with the binary or the central star. We calculate the relevant timescales for the alignment. We find the alignment is effective for sufficiently cold disks with strong external torques, especially for systems with rapidly rotating stars, but is ineffective for the majority of star-disk-binary systems. Viscous warp driven alignment may be necessary to account for the observed spin-orbit alignment in multi-planet systems if these systems are accompanied by an inclined binary companion. Key words: hydrodynamics - planets and satellites: formation - protoplanetary discs - stars: binaries: general 1 INTRODUCTION Circumstellar disks in young protostellar binary systems are likely to form with an inclined orientation relative to the binary orbital plane, as a result of the complex star/binary/disc formation processes (e.g. Bate, Bonnell & Bromm 2003; McKee & Ostriker 2007; Klessen 2011). In- deed, many misaligned circumstellar disks in protostellar bi- naries have been found in recent years (e.g. Stapelfeldt et al. 1998, 2003; Neuhauser et al. 2009; Jensen & Akeson 2014; Williams et al. 2014; Brinch et al. 2016; Fern´andez-L´opez, Zapata & Gabbasov 2017; Lee et al. 2017). Such misaligned disks experience differential gravitational torques from the binary companion, and are expected to be twisted/warped while undergoing damped precession around the binary (e.g. Lubow & Ogilvie 2000; Bate et al. 2000; Foucart & Lai 2014). On the other hand, a spinning protostar has a rotation-induced quadrupole, and thus exerts a torque on the disk (and also receives a back-reaction torque) when the stellar spin axis and the disk axis are misaligned. This torque tends to induce warping in the inner disk and drives mutual precession between the stellar spin and disk. In the presence of both torques on the disk, from the binary and from the (cid:63) Email: [email protected] c(cid:13) 0000 RAS central star, how does the disk warp and precess? What is the long-term evolution of the disk and stellar spin in such star-disk-binary systems? These are the questions we intend to address in this paper. Several recent studies have examined the secular dy- namics of the stellar spin and circumstellar disk in the pres- ence of an inclined binary companion (Batygin 2012; Baty- gin & Adams 2013; Lai 2014a; Spalding & Batygin 2014, 2015). These studies were motivated by the observations of spin-orbit misalignments in exoplanetary systems con- taining hot Jupiters, i.e., the planet's orbital plane is often misaligned with the stellar rotational equator (see Winn & Fabrycky 2015 and Triaud 2017 for recent reviews). It was shown that significant "primordial" misalignments may be generated while the planetary systems are still forming in their natal protoplanetary disks through secular star-disk- binary gravitational interactions (Batygin & Adams 2013; Lai 2014a; Spalding & Batygin 2014, 2015). In these studies, various assumptions were made about the star-disk inter- actions, and uncertain physical processes such as star/disk winds, magnetic star-disk interactions, and accretion of disk angular momentum onto the star were incorporated in a pa- rameterized manner. Nevertheless, the production of spin- orbit misalignments seems quite robust. In Zanazzi & Lai (2018b), we showed that the formation 2 J. J. Zanazzi and Dong Lai of hot Jupiters in the protoplanetary disks can significantly suppress the excitation of spin-orbit misalignment in star- disk-binary systems. This is because the presence of such close-in giant planets lead to strong spin-orbit coupling be- tween the planet and its host star, so that the spin-orbit misalignment angle is adiabatically maintained despite the gravitational perturbation from the binary companion. How- ever, the formation of small planets or distant planets (e.g. warm Jupiters) do not affect the generation of primordial misalignments between the host star and the disk. A key assumption made in all previous studies on mis- alignments in star-disk-binary systems (Batygin & Adams 2013; Lai 2014a; Spalding & Batygin 2015) is that the disk is nearly flat and behaves like a rigid plate in response to the external torques from the binary and from the host star. The rationale for this assumption is that different regions of the disk can efficiently communicate with each other through hy- drodynamical forces and/or self-gravity, such that the disk stays nearly flat. However, to what extent this assumption is valid is uncertain, especially because in the star-disk-binary system the disk experiences two distinct torques from the oblate star and from the binary which tend to drive the disk toward different orientations (see Tremaine & Davis 2014 for examples of non-trivial disk warps when a disk is torqued by different forces). Moreover, the combined effects of disk warps/twists (even if small) and viscosity can lead to non- trivial long-term evolution of the star-disk-binary system. Previous works on warped disks in the bending wave regime have considered a single external torque, such as an ext bi- nary companion (Lubow & Ogilvie 2000; Bate et al. 2000; Foucart & Lai 2014), an inner binary (Facchini, Lodato, & Price 2013; Lodato & Facchini 2013; Foucart & Lai 2014; Zanazzi & Lai 2018a), magnetic torques from the central star (Foucart & Lai 2011), a central spinning black hole (Demianski & Ivanov 1997; Lubow, Ogilvie, & Pringle 2002; Franchini, Lodato, & Facchini 2016; Chakraborty & Bhat- tacharyya 2017), and a system of multiple planets on nearly coplanar orbits (Lubow & Ogilvie 2001). In this paper, we will focus on the hydrodynamics of warped disks in star- disk-binary systems, and will present analytical calculations for the warp amplitudes/profiles and the rate of evolution of disk inclinations due to viscous dissipation associated with these warps/twists. This paper is organized as follows. Section 2 describes the setup and parameters of the star-disk-binary system we study. Section 3 presents all the technical calculations of our paper, including the disk warp/twist profile and effect of vis- cous dissipation on the evolution of system. Section 4 exam- ines how viscous dissipation from disk warps modifies the long-term evolution of star-disk-binary systems. Section 5 discusses theoretical uncertainties of our work. Section 6 contains our conclusions. ion exerts a torque on the disk, driving it into differential precession around the binary angular momentum axis lb. Averaging over the orbital period of the disk annulus and binary, the torque per unit mass is where Ω(r) (cid:39)(cid:112)GM(cid:63)/r3 is the disk angular frequency, l = Tdb = −r2Ωωdb(l·lb)lb×l, (1) l(r, t) is the unit orbital angular momentum axis of a disk "ring" at radius r, and ωdb(r) = 3GM(cid:63) 4a3 bΩ (2) is the characteristic precession frequency of the disk "ring" at radius r. Similarly, the rotation-induced stellar quadrapole drives the stellar spin axis s and the disk onto mutual precession. The stellar rotation leads to a difference in the principal components of the star's moment of inertia (cid:63), where kq (cid:39) 0.1 for fully convective of I3 − I1 = kqM(cid:63)R2 ¯Ω2 stars (Lai, Rasio, & Shapiro 1993). Averaging over the or- bital period of the disk annulus, the torque on the disk from the oblate star is (cid:63) Tds(r, t) = −r2Ωωds(s·l)s×l, where ωds(r) = 3G(I3 − I1) 2r5Ω = 3GkqM(cid:63)R2 (cid:63) ¯Ω2 (cid:63) 2r5Ω (3) (4) is the characteristic precession frequency of the disk ring at radius r. Since ωdb and ωds both depend on r, the disk would quickly lose coherence if there were no internal coupling be- tween the different "rings." We introduce the following rescaled parameters typical of protostellar systems: Ω(cid:63)(cid:112)GM(cid:63)/R3 (cid:63) , ¯Ω(cid:63) = , ¯rout = rout 50 au , ¯M(cid:63) = ¯Md = M(cid:63) 1 M(cid:12) Md 0.1 M(cid:12) Mb 1 M(cid:12) , , ¯R(cid:63) = , ¯rin = , R(cid:63) 2 R(cid:12) rin 8 R(cid:12) ab ¯Mb = . ¯ab = 300 au (5) The rotation periods of T Tauri stars vary from P(cid:63) ∼ 1−10 days (Bouvier 2013), corresponding to ¯Ω(cid:63) ∼ 0.3−0.03. We fix the canonical value of ¯Ω(cid:63) to be 0.1, corresponding to a stellar rotation period of P(cid:63) = 3.3 days. The other canon- ical values in Eq. (5) are unity, except the disk mass, which can change significantly during the disk lifetime. Our choice of rin is motivated by typical values of a T Tauri star's mag- netospheric radius rm, set by the balance of magnetic and plasma stresses (see Lai 2014b for review) (cid:19)1/7 (cid:18) µ4 (cid:19)4/7(cid:18) 10−7 M(cid:12)/yr GM(cid:63) M 2 (cid:63) rin ≈ rm = η (cid:18) B(cid:63) = 7.4 η 1 kG M (cid:19)2/7 ¯R12/7 (cid:63) ¯M 1/7 (cid:63) R(cid:12). (6) 2 STAR-DISK-BINARY SYSTEM AND GRAVITATIONAL TORQUES Consider a central star of mass M(cid:63), radius R(cid:63), rotation rate Ω(cid:63), with a circumstellar disk of mass Md, and inner and outer truncation radii of rin and rout, respectively. This star-disk system is in orbit with a distant binary companion of mass Mb and semimajor axis ab. The binary compan- Here, µ(cid:63) = B(cid:63)R3 (cid:63) is the stellar dipole moment, B(cid:63) is the M is the accretion rate onto the cen- stellar magnetic field, tral star (e.g. Rafikov 2017), and η is a parameter of order unity. We note that we take the stellar radius to be fixed, in contrast to the models of Batygin & Adams (2013) and Spalding & Batygin (2014, 2015), but we argue this will not change our results significantly. We are primarily concerned c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Disk Warp in Star-Disk-Binary Systems 3 interplay between the disk warp/twist and viscous dissipa- tion can lead to appreciable damping of the misalignment between the disk and the external perturber (i.e., the oblate star or the binary companion). In particular, when an ex- ternal torque Text (per unit mass) is applied to a disk in the bending wave regime (which could be either Tdb or Tds), the disk's viscosity causes the disk normal to develop a small twist, of order ∂l ∂ ln r ∼ 4α c2 s Text. (14) (cid:29) The detailed derivation of Eq. (14) is contained in Sec- tions 3.1-3.3. Since Text ∝ lext×l (lext is the axis around which l precesses), where the viscous twist interacts with the external torque, effecting the evolution of l over viscous timescales. To an order of magnitude, we have (cid:12)(cid:12)(cid:12)(cid:12) dl (cid:12)(cid:12)(cid:12)(cid:12)visc dt ∼ (cid:28)(cid:18) 4α (cid:19) T 2 (cid:29) (cid:28) 4α ∼ (15) where ωext is either ωds or ωdb, and (cid:104)···(cid:105) implies proper average over r. c2 s c2 s ext , (r2Ω)ω2 ext r2Ω We now study the disk warp and viscous evolution quantitatively, using the formalism describing the structure and evolution of circular, weakly warped disks in the bend- ing wave regime. The relevant equations have been derived by a number of authors (Papaloizou & Lin 1995; Demianski & Ivanov 1997; Lubow & Ogilvie 2000). We choose the for- malism of Lubow & Ogilvie (2000) and Lubow, Ogilvie, & Pringle (2002) (see also Ogilvie 2006 when ∂l/∂ ln r ∼ 1), where the evolution of the disk is governed by Σr2Ω ∂l ∂t ∂G ∂t = ΣText + (cid:18) Ω2 − κ2 1 r = 2Ω (cid:19) , ∂G ∂r l×G − αΩG + (16) (17) Σc2 s r3Ω 4 ∂l , ∂r where Text is the external torque per unit mass acting on the disk, κ = (2Ω/r)∂(r2Ω)/∂rz=0 is the epicyclic frequency, and G is the internal torque, which arises from slightly eccentric fluid particles with velocities sheared around the disk mid-plane (Demianski & Ivanov 1997). Eq. (16) is the 2D momentum equation generalized to non-coplanar disks. Eq. (17) is related to how internal torques generated from disk warps are communicated across the disk under the influ- ence of viscosity and precession from non-Keplarian epicyclic frequencies. See Nixon & King (2016) for a qualitative dis- cussion and review of Eqs. (16)-(17). We are concerned with two external torques acting on different regions of the disk. For clarity, we break up our calculations into three subsections, considering disk warps produced by individual torques before examining the com- bined effects. (cid:16) rout (cid:17)p r with the effects of viscous dissipation from disk warping, and a changing stellar radius will not affect the viscous torque calculations to follow. We parameterize the disk surface density Σ = Σ(r, t) as Σ(r, t) = Σout(t) . (7) We take p = 1 unless otherwise noted. The disk mass Md is then (assuming rin (cid:28) rout) Md = 2πΣrdr (cid:39) 2πΣoutr2 2 − p out . (8) (cid:90) rout rin The disk angular momentum vector is Ld = Ldld (assum- ing a small disk warp), and stellar spin angular momentum vector is S = S s, where ld and s are unit vectors, and 2πΣr3Ωdr (cid:39) 2 − p 5/2 − p √ GM(cid:63)rout, Md (9) (cid:90) rout rin Ld = S = k(cid:63)M(cid:63)R2 (cid:63)Ω(cid:63). (10) Here k(cid:63) (cid:39) 0.2 for fully convective stars ( e.g. Chan- drasekhar 1939). The binary has orbital angular momentum Lb = Lblb. Because typical star-disk-binary systems satisfy Lb (cid:29) Ld, S, we take lb to be fixed for the remainder of this work. 3 DISK WARPING When α (cid:46) H/r (α is the Shakura-Sunyaev viscosity param- eter, H is the disk scaleheight), which is satisfied for pro- tostellar disks (e.g. Rafikov 2017), the main internal torque enforcing disk rigidity and coherent precession comes from bending wave propagation (Papaloizou & Lin 1995; Lubow & Ogilvie 2000). As bending waves travel at 1/2 of the sound speed, the wave crossing time is of order tbw = 2(r/H)Ω−1. When tbw is longer than the characteristic precession times −1 ω from an external torque, significant disk warps db or ω can be induced. In the extreme nonlinear regime, disk break- ing may be possible (Larwood et al. 1996; Dogan et al. 2015). −1 To compare tbw with ω db , we adopt the disk sound speed profile −1 ds and ω −1 ds (cid:16) rout (cid:17)q (cid:114) GM(cid:63) (cid:17)q rout , r (11) cs(r) = H(r)Ω(r) = hout (cid:114) GM(cid:63) (cid:16) rin rin r = hin where hin = H(rin)/rin and hout = H(rout)/rout. Passively heated disks have q ≈ 0.0 − 0.3 (Chiang & Goldreich 1997), while actively heated disks have q ≈ 3/8 (Lynden-Bell & Pringle 1974). We find (cid:18) 0.1 (cid:18) 0.1 hin (cid:19)(cid:18) kq (cid:19) ¯R2 (cid:18) r (cid:19) ¯Mb ¯r3 (cid:18) r (cid:63) ¯r2 in 0.1 rin (cid:19)q−7/2 (cid:19)q+3/2 hout out ¯M(cid:63)¯a3 b rout . (13) tbwωds = 4.7 × 10 −4 tbwωdb = 1.7 × 10 −2 Thus, we expect the small warp approximation to be valid everywhere in the disk. This expectation is confirmed by our detailed calculation of disk warps presented later in this section. , (12) 3.1 Disk Warp Induced by Binary Companion The torque from an external binary companion is given by Eq. (1). The companion also gives rise to a non-Keplarian epicyclic frequency, given by Ω2 − κ2 2Ω = ωdbP2(l·lb), (18) where Pl are Legendre polynomials. Although the disk is flat to a good approximation, the To make analytic progress, we take advantage of our c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 4 J. J. Zanazzi and Dong Lai Figure 1. The rescaled radial functions [see Eq. (34) for rescaling] τb [Eq. (30)], Vb [Eq. (31)], and Wbb [Eq. (32)]. We take (p, q) values [Eq. (7) and (11)] of p = 0.5 (solid), p = 1.0 (dashed), and p = 1.5 (dotted) with q = 0.0 (blue) and q = 0.5 (red). All other parameters take their cannonical values [Eq. (5)]. The re-scaled radial functions trace out the viscous twist ( Vb) and warp (τb, Wbb) profiles of the disk due to the gravitational torque from the binary companion. p 0.5 1.0 1.5 0.5 1.0 1.5 q 0.5 0.5 0.5 0.0 0.0 0.0 Ub 0.857 1.00 1.20 1.65 1.93 2.31 Vb 0.857 1.00 1.20 3.67 4.26 5.02 Wbb 0.857 1.00 1.20 1.32 1.54 1.85 The physical meaning of ld thus becomes clear: ld is the unit total angular momentum vector of the disk, or Σr3Ωl(r, t)dr. (25) (cid:90) rout rin ld ≡ 2π Ld Using Eqs. (22) and (23), we may solve Eq. (21) for G0(r, t): G0(r, t) = gb(r)(ld·lb)lb×ld, Table 1. Dimensionless coefficients Ub [Eq. (35)], Vb [Eq. (36)], and Wbb [Eq. (37)] tabulated for various p and q values [Eqs. (7) and (11)]. All the parameter values are canonical [Eq. (5)]. When varying q, we fix hout = 0.1. where gb(r) = (ωdb − ωdb)Σr (cid:48)3Ωdr (cid:48) . expectation that ∂l/∂ ln r (cid:28) 1 since tbwωdb (cid:28) 1 [see Eq. (13)]. Specifically, we take l(r, t) = ld(t) + l1(r, t) + . . . , G(r, t) = G0(r, t) + G1(r, t) + . . . , (20) where l1 (cid:28) ld = 1. Here, G0(r, t) is the internal torque maintaining coplanarity of ld(t), G1(r, t) is the internal torque maintaining the leading order warp profile l1(r, t), etc. To leading order, Eq. (16) becomes Σr2Ω dld dt = −Σr2Ωωdb(ld·lb)lb×ld + 1 r ∂G0 ∂r . (21) Integrating (21) over rdr, and using the boundary condition G0(rin, t) = G0(rout, t) = 0, (22) we obtain where ωdb is given by dld dt (cid:90) rout ωdb = 2π Ld ωdbΣr3Ωdr (cid:18) Mb (cid:19)(cid:18) ab M(cid:63) rout (cid:19)3(cid:115) rin (cid:39) 3(5/2 − p) 4(4 − p) = −ωdb(ld·lb)lb×ld, (23) and the constants X0 of the functions X(r) (either τb(r), Vb(r), or Wbb(r)) are determined by requiring Σr3ΩXdr = 0. (33) rin GM(cid:63) r3 out . (24) Notice the radial functions τb, Vb, and Wbb trace out the disk's warp profile l1(r) due to the binary companion's gravitational torque. Because the magnitudes for the radial functions (2π/Myr)τb, Vb, and Wbb are much smaller than c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Using Eqs. (26) and (17), and requiring that l1 not con- tribute to the total disk angular momentum vector, or (19) we obtain the leading order warp l1(r, t): rin Σr3Ωl1(r, t)dr = 0, (28) l1(r, t) = − ωdbτb(ld·lb)2lb×(lb×ld) − Wbb(ld·lb)P2(ld·lb)ld×(lb×ld) + Vb(ld·lb)lb×ld, where τb(r) = Vb(r) = Wbb(r) = 4gb s r(cid:48)3Ω Σc2 4αgb s r(cid:48)3 dr Σc2 4ωdbgb s r(cid:48)3Ω Σc2 (cid:48) − τb0, dr (cid:48) − Vb0, (cid:48) − Wbb0, dr (26) (27) (29) (30) (31) (32) (cid:90) r rin (cid:90) rout (cid:90) r (cid:90) r (cid:90) r rin rin rin (cid:90) rout 10−1100101Radius(au)−0.8−0.6−0.4−0.20.00.20.40.6τb10−1100101Radius(au)−1.0−0.8−0.6−0.4−0.20.00.20.4Vb10−1100101Radius(au)−0.4−0.20.00.20.40.60.8Wbb p 0.5 1.0 1.5 0.5 1.0 1.5 q 0.5 0.5 0.5 0.0 0.0 0.0 Us 2.66 1.00 0.400 24.2 4.28 1.20 Vs 0.315 1.00 6.18 0.0735 0.110 0.207 Wss 0.800 1.00 1.33 0.457 0.533 0.639 Table 2. Dimensionless coefficients Us [Eq. (50)], Vs [Eq. (51)], and Wss [Eq. (52)], for different values of p and q [Eqs. (7) and (11)]. All other parameter values are canonical [Eq. (5)]. When varying q, we fix hout = 0.1. Disk Warp in Star-Disk-Binary Systems 5 = −ωds(ld·s)s×ld, where (assuming rin (cid:28) rout) dld dt (cid:90) rout rin ωds = 2π Ld ωdsΣr3Ωdr ¯Ω2 (cid:63) R2 (cid:63) r1−p out r1+p in 2(1 + p) (cid:39) 3(5/2 − p)kq (cid:18) Md (cid:39) 3(2 − p)kq 2(1 + p)k(cid:63) M(cid:63) (cid:19) ωsd = (Ld/S)ωds (cid:115) (cid:112) (41) GM(cid:63) r3 out , (42) ¯Ω(cid:63) GM(cid:63)R3 (cid:63) r2−p out r1+p in . (43) unity everywhere [see Eqs. (35)-(37)], we define the re-scaled radial function X(r) = τb, Vb, and Wbb as (cid:46)(cid:2)X(rout) − X(rin)(cid:3). X(r) ≡ X(r) Figure 1 plots the dimensionless radial functions τb, Vb, and Wbb for the canonical parameters of the star-disk-binary sys- tem [Eq. (5)]. The scalings of the radial functions evaluated at the outer disk radius are τb(rout) − τb(rin) = −1.82 × 10 −5Ub (34) where , (35) Vb(rout) − Vb(rin) = −1.54 × 10 −3Vb hout × (cid:18) 0.1 ×(cid:16) α (cid:18) 0.1 0.01 × hout ¯M 3/2 out (cid:63) ¯a6 b (cid:19)2 ¯Mb ¯r9/2 (cid:17)(cid:18) 0.1 (cid:19)2 ¯M 2 hout b ¯r6 out ¯M 2 (cid:63) ¯a6 b Myr 2π (cid:19)2 ¯Mb ¯r3 out ¯M(cid:63)¯a3 b Wbb(rout) − Wbb(rin) = −8.93 × 10 −5Wbb . (37) , (36) where With dld/dt and ds/dt determined, Eq. (16) may be inte- grated to obtain the leading order internal torque: G0(r, t) = gs(r)(ld·s)s×ld, (cid:90) r rin gs(r) = (ωds − ωds)Σr (cid:48)3Ωdr (cid:48) . Similarly, the leading order warp profile is l1(r, t) = − ωsdτs(ld·s)2(ld×s)×ld − ωdsτs(ld·s)2s×(s×ld) − Wss(ld·s)P2(ld·s)ld×(s×ld) + Vs(ld·s)s×ld, (44) (45) (46) (47) (48) (49) (cid:90) r (cid:90) r (cid:90) r rin rin rin τs(r) = Vs(r) = Wss(r) = Σc2 4gs s r(cid:48)3Ω 4αgs s r(cid:48)3 dr Σc2 4ωdsgs s r(cid:48)3Ω Σc2 (cid:48) − τs0, dr (cid:48) − Vs0, (cid:48) − Wss0. dr In Figure 2, we plot the rescaled radial functions τs, Vs, and Wss for various p and q values, tracing out the re-scaled warp profile across the radial extent of the disk due to the oblate star's torque. The radial function differences evaluated at the disk's outer and inner truncation radii are τs(rout) − τs(rin) = 2.21 × 10 × (cid:63) ¯r3/2 out ¯M 1/2 Vs(rout) − Vs(rin) = 1.13 × 10 ¯r2 in ¯rin (cid:63) hout −6Us (cid:18) 0.1 (cid:18) ¯Ω(cid:63) (cid:19)2 Myr (cid:19)p−1 ¯R2 (cid:19) ¯R2 (cid:19)2 (cid:18) ¯Ω(cid:63) (cid:19)2(cid:18) kq (cid:18) ¯Ω(cid:63) (cid:19)2 ¯R4 (cid:19)2 0.1 −3Vs 0.1 −7Wss 0.1 2π , (cid:63) ¯r4 in . 0.1 (cid:18) 1358 ¯rout (cid:17)(cid:18) 0.1 ×(cid:16) α (cid:18) kq (cid:19)2(cid:18) 0.1 hin × 0.1 hin (cid:63) ¯r2 in Wss(rout) − Wss(rin) = 4.39 × 10 0.01 (cid:19)2(cid:18) kq (cid:19) 0.1 , (50) (51) (52) Equations (50)-(52) provide an estimate for the misalign- ment angle between the disk's outer and inner orbital an- gular momentum unit vectors l(rout, t)×l(rin, t) due to the oblate star's torque. The dimensionless coefficients Us, Vs, and Wss depend weakly on the parameters p, q, and rin/rout. In Table 2, we tabulate Us, Vs, and Wss for the p and q values indicated, with rin/rout taking the canonical value [Eq. (5)]. Equations (35)-(37) provide an estimate for the mis- alignment angle between the disk's inner and outer or- bital angular momentum vectors, or X(rout) − X(rin) ∼ l(rout, t)×l(rin, t), where X = (2π/Myr)τb, Vb, and Wbb. The dimensionless coefficients Ub, Vb, and Wbb depend weakly on the parameters p, q, and rin/rout. Table 1 tab- ulates Ub, Vb, and Wbb for values of p and q as indicated, with the canonical value of rin/rout [Eq. (5)]. 3.2 Disk Warp Indued by Oblate Star The torque on the disk from the oblate star is given by Eq. (3). The stellar quadrupole moment also gives rise to a non-Keplarian epicyclic frequency given by Ω2 − κ2 2Ω = ωdsP2(l·s). (38) Equations (16)-(17) are coupled with the motion of the host star's spin axis: 2πΣr3Ωωds(s·l)l×s dr, (39) (cid:90) rout (cid:104) rin S ds dt = − (cid:105) Expanding l and G according to Eqs. (19) and (20), inte- grating Eq. (16) over rdr, and using the boundary condi- tion (22), we obtain the leading order evolution equations = −ωsd(s·ld)ld×s, ds dt (40) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 6 J. J. Zanazzi and Dong Lai Figure 2. The rescaled radial functions [see Eq. (34) for rescaling] τs [Eq. (47)], Vs [Eq. (48)], and Wss [Eq. (49)]. We take (p, q) values [Eq. (7) and (11)] of p = 0.5 (solid), p = 1.0 (dashed), and p = 1.5 (dotted) with q = 0.0 (blue) and q = 0.5 (red). Other parameters assume their canonical values [Eq. (5)]. The re-scaled radial functions trace out the viscous twist ( Vs) and warp (τs, Wss) profiles of the disk due to the gravitational torque from the oblate star. p 0.5 1.0 1.5 0.5 1.0 1.5 q 0.5 0.5 0.5 0.0 0.0 0.0 Wbs Wsb 2.13 1.00 0.457 4.57 1.93 0.823 0.917 1.00 1.06 312 319 307 Table 3. Dimensionless coefficients Wbs [Eq. (56)] and Wsb [Eq. (57)] for values of p and q as indicated [Eqs. (7) and (11)]. All parameter values are canonical [Eq. (5)]. When varying q, we fix hout = 0.1. 3.3 Disk Warps Induced by Combined Torques The combined torques from the distant binary and oblate star are given by Eqs. (1) and (3), and the non-Keplarian epicyclic frequencies are given by Eqs. (18) and (38). Using the same procedure as Sections 3.1-3.2, the leading order correction to the disk's warp is (cid:104) (cid:104) (cid:105)lb×ld (cid:105) (s×ld)·lb l1(r, t) = (l1)bin + (l1)star − ωdsτb(ld · s) − ωdsτb(ld·lb)(ld·s)lb×(s×ld) − ωdbτs(ld·lb) s×ld − ωdbτs(ld·s)(ld·lb)s×(lb×ld) − Wsb(ld·lb)P2(ld·s)ld×(lb×ld) − Wbs(ld·s)P2(ld·lb)ld×(s×ld), (lb×ld)·s (53) where (l1)bin is Eq. (29), (l1)star is Eq. (46), τb and τs are given in Eqs. (30) and (47), and Wbs(r) = Wsb(r) = 4ωdbgs s r(cid:48)3Ω Σc2 4ωdsgb s r(cid:48)3Ω Σc2 (cid:48) − Wbs0, dr (cid:48) − Wsb0. dr (54) (55) (cid:90) r (cid:90) r rin rin ternal torque resisting Tds (Tdb) induced by Tdb (Tds). The cross Wbs (Wsb) terms come from the internal torque re- sisting Tds (Tdb) twisted by the non-Keplarian epicyclic fre- quency induced by the binary [Eq. (18)] [star, Eq. (38)]. In Figure 3, we plot the re-scaled radial functions Wbs and Wsb for various p and q values, tracing out the warp profile across the radial extent of the disk due to the combined binary and stellar torques. The radial functions Wbs and Wsb evaluated at the disk's outer and inner truncation radii are hout (cid:18) 0.1 (cid:19)2 (cid:19)2 (cid:18) ¯Ω(cid:63) (cid:19)2 0.1 , (56) Wbs(rout) − Wbs(rin) = −7.23 × 10 (cid:18) kq (cid:18) 0.1 0.1 −6Wbs (cid:19)(cid:18) 1358 ¯rout (cid:19)2(cid:18) kq (cid:19)p−1 ¯Mb ¯R2 (cid:18) ¯Ω(cid:63) (cid:19) ¯Mb ¯R(cid:63) ¯rout (cid:63) ¯r3 out ¯M(cid:63)¯a3 b ¯r2 in −9Wsb ¯rin Wsb(rout) − Wsb(rin) = 1.23 × 10 × × hout 0.1 ¯M(cid:63)¯a3 b 0.1 . (57) These provide an estimate for the misalignment angle be- tween the disk's outer and inner orbital angular momentum unit vectors l(rout, t)×l(rin, t) due to the binary and stellar torques. The dimensionless coefficients Wbs and Wsb depend on the parameters p, q, and rin/rout. Table 3 tabulates Wbs and Wsb for several p and q values, with rin/rout taking the canonical value [Eq. (5)]. 3.4 Disk Warp Profile: Summary In the previous subsections, we have derived semi-analytic expressions for the disk warp profiles due to the combined torques from the oblate host star and the binary companion. Our general conclusion is that the warp is quite small across the whole disk. We illustrate this conclusion with a few ex- amples (Figs. 4-5). We define the disk misalignment angle β = β(r, t) as the misalignment of the disk's local angular momentum unit vector l(r, t) by sin β(r, t) ≡(cid:12)(cid:12)l(r, t)×ld(t)(cid:12)(cid:12), (58) Notice l1 is not simply the sum l1 = (l1)bin + (l1)star. The cross ωdsτb (ωdbτs) terms come from the motion of the in- where ld is unit vector along the total angular momentum of the disk [Eq. (25)]. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10−1100101Radius(au)−1.0−0.8−0.6−0.4−0.20.00.20.4τs10−1100101Radius(au)−1.0−0.8−0.6−0.4−0.20.00.2Vs10−1100101Radius(au)−1.0−0.8−0.6−0.4−0.20.0Wss Disk Warp in Star-Disk-Binary Systems 7 Figure 3. The rescaled radial functions [see Eq. (34) for rescaling] Wbs [Eq. (54)], and Wsb [Eq. (55)]. We take (p, q) values [Eq. (7) and (11)] of p = 0.5 (solid), p = 1.0 (dashed), and p = 1.5 (dotted) with q = 0.0 (blue) and q = 0.5 (red). We take all parameters to be cannonical [Eq. (5)]. The re-scaled radial functions trace out the the warp ( Wbs, Wsb) profiles of the disk due to the interaction between the binary companion and oblate star torques (see text for discussion). Figure 4. Disk misalignment angle β [Eq. (58)] as a function of radius r, for the hin [Eq. (11)] values indicated, all for hout = 0.05 [Eq. (11)]. The disk masses are Md = 0.1 M(cid:12) (solid) and Md = 0.01 M(cid:12) (dashed), with p = 1 [Eq. (7)], α = 0.01, ab = 300 au, and s, ld, and lb lying in the same plane with θsd = θdb = 30◦. Figures 4-5 that the disk warp angle is less than a few degrees for the range of parameters considered. When hin = 0.05, the binary's torque has the strongest influence on the disk's warp profile. As a result, the disk warp (∂β/∂ ln r) is strongest near the disk's outer truncation radius (r (cid:38) 10 au). When hin = 0.01, the spinning star's torque has a strong influence on the disk's warp profile, and the warp becomes large near the inner truncation radius (r (cid:46) 1 au). Notice that the differences between the high disk-mass (Md = 0.1 M(cid:12), solid lines) and low disk-mass (Md = 0.01M(cid:12), dashed lines) are marginal. This is because only the precession rate of the star around the disk ωsd [Eq. (43)] depends on the disk mass, and it enters the disk warp profile only through the term ωsdτs [see Eq. (46)]. Because the disk's c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 5. Same as Fig. 4, except ab = 200 au. internal torque from bending waves is purely hydrodynami- cal, the other terms in the disk warp profile are independent of the disk mass. 3.5 Viscous Evolution As noted above, when a hydrodynamical disk in the bending wave regime is torqued externally, viscosity causes the disk to develop a small twist, which exerts a back-reaction torque on the disk. When torqued by a central oblate star and a distant binary, the leading order viscous twist in the disk is (l1)visc = Vb(lb·ld)lb×ld + Vs(s·ld)s×ld, (59) where Vb and Vs are defined in Eqs. (31) and (48). All other terms in Eq. (53) are non-dissipative, and do not contribute to the alignment evolution of the disk. Inserting (l1)visc into 10−1100101Radius(au)−1.0−0.8−0.6−0.4−0.20.0Wbs10−1100101Radius(au)−0.8−0.6−0.4−0.20.00.20.40.6Wsb10−1100101Radiusr(au)0.00.20.40.60.81.01.2β(degrees)hin=0.05hin=0.0110−1100101Radiusr(au)0.00.20.40.60.81.01.21.41.61.8β(degrees)hin=0.05hin=0.01 8 J. J. Zanazzi and Dong Lai p 0.5 1.0 1.5 0.5 1.0 1.5 q 0.5 0.5 0.5 0.0 0.0 0.0 Γb Γs 0.698 1.00 1.41 1.44 2.31 3.82 0.522 1.00 2.86 0.0964 0.0970 0.108 Γ(bs) 1.70 1.00 0.527 8.64 5.38 3.05 Table 4. Dimensionless viscosity coefficients Γb [Eq. (69)], Γs [Eq. (70)], and Γ(bs) [Eq. (71)], for various p and q values. All other parameter values are canonical [Eq. (5)]. Figure 6. The damping rate γb [Eq. (69)] as a function of the binary semi-major axis ab. We take the p [Eq. (7)] value to be p = 0.5 (solid), p = 1.0 (dashed), and p = 1.5 (dotted), with the q [Eq. (11)] value of q = 0.0 (blue) and q = 0.5 (red). We take all other parameter values to be canonical [Eq. (5)]. When varying q, we fix hout = 0.05 [Eq. (11)]. When the damping rate γb (cid:38) 0.1(2π/Myr), viscous torques from disk warping may significantly decrease the mutual disk-binary inclination θdb [Eq. (79)] over the disk's lifetime. Figure 7. The damping rate γsd [Eq. (72)] as a function of the normalized stellar rotation frequency ¯Ω(cid:63) [Eq. (5)]. We take the p [Eq. (7)] values to be p = 0.5 (solid), p = 1.0 (dashed), and p = 1.5 (dotted), with q [Eq. (11)] values of q = 0.0 (blue) and q = 0.5 (red). We take all other parameter values to be canonical [Eq. (5)]. When varying q, we fix hin = 0.03 [Eq. (11)]. When the damping rate γsd (cid:38) 0.1(2π/Myr), viscous torques from disk warp- ing may significantly decrease the mutual star-disk inclination θsd [Eq. (77)] over the disk's lifetime. rin (cid:90) rout (cid:90) rout (cid:90) rout (cid:90) rout (cid:90) rout rin rin rin rin γs ≡ 2π Ld = − 2π Ld γ(bs) ≡ 2π Ld = − 2π Ld = − 2π Ld 4αg2 s s r3 dr Σc2 Σr3Ω(ωds − ωds)Vsdr, 4αgbgs s r3 dr Σc2 Σr3Ω(ωds − ωds)Vbdr (63) Σr3Ω(ωdb − ωdb)Vsdr. (64) Eqs. (16) and (39), and integrating over 2πrdr, we obtain (cid:18) dLd (cid:19) dt visc (cid:18) dS (cid:19) dt visc where = Ldγb(ld·lb)2lb×(lb×ld) + Ldγs(ld·s)2s×(s×ld) + Ldγ(bs)(ld·lb)(ld·s)lb×(s×ld) + Ldγ(bs)(ld·s)(ld·lb)s×(lb×ld), = − Ldγs(ld·s)2s×(s×ld) − Ldγ(bs)(ld·s)(ld·lb)s×(lb×ld), (cid:90) rout (cid:90) rout rin rin γb ≡ 2π Ld = − 2π Ld 4αg2 b s r3 dr Σc2 Σr3Ω(ωdb − ωdb)Vbdr, (60) (61) (62) When deriving Eqs. (60) and (61), we have neglected terms proportional to l1·s or l1·lb, as these only modify the dy- namics by changing the star-disk and disk-binary preces- sional frequencies, respectively. Using − ld (cid:19) (65) = , (cid:18) dLd (cid:18) dS dt − s dt dLd dt (cid:19) dS dt dld dt ds dt = 1 Ld 1 S , (66) the leading order effect of viscous disk twisting on the time evolution of ld and s is (cid:18) dld (cid:19) dt visc = γb(ld·lb)3ld×(lb×ld) + γs(ld·s)3ld×(s×ld) + γ(bs)(ld·lb)(ld·s)2ld×(lb×ld) + γ(bs)(ld·s)(ld·lb)2ld×(s×ld), (67) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 200400800ab(au)10−610−510−410−310−210−1100101γb(2π/Myr)EfficientAlignment0.050.10.2¯Ω?10−510−410−310−210−1100101102103104γsd(2π/Myr)EfficientAlignment Disk Warp in Star-Disk-Binary Systems 9 function of the binary semi-major axis ab. In agreement with Foucart & Lai (2014), we find the damping rate to be small, and weakly dependent on the power-law surface density and sound-speed indices p and q. This is because the torque from the binary companion is strongest around r ∼ rout. The properties of the disk near rout are "global," since the amount of inertia of disk annuli near rout is set mainly by the total disk mass rather than the surface den- sity profile, and the disk sound-speed does not vary greatly around r ∼ rout. We conclude that viscous torques from disk warping are unlikely to significantly decrease the mu- tual disk-binary inclination θdb unless ab (cid:46) 200 au. Figure 7 plots the star-disk alignment rate γsd as a func- tion of the normalized stellar rotation frequency ¯Ω(cid:63). Un- like the disk-binary alignment rate γb (Fig. 6), γsd depends strongly on the surface density and sound-speed power-law indices p and q. The alignment rate of a circumbinary disk with its binary orbital plane has a similarly strong depen- dence on p and q (Foucart & Lai 2013, 2014; Lubow & Mar- tin 2018). This strong dependence arises because the torque on the inner part of a disk from an oblate star or binary is strongest near rin. The disk properties near r ∼ rin are very local (both the amount of inertia for disk annuli and disk sound-speed), and hence will depend heavily on p and q. Despite this uncertainty, Figure 7 shows that there are reasonable parameters for which viscous torques from disk warping can significantly reduce the star-disk inclination θsd [when γsd (cid:38) 0.1(2π/Myr)], especially when the stellar rota- tion rate is sufficiently high ( ¯Ω(cid:63) (cid:38) 0.2). (cid:18) ds (cid:19) dt visc = − Ld S − Ld S γs(ld·s)2s×(s×ld) γ(bs)(ld·s)(ld·lb)s×(lb×ld). (68) The four terms in (dld/dt)visc [Eq. (67)] arises from four different back-reaction torques of the disk in response to Tds [Eq. (3)] and Tdb [Eq. (1)]. To resist the influence of the two external torques Tds and Tdb, the disk develops two twists (∂l/∂ ln r)ds and (∂l/∂ ln r)db, given by Eqs. (46) and (29). The terms in Eqs. (67)-(68) proportional to γs arise from the back reaction of (∂l/∂ ln r)ds to Tds, and works to align s with ld. The term in Eq. (67) proportional to γb arises from the back reaction of Tdb to (∂l/∂ ln r)db , and works to align ld with lb. Because γ(bs) < 0, the terms in Eqs. (67)-(68) proportional to γ(bs) have different effects than the terms proportional to γs and γb. One of the terms in Eqs. (67)- (68) proportional to γ(bs) arises from the back reaction of Tds to (∂l/∂ ln r)db, and works to drive ld perpendicular to s, while the other arises from the back-reaction of Tdb to (∂l/∂ ln r)ds, and works to drive ld perpendicular to lb. Although typically γs > γ(bs) or γb > γ(bs) (so the dynamical effect of γ(bs) may be absorbed into γb and γs), the magnitude of γ(bs) is not negligible compared to γs and γb. For completeness, we include the effects of the γ(bs) terms in the analysis below. The damping rates (62)-(64) may be evaluated and (cid:17)(cid:18) 0.1 (cid:19)2 0.01 hout rescaled to give γb = 1.26 × 10 × ¯M 2 ¯a6 b b ¯r9/2 out ¯M 3/2 γs = 2.04 × 10 (cid:63) (cid:18) kq 0.1 −9Γb (cid:16) α (cid:18) 2π (cid:19) (cid:16) α (cid:19)2 ¯M 1/2 yr −10Γs , 0.01 ¯R4 (cid:63) (cid:63) in ¯r3/2 ¯r4 −10Γ(bs) out × × γ(bs) = − 2.04 × 10 (cid:18) kq (cid:19) ¯Mb ¯R2 0.1 ¯M 1/2 (cid:63) ¯a3 0.1 yr (cid:63) ¯r1/2 out b ¯r2 in ¯rin hin (cid:19)p−1 (cid:17)(cid:18) 0.1 (cid:19)2(cid:18) 1358 ¯rout (cid:19) (cid:19)4(cid:18) 2π (cid:18) ¯Ω(cid:63) (cid:17)(cid:18) 0.1 (cid:19)2(cid:18) 1358 ¯rout (cid:16) α (cid:19) (cid:19)2(cid:18) 2π (cid:18) ¯Ω(cid:63) hout 0.01 0.1 ¯rin yr , , (70) (cid:19)p−1 (71) where hin = (rin/rout)q−1/2hout. The rescaling above has removed the strongest dependencies of the damping rates on p, q, and rin/rout. Table 4 lists values of the dimensionless viscous coefficients Γb, Γs, and Γ(bs), varying p and q. Note that there are "mixed" terms in Eqs. (67)-(68): the counter-aligment rate of ld and lb depends on s, while the counter-alignment rate of ld and s depends on lb. Also note that net spin-disk alignment rate is given by Ld S 1 + γsd = γs. Assuming Ld (cid:29) S, γsd evaluates to be γsd (cid:39) 7.52 × 10 (cid:16) α −9 (2 − p)Γs 5/2 − p (cid:18) 2kq (cid:19)(cid:18) kq (cid:19) k(cid:63) 0.1 × 0.01 ¯Md ¯R7/2 (cid:63) (cid:63) ¯r4 in ¯rout ¯M 1/2 (cid:19) (cid:17)(cid:18) 0.1 (cid:19)2(cid:18) 1358¯rout (cid:19) (cid:19)3(cid:18) 2π (cid:18) ¯Ω(cid:63) hin ¯rin (72) (cid:19)p−1 0.1 yr . (73) (cid:18) Figure 6 plots the disk-binary damping rate γb as a c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 (69) 4 EVOLUTION OF THE STAR-DISK-BINARY SYSTEM WITH VISCOUS DISSIPATION FROM DISK WARPING This section investigates the evolution of star-disk-binary systems under gravitational and viscous torques: = − ωsd(s·ld)ld×s + = − ωds(ld·s)s×ld ds dt dld dt − ωdb(ld·lb)lb×ld + (cid:18) ds (cid:19) (cid:18) dld dt dt , (74) visc (cid:19) . visc (75) The viscous terms are given by Eqs. (67)-(68). As in Batygin & Adams (2013) and Lai (2014a), we assume the disk's mass is depleted according to Md(t) = Md0 1 + t/tv , (76) where Md0 = 0.1 M(cid:12) and tv = 0.5 Myr. See Lai (2014a) and Zanazzi & Lai (2018b) for discussions on the dynamical evolution of s and ld and secular resonance (ωsd ∼ ωdb) when viscous dissipation from disk warping is neglected. The effect of the γs term on the dynamical evolution of s over viscous timescales depends on the precessional dynam- ics of the star-disk-binary system. If ωsd (cid:29) ωdb, s rapidly precesses around ld, and the γs term works to align s with ld. If ωsd (cid:28) ωdb, s cannot "follow" the rapidly varying ld, and effectively precesses around lb. In the latter case, because of the rapid variation of ld around lb, s is only effected by the secular ld. As a result, γs works to drive θsb to θdb. The 10 J. J. Zanazzi and Dong Lai Figure 8. Inclination evolution of star-disk-binary systems. The top panels and bottom left panel plot the time evolution of the angles θsd [Eq. (77)], θsb [Eq. (78)], and θdb [Eq. (79)], integrated using Eqs. (74) and (75), with values of α and hin [Eq. (11)] as indicated. The bottom right panel shows the precession frequencies ωsd [Eq. (43)] and ωdb [Eq. (24)]. We take θdb(0) = 60◦, θsd(0) = 5◦, and hout = 0.05 [Eq. (11)]. The damping rates are γb = 5.05 × 10−9(2π/yr) [Eq. (69)], γsd(0) = 2.00 × 10−7(2π/yr) [Eq. (72)], and γbs = −8.18 × 10−10(2π/yr) [Eq. (71)] for hin = 0.05, and γb = 7.12 × 10−9(2π/yr), γsd(0) = 1.37 × 10−6(2π/yr), and γ(bs) = −1.51 × 10−9(2π/yr) for hin = 0.01. effect of the γb term is simpler: it always works to align ld with lb. Figure 8 shows several examples of the evolution of star- disk-binary systems. The top panels and bottom left panel of Fig. 8 show the time evolution of the angles θsd = cos θsb = cos θdb = cos −1(s·ld), −1(s·lb), −1(ld·lb), (77) (78) (79) from integrating Eqs. (74)-(75), while the bottom right panel plots the characteristic precession frequencies ωsd and ωdb. The top left panel of Fig. 8 does not include viscous torques (α = 0). Because the damping rates γb [Eq. (69)] and γsd [Eq. (72)] are much less than 0.1(2π/Myr) over most of the system's lifetime (10 Myr), viscous torques have a negligible effect on the evolution of θsd, θsb, and θdb. The bottom left panel of Fig. 8 shows the evolution of θsd, θsb, and θdb with α = 0.01 and hin = 0.01. Because the inner edge of the disc has a much smaller scaleheight, the oblate star warps the inner edge of the disk more [Eq. (14)], resulting in γsd taking a value larger than 0.1(2π/Myr). This increase in γsd causes a much tighter coupling of s to ld before secular resonance (ωsd (cid:38) ωdb), evidenced by the damped oscillations in θsd. After secular resonance (ωsd (cid:46) ωdb), the γs term damps s toward lb. Notice θsb approaches θdb because of the rapid precession of ld around lb after secular resonance, not θsb → 0. To gain insight to how the disk warp evolves during the star-disk-binary system's evolution, we introduce the mis- alignment angle ∆β between the disk's outer and inner or- bital angular momentum unit vectors: sin ∆β(t) =(cid:12)(cid:12)l(rout, t)×l(rin, t)(cid:12)(cid:12) (cid:39)(cid:12)(cid:12)[l1(rout, t) − l1(rin, t)]×ld(t)(cid:12)(cid:12) (80) Figure 9 plots ∆β as a function of time, for the examples considered in Fig. 8. We see even when viscous torques from disk warping significantly alter the star-disk-binary system dynamics (e.g. α = 0.01 and hin = 0.01), ∆β < 1.2◦ over the disk's lifetime, indicating a high degree of disk coplanarity throughout the system's evolution. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 0246810020406080100120140160180θsd,θsb,θdb(degrees)α=0θsdθsbθdb0246810020406080100120140160180θsd,θsb,θdb(degrees)α=0.01,hin=0.05θsdθsbθdb0246810Time(Myr)020406080100120140160180θsd,θsb,θdb(degrees)α=0.01,hin=0.01θsdθsbθdb0246810Time(Myr)100101102ωsd,ωdb(2π/Myr)ωsdωdb Disk Warp in Star-Disk-Binary Systems 11 through resonant Lindblad torques (Lubow 1991) or the Lidov-Kozai effect (Martin et al. 2014; Fu, Lubow, & Martin 2015a; Zanazzi & Lai 2017; Lubow & Ogilvie 2017). Lind- blad torques only cause eccentricity growth where the binary orbital frequency is commensurate with the disk orbital fre- quency, so they are unlikely to be relevant unless the outer edge of the disk is close to tidal truncation by the binary companion. Lidov-Kozai oscillations are a much more likely culprit for causing eccentricity growth of circumstellar disks in binaries when θdb (cid:38) 40◦. Lidov-Kozai oscillations may be suppressed by the disk's self-gravity when (Fu, Lubow, & Martin 2015b) Md (cid:38) 0.04 Mb , (81) (cid:18) 3rout (cid:19)3 ab and by the disk's pressure gradients when (Zanazzi & Lai 2017; Lubow & Ogilvie 2017) (cid:18) Mb (cid:19)1/3(cid:18) hout (cid:19)−2/3 M(cid:63) 0.1 ab (cid:38) 4.2 rout . (82) For our canonical parameters [Eq. (5)], the Lidov-Kozai ef- fect is unlikely to be relevant unless ab (cid:46) 4rout. 5.2 Observational Implications In our companion work (Zanazzi & Lai 2018b), we show that the formation of a short-period (orbital periods less than 10 days) massive planet in many instances significantly reduces or completely suppresses primordial misalignments gener- ated by the gravitational torque from an inclined binary companion. Primordial misalignments are still robustly gen- erated in protostellar systems forming low-mass (∼ 1 M⊕) multiple planets, and systems with cold (orbital periods greater than one year) Jupiters. On the other hand, ob- servations suggest that most Kepler compact multi-planet systems have small stellar obliquities (e.g. Albrecht et al. 2013; Winn et al. 2017). A major goal of this work was to examine if viscous torques from disk warping may reduce or suppress the generation of primordial misalignments in star-disk-binary systems. We find that for some parameters, the star-disk inclination damping rate can be significant (see Fig. 7); in particular, the star-disk misalignment may be re- duced when the disk is sufficiently cold with strong external torques (Figs. 8 & 10). Observational evidence is mounting which suggests hot stars (effective temperatures (cid:38) 6000◦ K) have higher obliq- uities than cold stars (Winn et al. 2010; Albrecht et al. 2012; Mazeh et al. 2015; Li & Winn 2016). Since all damping rates from viscous disk-warping torques in star-disk-binary sys- tems are inversely proportional to the disk's sound-speed squared [see Eqs. (69)-(72)], a tempting explanation for this correlation is that hot stars have hot disks with low damp- ing rates which remain misaligned, while cold stars have cold disks with high damping rates which have star-disk misalignments significantly reduced over the disk's lifetime. However, we do not believe this is a likely explanation, since the protostellar disk's temperature should not vary strongly with the T-Tauri stellar mass. If a disk is passively heated from irradiation by its young host star (Chiang & Goldreich 1997), low mass ((cid:46) 3 M(cid:12)) pre-main sequence stars have ef- fective temperatures which are not strongly correlated with Figure 9. Total disk warp ∆β [Eq. (80)] for the integra- tions of Fig. 8. The blue curve denotes the integration where (hin, α) = (0.05, 0.0), the red is (hin, α) = (0.05, 0.01), and the green is (hin, α) = (0.01, 0.01). All other parameters are listed in Fig. 8. All examples considered have ∆β < 1.2◦, indicating the disk remains highly coplanar throughout the system's evolution. Notice ∆β (cid:28) 1◦ when α = 0 (blue, hugs the x-axis). Figure 10 is identical to Fig. 8, except we take ab = 200 au instead of ab = 300 au. Since γb is greater than 0.1(2π/Myr), ld aligns with γb over the disk's lifetime. In the top right panel of Fig. 10, γsd is less than 0.1(2π/Myr) for most of the disk's lifetime, so s stays misaligned with both ld and lb. At the end of the disk's lifetime, s precesses around lb, which is aligned with ld. In the bottom left panel, both γb and γsd are greater than 0.1(2π/Myr) for most of the disk's lifetime. This results in alignment of ld, s, and lb over 10 Myr. Figure 11 shows the evolution of disk misalign- ment angles for the examples considered in Fig. 10. We see ∆β < 1.9◦ for all examples considered, indicating the disk remains highly co-planar throughout the system's evolution. 5 DISCUSSION 5.1 Theoretical Uncertainties Our study of warped disks in star-disk-binary systems relies critically on the warp evolution equations derived in Lubow & Ogilvie (2000) for disks in the bending wave regime (α (cid:46) H/r), assuming a small disk warp (∂l/∂ ln r (cid:28) 1). A non- linear disk warp will change the surface density evolution of the disk through advection and viscosity where the warp is strongest (e.g. Ogilvie 1999; Tremaine & Davis 2014). In addition, even a small warp may interact resonantly with inertial waves, resulting in a parametric instability which enhances the disk's dissipation rate (Gammie, Goodman, & Ogilvie 2000; Ogilvie & Latter 2013). Because we have found for typical parameters, the warp in the disk torqued externally by a central oblate star and distant binary is small [see Eqs. (35)-(37), (50)-(52), and (54)-(55)], such effects are unlikely to change the main results of this paper. In this study, we have assumed that the circumstellar disk in a binary system is circular. This may not be a valid assumption, as the disk may undergo eccentricity growth c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 0246810Time(Myr)0.00.20.40.60.81.01.2∆β(degrees) 12 J. J. Zanazzi and Dong Lai Figure 10. Same as Figure 8, except ab = 200 AU. The damping rates are γb = 5.75 × 10−8(2π/yr), γsd(0) = 2.00 × 10−7(2π/yr), and γ(bs) = −2.76× 10−9(2π/yr) for hin = 0.05, and γb = 8.11× 10−8(2π/yr), γsd(0) = 1.37× 10−6(2π/yr), and γ(bs) = −5.10× 10−9(2π/yr) for hin = 0.01. their masses (Hayashi 1961). If the disk is actively heated by turbulent viscosity (Lynden-Bell & Pringle 1974), the disk's accretion rate does not vary enough between different host star masses to create a difference in disk temperature (Rafikov 2017). Even in systems where viscous torques from disk warp- ing alter the dynamics of the star-disk-binary system over the disk's lifetime (Figs. 8 & 10), we find the misalignment angle between the outer and inner disk orbital angular mo- mentum unit vectors to not exceed a few degrees (Figs. 9 & 11). Therefore, it is unlikely that the disk warp profile plays a role in setting the mutual inclinations of forming exoplanetary systems with inclined binary companions. 6 CONCLUSIONS We have studied how disk warps and the associated vis- cous dissipation affect the evolution of star-disk inclinations in binary systems. Our calculation of the disk warp pro- file shows that when the circumstellar disk is torqued by both the exterior companion and the central oblate star, the deviation of the disk angular momentum unit vector from coplanarity is less than a few degrees for the entire c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 11. Same as Fig. 9, except for the examples considered in Fig. 10. All examples considered have ∆β < 1.9◦, indicating the disk remains highly coplanar throughout the disk's lifetime. 0246810020406080100120140160180θsd,θsb,θdb(degrees)α=0θsdθsbθdb0246810020406080100120140160180θsd,θsb,θdb(degrees)α=0.01,hin=0.05θsdθsbθdb0246810Time(Myr)020406080100120140160180θsd,θsb,θdb(degrees)α=0.01,hin=0.01θsdθsbθdb0246810Time(Myr)100101102ωsd,ωdb(2π/Myr)ωsdωdb0246810Time(Myr)0.00.51.01.52.0∆β(degrees) parameter space considered (Figs. 9 & 11). This indicates that disk warping in star-disk-binary systems does not al- ter exoplanetary architectures while the planets are form- ing in the disk. We have derived analytical expressions for the viscous damping rates of relative inclinations (Sec. 3.5), and have examined how viscous dissipation affects the in- clination evolution of star-disk-binary systems. Because the star-disk [Eq. (72), Fig. 7] and disk-binary [Eq. (69), Fig. 6] alignment timescales are typically longer than the proto- planetary disk's lifetime ((cid:46) 10 Myrs), viscous dissipation from disk warping does not significantly modify the long- term inclination evolution of most star-disk-binary systems (Fig. 8, top left panel). However, in sufficiently cold disks (small H/r) with strong external torques from the oblate star or inclined binary companion, the star-disk-binary evo- lution may be altered by viscous dissipation from disk warp- ing, reducing the star-disk misalignment generated by star- disk-binary interactions (Figs. 8 & 10). In particular, we find when the stellar rotatation rate is sufficiently high (rotation periods (cid:46) 2 days), the star-disk damping is particularly ef- ficient (Fig. 7). This viscous damping may explain the ob- served spin-orbit alignment in some multiplanetary systems (e.g. Albrecht et al. 2013; Winn et al. 2017) in the presence of inclined binary companions. ACKNOWLEDGEMENTS We thank the referee, Christopher Spalding, for many com- ments which improved the presentation and clarity of this work. JZ thanks Re'em Sari and Yoram Lithwick for helpful discussions. This work has been supported in part by NASA grants NNX14AG94G and NNX14AP31G, and NSF grant AST- 1715246. JZ is supported by a NASA Earth and Space Sciences Fellowship in Astrophysics. REFERENCES Albrecht S., Winn J. N., Marcy G. W., Howard A. W., Isaacson H., Johnson J. A., 2013, ApJ, 771, 11 Albrecht S., et al., 2012, ApJ, 757, 18 Bate M. R., Bonnell I. A., Bromm V., 2003, MNRAS, 339, 577 Bate M. R., Bonnell I. A., Clarke C. J., Lubow S. H., Ogilvie G. I., Pringle J. E., Tout C. A., 2000, MNRAS, 317, 773 Batygin K., 2012, Natur, 491, 418 Batygin K., Adams F. C., 2013, ApJ, 778, 169 Bouvier J., 2013, EAS, 62, 143 Brinch C., Jørgensen J. K., Hogerheijde M. R., Nelson R. P., Gressel O., 2016, ApJ, 830, L16 Chandrasekhar S., 1939, An introduction to the study of stellar structure, The University of Chicago press Chakraborty C., Bhattacharyya S., 2017, MNRAS, 469, 3062 Chiang E. I., Goldreich P., 1997, ApJ, 490, 368 Demianski M., Ivanov P. B., 1997, A&A, 324, 829 Dogan S., Nixon C., King A., Price D. J., 2015, MNRAS, 449, 1251 Franchini A., Lodato G., Facchini S., 2016, MNRAS, 455, 1946 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Disk Warp in Star-Disk-Binary Systems 13 Facchini S., Lodato G., Price D. J., 2013, MNRAS, 433, 2142 Fern´andez-L´opez M., Zapata L. A., Gabbasov R., 2017, ApJ, 845, 10 Foucart F., Lai D., 2014, MNRAS, 445, 1731 Foucart F., Lai D., 2013, ApJ, 764, 106 Foucart F., Lai D., 2011, MNRAS, 412, 2799 Fu W., Lubow S. H., Martin R. G., 2015, ApJ, 807, 75 Fu W., Lubow S. H., Martin R. G., 2015, ApJ, 813, 105 Gammie C. F., Goodman J., Ogilvie G. I., 2000, MNRAS, 318, 1005 Hayashi C., 1961, PASJ, 13, Jensen E. L. N., Akeson R., 2014, Natur, 511, 567 Klessen R. S., 2011, EAS, 51, 133 Lai, D. 2014a, MNRAS, 440, 3532 Lai D., 2014b, EPJWC, 64, 01001 Lai D., Rasio F. A., Shapiro S. L., 1993, ApJS, 88, 205 Larwood J. D., Nelson R. P., Papaloizou J. C. B., Terquem C., 1996, MNRAS, 282, 597 Lee J.-E., Lee S., Dunham M. M., Tatematsu K., Choi M., Bergin E. A., Evans N. J., 2017, NatAs, 1, 0172 Li G., Winn J. N., 2016, ApJ, 818, 5 Lodato G., Facchini S., 2013, MNRAS, 433, 2157 Lubow S. H., Martin R. G., 2018, MNRAS, 473, 3733 Lubow S. H., Ogilvie G. I., 2017, MNRAS, 469, 4292 Lubow S. H., Ogilvie G. I., Pringle J. E., 2002, MNRAS, 337, 706 Lubow S. H., Ogilvie G. I., 2001, ApJ, 560, 997 Lubow S. H., Ogilvie G. I., 2000, ApJ, 538, 326 Lubow S. H., 1991, ApJ, 381, 259 Lynden-Bell D., Pringle J. E., 1974, MNRAS, 168, 603 Martin R. G., Nixon C., Lubow S. H., Armitage P. J., Price D. J., Dogan S., King A., 2014, ApJ, 792, L33 Mazeh T., Perets H. B., McQuillan A., Goldstein E. S., 2015, ApJ, 801, 3 McKee C. F., Ostriker E. C., 2007, ARA&A, 45, 565 Neuhauser R., et al., 2009, A&A, 496, 777 Nixon C., King A., 2016, LNP, 905, 45 Ogilvie G. I., Latter H. N., 2013, MNRAS, 433, 2420 Ogilvie G. I., 2006, MNRAS, 365, 977 Ogilvie G. I., 1999, MNRAS, 304, 557 Papaloizou J. C. B., Lin D. N. C., 1995, ApJ, 438, 841 Rafikov R. R., 2017, ApJ, 837, 163 Spalding, C., & Batygin, K. 2015, ApJ, 811, 82 Spalding C., Batygin K., 2014, ApJ, 790, 42 Stapelfeldt K. R., M´enard F., Watson A. M., Krist J. E., Dougados C., Padgett D. L., Brandner W., 2003, ApJ, 589, 410 Stapelfeldt K. R., Krist J. E., M´enard F., Bouvier J., Pad- gett D. L., Burrows C. J., 1998, ApJ, 502, L65 Tanaka H., Ward W. R., 2004, ApJ, 602, 388 Tremaine S., Davis S. W., 2014, MNRAS, 441, 1408 Triaud A. H. M. J., 2017, arXiv, arXiv:1709.06376 Williams J. P., et al., 2014, ApJ, 796, 120 Winn J. N., et al., 2017, AJ, 154, 270 Winn J. N., Fabrycky D. C., 2015, ARA&A, 53, 409 Winn J. N., Fabrycky D., Albrecht S., Johnson J. A., 2010, ApJ, 718, L145 Zanazzi J. J., Lai D., 2017, MNRAS, 467, 1957 Zanazzi J. J., Lai D., 2018a, MNRAS, 473, 603 Zanazzi J. J., Lai D., 2018b, arXiv, arXiv:1711.03138
1504.03237
1
1504
2015-04-13T16:10:45
A reassessment of the in situ formation of close-in super-Earths
[ "astro-ph.EP" ]
A large fraction of stars host one or multiple close-in super-Earth planets. There is an active debate about whether these planets formed in situ or at greater distances from the central star and migrated to their current position. It has been shown that part of their observed properties (e.g., eccentricity distribution) can be reproduced by N-body simulations of in situ formation starting with a population of protoplanets of high masses and neglecting the effects of the disk gas. We plan to reassess the in situ formation of close-in super-Earths through more complete simulations. We performed N-body simulations of a population of small planetary embryos and planetesimals that include the effects of disk-planet interactions (e.g., eccentricity damping, type I migration). In addition, we also consider the accretion of a primitive atmosphere from a protoplanetary disk. We find that planetary embryos grow very quickly well before the gas dispersal, and thus undergo rapid inward migration, which means that one cannot neglect the effects of a gas disk when considering the in-situ formation of close-in super-Earths. Owing to their rapid inward migration, super-Earths reach a compact configuration near the disk's inner edge whose distribution of orbital parameters matches the observed close-in super-Earths population poorly. On the other hand, simulations including eccentricity damping, but no type I migration, reproduce the observed distributions better. Including the accretion of an atmosphere does not help reproduce the bulk architecture of observations. Interestingly, we find that the massive embryos can migrate inside the disk edge while capturing only a moderately massive hydrogen/helium atmosphere. By this process they avoid becoming giant planets. The bulk of close-in super-Earths cannot form in situ, unless type I migration is suppressed in the entire disk inside 1 AU.
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. ogihara_etal_2015 July 11, 2021 c(cid:13)ESO 2021 5 1 0 2 r p A 3 1 . ] P E h p - o r t s a [ 1 v 7 3 2 3 0 . 4 0 5 1 : v i X r a A reassessment of the in situ formation of close-in super-Earths Masahiro Ogihara, Alessandro Morbidelli, and Tristan Guillot Observatoire de la Côte d'Azur, Boulevard de l'Observatoire, 06304 Nice Cedex 4, France e-mail: [email protected] Received 13 February 2015 / Accepted 26 March 2015 ABSTRACT Context. A large fraction of stars host one or multiple close-in super-Earth planets. There is an active debate about whether these planets formed in situ or at greater distances from the central star and migrated to their current position. It has been shown that part of their observed properties (e.g., eccentricity distribution) can be reproduced by N-body simulations of in situ formation starting with a population of protoplanets of high masses and neglecting the effects of the disk gas. Aims. We plan to reassess the in situ formation of close-in super-Earths through more complete simulations. Methods. We performed N-body simulations of a population of small planetary embryos and planetesimals that include the effects of disk-planet interactions (e.g., eccentricity damping, type I migration). In addition, we also consider the accretion of a primitive atmosphere from a protoplanetary disk. Results. We find that planetary embryos grow very quickly well before the gas dispersal, and thus undergo rapid inward migra- tion, which means that one cannot neglect the effects of a gas disk when considering the in-situ formation of close-in super-Earths. Owing to their rapid inward migration, super-Earths reach a compact configuration near the disk's inner edge whose distribution of orbital parameters matches the observed close-in super-Earths population poorly. On the other hand, simulations including eccentric- ity damping, but no type I migration, reproduce the observed distributions better. Including the accretion of an atmosphere does not help reproduce the bulk architecture of observations. Interestingly, we find that the massive embryos can migrate inside the disk edge while capturing only a moderately massive hydrogen/helium atmosphere. By this process they avoid becoming giant planets. Conclusions. The bulk of close-in super-Earths cannot form in situ, unless type I migration is suppressed in the entire disk inside 1 AU. Key words. Planets and satellites: formation -- Planets and satellites: atmospheres -- Planet-disk interactions -- Methods: numerical 1. Introduction Recent observations of exoplanets have revealed a large num- ber of close-in low-mass planets (e.g., Schneider et al. 2011; Wright et al. 2011). As of January 2015, 337 systems harbor 839 planets with masses M < 100M⊕ (or with radii R < 10R⊕) and semimajor axes a < 1AU (or with orbital periods P < 200day). We define these as "close-in super-Earths." They have a semimajor-axis distribution centered on 0.1 AU. Super-Earths in each system are generally confined within a few tenths of an AU. Although some of these close-in super- Earths are in or near first-order mean motion resonances (es- pecially in 3:2 resonances), a large number of close-in super- Earths are not in mean motion resonances (e.g., Mayor et al. 2009; Lissauer et al. 2011b). Typical orbital separations between these planets are 10 − 30rH, where rH is the mutual Hill radius, which is similar to orbital separations between the solar system terrestrial planets. Period ratios of adjacent pairs lie between the 4:3 and 3:1 resonances, respectively. It is estimated that a large number of super-Earths have small eccentricities e ∼ 0.01 − 0.1, while some of them can have high eccentricities up to e ∼ 0.5 (e.g., Moorhead et al. 2011). The mutual inclinations between planetary orbits could be estimated in a fraction of transiting systems from the Kepler catalog and appear to be low with an average of i ∼ 0.03 (Fabrycky et al. 2014). Hansen & Murray (2012, 2013) present N-body simulations of the in situ formation of close-in super-Earths from embryos that are placed between 0.05 AU and 1 AU. To account for the masses of the known planetary systems, they had to assume that up to 100 Earth masses of solids existed in the disk within 1 AU, implying a surface density of solids much higher than for the minimum mass solar nebula (MMSN) (Weidenschilling 1977; Hayashi 1981). Assuming a nominal gas/solid density ratio of 100, this would imply a very massive protoplanetary disk that would probably be gravitationally unstable. However, Chatterjee & Tan (2014, 2015) propose that the disk may be en- riched in solids relative to the gas thanks to the inward migration of dust grains, pebbles, and planetesimals. Thus, it may be legit- imate to assume a very high surface density of solids embedded in the inner protoplanetary disk with a mass of gas comparable to gas in the MMSN model. The simulations of Hansen & Murray (2012, 2013), never- theless, are quite simplistic. They start from a population of protoplanets of high masses (sometimes multiple Earth masses), which are supposed to have already achieved the completion of the oligarchic growth process (Kokubo & Ida 1998). No plan- etesimals are considered, and the effects of the gas, in relation to the migration and eccentricity or to the inclination damping of the protoplanets' orbits, are not taken into account. With these assumptions, their simulations suggest that an in situ accretion without gas can explain the distributions of orbital periods and eccentricities of observed super-Earths. However, their simula- tions did not reproduce the distribution of period ratios between adjacent super-Earths, because they had a deficit of close pairs and resonant pairs (see Fig. 15 in Hansen & Murray (2013)). Interestingly, the same deficit was found by Cossou et al. (2014) in a different model accounting for the migration of plan- etary embryos from several AUs away. On the other hand, a Article number, page 1 of 9 A&Aproofs:manuscript no. ogihara_etal_2015 model by Ogihara & Ida (2009) with gas drag and type I mi- gration led to more separated non-resonant pairs, when type I migration was reduced by about a factor of 100. In this paper, we revisit the process of in situ formation (Hansen & Murray 2012, 2013) with more complete simulations that start from a population of small planetary embryos (i.e., Mars-mass) and planetesimals, both carrying cumulatively 50% of the solid mass in the disk. These initial conditions are typ- ical of terrestrial-planet simulations (e.g., O'Brien et al. 2006). Thus, we do not assume that the protoplanets have reached the completion of the oligarchic growth process, but we sim- ulate that process from a much more primordial state. More- over, we consider the action of the gas that forces a migration and eccentricity/inclination damping of the embryos and plan- etesimals. Our aim is to clarify differences from the studies of Hansen & Murray (2012, 2013) and to understand, with im- proved and more realistic simulations, whether in situ formation can explain the observed properties of the close-in super-Earth systems. We performed simulations with different disk models (i.e., the amount of solids available), but we also present a model without migration and one without gas for comparison. In addition, we extend our model by including the accre- tion of primitive atmospheres onto super-Earths. The analyses of transit observations with radial velocity measurements or analy- ses of transit-time variations show that most of the planets larger than 2.5 Earth radii have very low densities (Marcy et al. 2014; Hadden & Lithwick 2014), so are thought to have thick H/He atmospheres (up to 10-20% by mass). We therefore consider the acquisition of H/He envelopes and discuss whether observed low-density super-Earths can be explained by our in situ accre- tion model. The rest of the paper is organized as follows. In Sect. 2, we describe our model and methods of N-body simulations; in Sect. 3 we give a series of the results of our simulations; in Sect. 4 we present results of simulations that include accretion of H/He atmosphere; in Sect. 5 the conclusions are provided. 2. Model and methods 2.1. Diskmodel We assume that the gas distribution is similar to the one of the classical MMSN model. Thus, for the gas surface density, we assume Σg = 2400(cid:18) r 1AU(cid:19)−3/2 exp − t 1Myr! g cm−2, (1) where Σg and r are the gas surface density and the radial distance from the central star, respectively. The value of Σg is 1.4 times the value in the MMSN. The gas dissipation is modeled as an exponential decay with the depletion timescale of 1 Myr. We set the disk inner edge at r = 0.1AU. The inner edge of the disk is expected to be at the radius where the orbital period equals the stellar rotation period. This could be a factor of two-three smaller than we assume. Our assumption of an inner edge at 0.1 AU is dictated by constraints due to computational time. Our results concerning the position of the final planets relative to the disk's inner edge can scale with the assumed edge's position. The tem- perature distribution is that for an optically-thin disk (Hayashi 1981), so that the disk scale height is h = 0.047(cid:18) r 1AU(cid:19)5/4 AU. Article number, page 2 of 9 (2) Table 1. List of models. In Model 2, the initial solid amount is decreased by a factor of two from Model 1. Model 4 corresponds to the model of Hansen & Murray (2012, 2013). Model Type I migration 1 2 3 4 yes yes no no e, i-damping yes yes yes no 2.2. Initialconditionsandnumericalmethod We chose an initial solid distribution similar to that of Hansen & Murray (2012); that is, 50M⊕ in total are placed be- tween 0.1 and 1AU. In our standard model, we set 250 embryos with a mass of M = 0.1M⊕ and 1250 planetesimals with a mass of M = 0.02M⊕ in such a way as to keep the radial distribution of the solid surface density proportional to r−3/2. Planetesimals gravitationally interact with the embryos but not with each other. This set-up is typical of successful simulations for the growth of the terrestrial planets in our solar system. By adopting the same set-up, we explore the effects of the enhanced solid distribution and much shorter orbital distances. There may be some caveats associated with the initial condi- tion. As shown in Sect. 3, once a disk of planetesimals and plane- tary embryos is set, the growth of planets in the close-in region is quite rapid. If the planetesimal formation process takes a longer time than the planet-growth timescale, it would be important to take planetesimal formation into account in simulations. How- ever, the formation of planetesimals is not yet fully understood. Therefore, in this study, we do not attempt to include the plan- etesimal formation; instead, we start simulations with already formed planetesimals. We note, however, that it is unlikely that planetesimal formation takes a timescale comparable to the disk lifetime (∼ Myr) for two reasons. First, this timescale would cor- respond to 30 million orbital periods, and it is difficult to under- stand why it should take so long. For instance, in the solar sys- tem chondritic planetesimals formed in about one million orbits. Second, the planetesimal formation process is most likely related to the drift of small particles through the disk (Johansen et al. 2014), which is due to gas drag, so very likely most of the mass was fed to the inner disk at an early time, when the gas density was higher. A late formation of the planetesimals would require that the small particles drift into the inner part of the disk only when the disk is disappearing. We think that this possibility is difficult to envision. Table 1 shows the list of simulations for each model. In Model 2, the total mass inside 1AU is decreased by a factor of two (the total mass is 25M⊕). In Model 3, type I migration is ne- glected, but eccentricity damping is still considered. In Model 4, the effect of gas is ignored the same as in the simulation by Hansen & Murray (2012, 2013). Our N-body code is based on SyMBA (Duncan et al. 1998), modified so that the effects of the gas disk are included accord- ing to the formulae reported in Sect. 2.3. When bodies collide with each other, the bodies are merged by assuming perfect ac- cretion. The physical radius of a body is determined by its mass, assuming an internal density of ρ = 3 g cm−3. The inner bound- ary of the simulation is set to r = 0.05AU. We use a 0.0004-year timestep for integrations. Masahiro Ogihara et al.: In situ formation of close-in super-Earths 2.3. Effectsofgas Eccentricities, inclinations, and semimajor axes are damped by disk interaction. Planets with more than roughly 0.1M⊕ suffer the tidal damping by the density wave, while planetesimals undergo aerodynamical gas drag. 2.3.1. Damping for embryos The eccentricity damping timescale for embryos, te, is given by (Tanaka & Ward 2004) where CD, rp, ρg, and ∆u are the gas drag coefficient, the physical radius, the density of gas disk, and the velocity of the body rela- tive to the disk of gas, respectively. Although our planetesimals have a mass of 0.02 Earth masses, we consider them as super- planetesimals, representing a swarm of much smaller objects cu- mulatively carrying the same total mass. From the size distribu- tion of the asteroid belt, we think that most planetesimals had a physical size of 50km in radius (Morbidelli et al. 2009). For the calculation of gas drag, this is the size we assume. For the value of CD, we use the same definition as in previous studies (e.g., Adachi et al. 1976; Brasser et al. 2007), which depends on the Mach number, the Knudsen number, and the Reynolds number. te = −1 1 0.78 M ≃ 3 × 102(cid:18) M∗!−1 Σgr2 vK!4 cs M∗ ! M⊕!−1 M∗ M⊙!−1/2 1AU(cid:19)2 M Ω−1 r yr, (3) 3.1. OutcomesofModels1and2 3. Results where M∗ is the stellar mass, L∗ the stellar luminosity, cs the sound speed, vK the Keplerian velocity, and Ω the orbital fre- quency, respectively. Here the relative motion between gas and planets is assumed to be subsonic (evK . cs). For planets with high eccentricities and inclinations, we include a correction fac- tor according to Eqs. (11) and (12) of Creswell & Nelson (2008). The migration timescale ta is given by (Tanaka et al. 2002; Paardekooper et al. 2011) ta = −1 1 M∗!−1 Σgr2 vK!2 β M cs M∗ ! M⊕!−1 M∗ 1AU(cid:19)3/2 M ≃ 2 × 105β−1(cid:18) Ω−1 r M⊙!1/2 yr, (4) where β is a coefficient that determines the direction and speed of type I migration. The type I migration torque depends on the Lindblad torque, the barotropic part of the horseshoe drag, the entropy-related part of the horseshoe drag, the barotropic part of the linear corotation torque, and the entropy-related part of the linear corotation torque. Paardekooper et al. (2011) derived the total type I migration torque, including both saturation and the cutoff at high viscosity. We write the migration coefficient β entering in Eq. (4) in the form (5) β = βL + βc,baro + βc,ent, where βL, βc,baro, and βc,ent are related to the Lindblad torque, the barotropic part of the corotation torque, and the entropy-related part of the corotation torque, respectively. Each formula is given by Eqs. (11)-(13) in Ogihara et al. (2015). In addition, the coro- tation torque decreases as the planet eccentricity increases (e.g., Bitsch & Kley 2010). We also consider this effect using the fol- lowing formula (Fendyke & Nelson 2014): e βC,baro(e) = βC,baro exp − βC,ent(e) = βC,ent exp − ef! , ef! , e (6) (7) where ef = 0.5h/r + 0.01. 2.3.2. Damping for planetesimals The aerodynamical gas drag force per unit mass is (Adachi et al. 1976) Faero = − 1 2M CDπr2 pρg∆u∆u, (8) The formation of close-in super-Earth systems can be schemati- cally divided into three phases: (i) growth of embryos from plan- etesmals, (ii) migration of embryos, and (iii) gas depletion and long-term orbital evolution. Through a series of simulations, we find that the first accretion stage is extremely rapid (∼0.1 Myr, much faster than the typical timescale of several 10 Myr ob- served in terrestrial planet simulations). Thus, the effect of the enhanced amount of solid is not just that of producing bigger planets: it also reduces the formation timescale. The closer prox- imity to the central star (i.e., shorter orbital periods) also favors a much faster accretion (see also Lee et al. 2014). Figure 1 shows snapshots of the evolution of one simulation for Model 1and indicates the gas surface density at each time (right axis). Fig- ure 2(a) shows the time evolution of the semimajor axis for this run. The color of lines indicates the eccentricity of the planets (see color bar). At t = 103yr, almost all planetesimals initially placed inside r ≃ 0.2AU have been accreted by embryos. The first accretion phase ends before t = 0.01Myr and 0.1Myr inside r ≃ 0.4AU and ≃ 0.7AU, respectively. Because of the short accretion timescale, it is thus not correct to neglect the gas effects as in Hansen and Murray's works. In fact, the protoplanets become massive well before the gas disk is substantially depleted. The gas forces the plan- ets to migrate inward. All planets would be lost into the star if there were no inner edge of the disk. With a sharp disk inner edge, the innermost planet is trapped at the edge by the planet-trap effect (Masset et al. 2006). The other plan- ets pile up in mutual mean motion resonances with the for- mer. The typical commensurabilities are between 5:4 and 6:5 with orbital separations of ≃ 5 − 10rH, thus the final orbits are packed near the disk's edge. The resonant configurations of two bodies that undergo convergent migration are deter- mined by the mass of the bodies and the relative migration speed (e.g., Mustill & Wyatt 2011; Ogihara & Kobayashi 2013). Ogihara & Kobayashi (2013) have derived the critical migration timescale for capture into first-order mean motion resonances. They found that if the relative migration timescale is shorter than ta,crit ≃ 1 × 105(M1/M⊕)−4/3TK, where M1 is the mass of larger body and TK the Keplerian period, bodies can only be captured in resonances closer than the 4:3 resonance (see Eq. (6) and Ta- ble 2 in Ogihara & Kobayashi 2013). In the result of Model 1, the mass of migrating embryos is ≃ 0.5M⊕ and the mass of plan- ets near the edge is ∼ 1M⊕, so the relative migration timescale is ≃ 105TK (Eq. (4)). Migrating embryos therefore settle into closely packed configurations. Five planets form at the end of the simulation presented in Figs. 1 and 2(a). In this run, the planets do not exhibit orbital instability after gas depletion because the Article number, page 3 of 9 A&Aproofs:manuscript no. ogihara_etal_2015 y t i c i r t n e c c E 10-1 103 yr 10-2 10-3 10-1 104 yr 10-2 10-3 10-1 105 yr 10-2 10-3 10-1 106 yr 10-2 10-3 10-1 107 yr 10-2 10-3 105 104 103 105 104 103 105 104 103 105 104 103 105 104 103 G a s S u r f a c e D e n s i t y ( g c m - 2 ) 0.05 0.1 0.3 1 Semimajor Axis (AU) Fig. 1. Snapshots of a system for Model 1. Filled circles represent bod- ies. The size of the circles is proportional to the radius of the body. The smallest circle represents a 0.02 Earth-mass body, while the largest one represents a 33 Earth-mass body. The solid line indicates the gas surface density (right axis). number of planets in the system is small and the planets are in mean motion resonances, leading to a long-lasting orbital stabil- ity (Chambers et al. 1996; Matsumoto et al. 2012). We performed ten simulation runs for each model with dif- ferent initial positions of embryos and planetesimals. The results are qualitatively the same: final orbits are compact near the disk's inner edge. In some runs, the system undergoes orbital instability during the third phase after 1 Myr, resulting in non-resonant and relatively separated orbits with a smaller number of planets (see Fig. 2(b) for example). Figure 3(a) shows the final orbital con- figurations of all ten runs, where the planets that formed through the same run are connected with a line. We observe steep mass gradients in this figure; the largest bodies are located near the edge, and the planetary mass monotonically decreases when the semimajor axis increases. This is because, in the presence of a strong migration torque, the resonant system is stable only if the innermost planet is the most massive, as already found by Morbidelli et al. (2008). If originally the innermost planet is not the most massive, an instability typically occurs. The first and the second planets have encounters with each other, and the sys- tem stabilizes in resonance only after that the two planets have exchanged their relative positions. The same is true for the sec- ond, relative to the third planet and so forth. Figure 3(b) shows orbital configurations of observed close- in super-Earth systems, in which the mass and semimajor axis of all planets are known. It is clear that the bulk architecture of the observed systems is inconsistent with the simulated plan- Article number, page 4 of 9 (a) 1 ) U A ( s i x A r o j a m m e S i 0.3 0.1 103 104 105 106 107 Time (yr) (b) 1 ) U A ( s i x A r o j a m m e S i 0.3 0.1 103 104 105 106 107 Time (yr) 1 10-1 10-2 E c c e n t r i c i t y 10-3 1 10-1 10-2 10-3 E c c e n t r i c i t y Fig. 2. Time evolution of planets for Model 1. The filled circles con- nected with solid lines represent the sizes of planets. The smallest circle represents a 0.2 Earth-mass embryo, while the largest ones represent a 33 Earth-mass planet in panel (a) and 35 Earth-mass planet in panel (b). The color of line indicates the eccentricity (color bar). etary systems. No steep mass gradients are observed1, and the bulk architecture of the observed systems cannot be reproduced through the simulations of model 1. If embryos migrate from beyond 1 AU, the mass gradient would be shallower and/or the outer boundary of the final planet distribution at around 0.2 AU would be removed. This would suggest that a "migration model" yields better results in reproducing the observed close-in super- Earth systems, which should be investigated by future N-body simulations. Figure 4(a) shows period ratios of adjacent pairs of observed close-in super-Earths indicating period ratios of first-order mean motion resonances (e.g., 2:1 and 3:2) and the cumulative dis- tribution of the period ratios (right axis). Most pairs have period ratios between 4:3 and 3:1, and there are a few pairs that have pe- riod ratios lower than 4:3. Figure 4(b) shows the period ratio dis- tribution for the ten runs for Model 1 and a copy of the observed 1 The mass gradient is steep in the Kepler-101 system, where the in- nermost planet has 51 Earth mass and the second planet is about 3 Earth mass. s r i a P f o r e b m u N 30 25 20 15 10 5 0 10 8 6 4 2 0 10 8 6 4 2 0 10 8 6 4 2 0 10 8 6 4 2 0 Masahiro Ogihara et al.: In situ formation of close-in super-Earths (a) 6:5 4:3 3:2 2:1 5:4 (b) (c) (d) (e) 1 2 3 5 7 10 Period Ratio 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 C u m u l a t i v e D i s t r i b u t i o n Fig. 4. Comparison of the distributions of period ratios of adjacent pairs of planets for (a) observation and (b)-(d) simulations. The distribution of period ratios is presented as a histogram (see the left y-axis) and as a cumulative distribution (see the right y-axis). Panels (b), (c), (d), and (e) show the results of Models 1, 2, 3, and 4, respectively. The dashed lines in panels (b)-(d) represent the cumulative distribution of observed planets shown in red in panel (a). The vertical lines indicate locations of mean motion resonances. distribution. Some pairs are in closely-spaced resonances (e.g., 5:4), and others have been knocked out of resonant orbits during the late instability. Although very closely spaced pairs (. 4:3) can form in Model 1, the cumulative distribution of observed close-in super-Earths is not matched by the results of Model 1. In fact, a Kolmogorov-Smirnov (K-S) test indicates that the cu- mulative distributions are statistically different (QKS ≪ 0.01). Figure 5 shows cumulative distributions of the eccentricities of the observed close-in super-Earths and of all the results pro- duced in the simulations corresponding to the same model. The general trend is that the eccentricity of the results of Model 1 is smaller than for close-in super-Earths. In Model 1, planets with e < 0.03 account for 66 percent of all bodies, while exoplanets with e < 0.03 make up only 26 percent of all planets. One reason for the small eccentricity in Model 1 is that the number of plan- ets in a system is small (N = 3 − 6), and the orbital stability time is long (Chambers et al. 1996), which inhibits giant impacts dur- ing gas dispersal (see Fig. 2(a) for example). In addition, even if planets undergo orbital instability as in Fig. 2(b), the eccentricity is not highly excited. This is because the effect of mutual scat- tering between planets is limited by the small number of planets. Eccentricity damping also operates during the gas dissipation phase due to remnant gas (see Fig. 2(b), for example). A K-S test indicates that the observed and simulated eccentricity distri- butions are statistically different (QKS ≪ 0.01). In summary, the results of Model 1 cannot reproduce the bulk properties (period ratio, eccentricity) of observed close-in super-Earth systems. Ec- centricities of exoplanets can be overestimated (see Sect. 3.2 for discussion). Thus, our results disagree with those of Hansen and Murray, which is not surprising given that the latter neglected the effect exerted by the disk of gas (particularly the inward mi- gration of proroplanets), not realizing that the super-Earths must form well within the gas-disk lifetime. We also performed ten runs for Model 2, where the initial solid amount is reduced by a factor of two from Model 1, in the hope of observing a slower accretion rate and consequently weaker migration effects. However, the results are qualitatively the same with those of Model 1. In Fig. 4(c), planetary pairs with their period ratio of < 4:3 account for 78 percent of all pairs, and the cumulative distribution is different from that of observed close-in super-Earths. The results are even more closely packed than those of Model 1 because planets are less vulnerable to or- bital instability during the third phase. The eccentricity distribu- tion in Fig. 5 also differs from that of close-in super-Earths. Article number, page 5 of 9 (a) 100 ) h t r a E M ( s s a M 10 1 0.1 0.01 (b) 100 ) h t r a E M ( s s a M 10 1 0.1 0.01 A&Aproofs:manuscript no. ogihara_etal_2015 0.1 1 Semimajor Axis (AU) 0.1 1 Semimajor Axis (AU) Fig. 6. Same as Fig. 2 but for a representative simulation of Model 3 (panel (a)) and Model (4) (panel (b)). Fig. 3. Results of 10 simulations of Model 1 (panel (a)). Observed close- in super-Earths systems (panel (b)). 3.2. OutcomesofModels3and4 observation e - σ model 1 model 2 model 3 model 4 n o i t u b i r t s i D e v i t a l u m u C 1 0.8 0.6 0.4 0.2 0 0.001 0.01 0.1 Eccentricity 1 Fig. 5. Comparison of cumulative eccentricity distributions between ob- served close-in super-Earths (thin solid line) and the planets formed through simulations (thick lines, see legend). The thin dotted line indi- cates the cumulative eccentricity distribution of observed super-Earths, in which each eccentricity is assumed to be e − σ. We then present results of Models 3 and 4, in which we sup- press the migration torques exerted by the gas-disk interaction onto the planets (Model 3) or we neglect the presence of gas altogether (Model 4). Clearly, both models are academic. In fact, a general mechanism that suppresses type I migration has never been found. Locally, type I migration can be halted or re- versed (Paardekooper & Mellema 2006; Bitsch et al. 2014), but no global weakening of type I migration has even been demon- strated. As for the absence of gas, this seems inconsistent with the fast growth timescale for the super-Earths. It is difficult to imagine that the gas disappears significantly faster than what we assume above (1 Myr). The reason we present these models is to highlight the role of migration and eccentricity/inclination damping in the results we presented before. Figures 6(a) and (b) show the typical orbital evolution for Models 3 and 4, respectively. In both cases, planetary systems are not as compact when compared with the results of Models 1 and 2. In the simulations for Model 3, in which type I migra- tion is neglected, planets undergo slow inward migration due to eccentricity damping (see also Sect. 5.2 in Ogihara et al. 2014), Article number, page 6 of 9 Masahiro Ogihara et al.: In situ formation of close-in super-Earths and planets are temporarily captured in mutual mean motion res- onances before a few Myr. Then, they undergo orbital instabil- ity and collide with each other, resulting in non-resonant orbital configurations, which are qualitatively similar to those obtained in the slow-migration simulation of Ogihara & Ida (2009). In the results for Model 4, planets are never in resonances, which is almost the same as in the simulations of Cossou et al. (2014) and Hansen & Murray (2012, 2013). The averaged mass of the largest bodies is ≃ 13 M⊕ (Model 3) and ≃ 14 M⊕ (Model 4), which are lower than for Model 1. Figures 4(d) and (e) show the period ratio distribution of Models 3 and 4, respectively. Interestingly, the results match the observations much better than those of Models 1 and 2. In the results of Model 3, most pairs have period ratios between 4:3 and 3:1, while some pairs are relatively separated (> 3:1). In the results of Model 4, 85 percent of pairs lies between 2:1 and 3:1. In comparison with the observed distribution, Model 3 is a good match to the distribution of exoplanets. A K-S test shows that the distributions are similar with a significance level of QKS = 0.24 for Model 3, while the distribution of Model 4 is not similar to the observed distribution (QKS ≪ 0.01). In Fig. 5, the eccentricity of the planets produced in Model 3 is generally smaller than in Model 4 because of the eccentric- ity damping. Compared with the distribution of exoplanets, the K-S test shows that the distributions are different for Model 3 (QKS ≪ 0.01), while the distributions for Model 4 are closer, but are still not satisfactory (QKS = 0.026). Thus, neither Model 3 nor Model 4 explains the observation, because the first has prob- lems with the eccentricity distribution, the second with the or- bital period distribution. However, it is fair to say that the measurement of the eccen- tricity of exoplanets is still difficult, and the uncertainty on the results is quite large. In particular, it has been shown that ec- centricities, which are derived from radial velocity surveys, can be overestimated (e.g., Shen & Turner 2008; Zakamska et al. 2011). Therefore, the observed eccentricity distribution in Fig. 5 may shift to lower values. In this case the results of Model 3 might match the eccentricity distribution as well. As an example, we recalculated the eccentricity distribution in a simple way. The eccentricity of each exoplanet is set to e−σ, where σ is the estimated error. The new distribution is indicated in Fig. 5. We find that the new distribution gives a better match to Model 3 (QKS = 0.30) rather than Model 4 (QKS = 0.010). 4. Accretion of primitive atmospheres An objection often expressed against the in situ accretion model for super-Earths is that observations indicate in many cases (e.g., Kepler-11, see Lissauer et al. 2011a) that these planets have a low bulk density. One would expect planets grown in the inner part of the disk to be rocky, given that the high local disk tem- perature should not have allowed ice condensation. A possibility, however, is that super-Earths accreted prim- itive H/He atmospheres, leading to low bulk densities. Recent simulations of the structure and evolution of planetary atmo- spheres have demonstrated that super-Earths can indeed accrete primordial atmospheres from the protoplanetary disk provided that they do not accrete solids at a high rate (Lee et al. 2014). As we have seen above, we expect that in situ formation is ex- tremely rapid and reaches completion well within the lifetime of the gas disk. Thus, we expect our planets to be in the condi- tion of a very low accretion rate of solids when there is still gas in the disk, enabling the acquisition of significant atmospheres. In what follows we implement the most recent recipes for gas accretion to compute the mass of atmosphere expected for our super-Earths. 4.1. Model Once planetary embryos embedded in a gas disk are sufficiently massive (typically the mass of Mars or more), they capture part of the disk gas to have an atmosphere of their own (e.g., Wuchterl et al. 2000). This atmosphere or envelope grows with the mass of the embryo itself until it becomes greater than the mass of the solid core. At this so-called crossover mass (Menv ∼ Mcore), accretion enters a runaway phase with the envelope mass increasing exponentially and the planet becoming a giant planet (Pollack et al. 1996). In this process, the accretion of solids has two effects: it increases the mass of the planet and heats the enve- lope. The first favors the growth of the envelope, but the second leads to an increase in the crossover mass because of a more ten- uous envelope. Most works on the growth of giant planet cores have thus focused on obtaining expressions that depend on the accretion rate of planetesimals (e.g., Ikoma et al. 2001). However, since in our simulations the accretion of solids stalls while the gas disk is still present, at this later stage, the rate of cooling of the envelope becomes crucial for controlling the growth of the envelope (Pollack et al. 1996). We choose to em- pirically model the growth of the envelope by fitting the results of Ikoma & Hori (2012), Piso & Youdin (2014), and Lee et al. (2014), which are works that account for the planetary cooling to calculate the resulting envelope growth. Specifically, we model the envelope mass, Menv, as ek1(t/trun−1), (9) Menv Mcore = k2 1 + k3  1 + k3 t trun!1/3 is the core mass and trun the time to get to where Mcore the crossover mass. As for the coefficients, we use k1 = (Mcore/15M⊕)−1, k2 = 1 + (Mcore/40M⊕), and k3 = 9. Most of the uncertainty is linked to the value of trun, which depends crit- ically on the opacities chosen, the cooling rate of the core, and orbital distance. For the present simulations we choose to use as a fiducial value trun = 107(Mcore/5M⊕)−3 yr, (10) which lies in between the results obtained by Ikoma & Hori (2012) and Piso & Youdin (2014) and Lee et al. (2014). Then the accretion rate onto the planet is expressed by Menv = Menv" k3 3 1 t1/3 run t2/3 + k3t + k1 trun# . (11) Equation (9) implicitly assumes that the disk can supply all the gas that the planet is able to accrete. In reality, this amount is limited by the viscous inflow in the disk at the location of the planet (e.g., Tanigawa & Ikoma 2007). The accretion rate due to viscous diffusion is Mvis ≃ 3πνΣg, (12) where an "alpha model" for the disk viscosity is used with ν = αcsh (cs is the isothermal sound speed), and we adopt α = 10−3. The actual accretion rate is limited by the minimum of Eqs. (11) and (12). Based on these prescriptions, we recalculate our N-body sim- ulations for Model 1, this time accounting for the accretion of an atmosphere after the end of core accretion phase (t . 105yr). Article number, page 7 of 9 A&Aproofs:manuscript no. ogihara_etal_2015 In Fig. 7(a), the planet at a = 0.1AU with Mcore = 9.2M⊕ accretes gas from the disk and ends up retaining a thick atmo- sphere of Menv = 6.2M⊕ and Menv/Mtot = 0.30 at t = 10Myr. This planet migrates inside of the disk inner edge at t ≃ 2.6Myr by the interaction with four outer bodies. This migration prevents the former planet from accreting more gas. The innermost planet is also moved inside the disk inner edge before t = 0.1Myr and stops accreting gas at that point. In the simulation with a shorter value of trun shown in Fig. 7(b), a planet with Menv = 14M⊕ and Menv/Mtot = 0.56 eventually forms and also moves inside of the disk edge at t ≃ 0.3Myr. It is interesting to notice that as the envelope mass increases, systems become vulnerable to orbital instability. In panel (b), in fact, the system undergoes orbital instability earlier than in panel (a), and the number of final planets is also smaller. As for the first question posed in the beginning of this sec- tion, we find that in both simulations, one planet can acquire a thick H/He atmosphere from the disk, which may explain the ori- gin of the observed low density super-Earths. Regarding the sec- ond question, we also observe that the acquisition of a massive atmosphere by one planet destabilizes closely spaced systems, leading to relatively well separated systems with fewer planets or even single-planet systems. The properties (e.g., orbital separation and bulk density) of the final systems shown in Fig. 7(a) are reminiscent of some known planetary system. For example, two super-Earths were discovered in the Kepler-36 system, where the inner planet would be a rocky planet without a thick atmosphere, and the outer one would possess a thick atmosphere. However, the typi- cal properties of the observed close-in super-Earths are unlikely to be reproduced. This is because the systems are tightly packed near the edge before the acquisition of atmospheres (see t = 105 yr in Fig. 7), so that systems separated more than 2:1 resonances, which have been observed in super-Earth systems (see Fig. 4(a)), can hardly be reproduced by simulations for Model 1. Lee et al. (2014) point out that super-Earths with a mass of 10M⊕ tend to undergo runaway gas accretion, thus becoming gas giant planets. We do not observe this phenomenon in the results of Figs. 7(a) and (b), even though the core mass is high. This is because the massive planets move inside the disk inner edge, where there is no gas, and cease envelope accretion. 5. Discussion and conclusions We have re-examined in situ formation of close-in super-Earths with improved simulations, in which the effects of the disk of gas are considered. The simulations are started with small em- bryos and planetesimals. We find that the accretion of planets is extremely rapid owing to the large amount of solid material and short orbital periods. Thus, it is not correct to ignore the ef- fects of the gas disk for the investigation of in situ formation of close-in super-Earths. We performed ten runs of simulation for our fiducial model and found the following. 1) The orbital architecture of resultant systems is very compact near the disk inner edge. 2) The eccen- tricity of super-Earths is small because planets can be stable after gas depletion. If they undergo orbital instability, the eccentricity is not highly excited. 3) The masses of planets monotonically decrease when increasing the semimajor axis. These character- istics are not consistent with observed close-in super-Earths. In fact, the cumulative observed distributions of period ratios of ad- jacent pairs and of eccentricities are statistically different from those we produce. Fig. 7. Evolution of the semimajor axis and the envelope mass (color bar). Panel (a) shows the result of trun = 107(Mcore/5M⊕)−3 yr, while panel (b) indicates that of trun = 106(Mcore/5M⊕)−3 yr. The filled circles connected with solid lines represent the sizes of planets. The largest circles represent a 33 Earth-mass planet in panel (a) and 36 Earth-mass planet in panel (b). 4.2. Results We now present the results of these simulations with the set-up of Model 1 but accounting for atmosphere accretion. There are two crucial questions that we wish to address: 1) Can atmosphere accretion be substantial enough to significantly reduce the appar- ent bulk density of the planet and mimic low-density objects as in the Kepler-11 system? 2) Can atmosphere accretion change the masses of the planets enough to induce a late dynamical in- stability that can result in less compact systems. Figure 7 shows the orbital evolution of each system in which the color of each planet's lines indicates the ratio of enve- lope mass to total mass, Menv/Mtot (see color bar). We adopt trun = 107(Mcore/5M⊕)−3 yr as a fiducial value for trun in the result of panel (a). In addition, we also perform a simulation under more efficient conditions for envelope accretion trun = 106(Mcore/5M⊕)−3 yr, the results of which are shown in panel (b). Article number, page 8 of 9 Masahiro Ogihara et al.: In situ formation of close-in super-Earths Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J.J., Podolak, M., & Greenzweig, Y. 1996, Icarus, 124, 62 Schneider, J., Dedieu, C., Le Sidaner, P., Savalle, R., & Zolotukhin, I. 2011, A&A, 532, A79 Shen, Y., & Turner, E. L. 2008, ApJ, 685, 553 Suzuki, T. K., & Inutsuka, S. 2009, ApJ, 691, L49 Suzuki, T. K., Muto, T., & Inutsuka, S. 2010, ApJ, 718, 1289 Tanaka, H., Takeuchi, T., & Ward, W. R. 2002, ApJ, 565, 1257 Tanaka, H., & Ward, W. R. 2004, ApJ, 602, 388 Wright, J. T., Fakhouri, O., Marcy, G. W., et al. 2011, PASP, 123, 412 Tanigawa, T., & Ikoma, M. 2007, ApJ, 667, 557 Ward, W. R. 1986, Icarus, 67, 164 Weidenschilling, S. J. 1977, Ap&SS, 51, 153 Wuchterl, G., Guillot, T., & Lissauer, J. J. 2000, Protostars and Planets IV, 1081 Zakamska, N. L., Pan, M., & Ford, E. B. 2011, MNRAS, 410, 1895 We have also investigated orbital evolution including the ac- cretion process of primitive atmospheres onto the super-Earths. The results show that close-in super-Earths that formed in situ can acquire a thick H/He atmosphere in which the planets stop envelope accretion when they migrate inside the disk's inner edge. Interestingly, if type I migration is neglected (but the ec- centricity damping is included), the results match the observa- tions much better. However, no mechanism capable of suppress- ing type I migration over the whole inner disk has ever been found. Recent studies have shown that MRI-driven disk winds, in which gas material is blown away from the surface of the disk, can alter the density profile of the gas disk, potentially slow- ing down or even reversing the migration of the protoplanets (e.g., Suzuki & Inutsuka 2009; Suzuki et al. 2010; Ogihara et al. 2015). This possibility, however, requires further investigation. Unless a mechanism for a global reduction of type I migration is demonstrated, our results imply that in situ accretion of close-in super Earth is unlikely. Acknowledgements. We thank John Chambers for comments that helped us im- prove the manuscript. We also thank Yasunori Hori and Hiroki Harakawa for helpful comments. We thank the CRIMSON team, who manages the mesocen- tre of the OCA, on which most simulations were performed. Numerical com- putations were in part conducted on the general-purpose PC farm at CfCA of NAOJ. M.O. is supported by the JSPS Postdoctoral Fellowships for Research Abroad. A.M. and T.G. were supported by ANR, project number ANR-13 -- 13- BS05-0003-01 projet MOJO(Modeling the Origin of JOvian planets). References Adachi, I., Hayashi, C., & Nakazawa, K. 1976, Prog. Theor. Phys., 56, 1756 Allégre, C. J., Manhés, G. & Göpel, C. 2008, Earth Planet. Sci. Lett., 267, 386 Bitsch, B., & Kley, W. 2010, A&A, 523, A30 Bitsch, B., Morbidelli, A., Lega, E., Kretke, K., & Crida, A. 2014, A&A, 570, A75 Chambers, J. E., Wetherill, Q. W., & Boss, A. P. 1996, Icarus, 119, 261 Chatterjee, S., & Tan, J. 2014, ApJ, 780, 53 Chatterjee, S., & Tan, J. 2015, ApJ, 798, L32 Cossou, C., Raymond, S. N., Hersant, F., & Pierens, A. 2014 A&A, 569, A56 Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., et al. 2014, ApJ, 790, 146 Fendyke, S. M., & Nelson, R. P. 2014, MNRAS, 437, 96 Hadden, A., & Lithwick, Y. 2014, ApJ, 787, 80 Hansen, B. M., & Murray, N. 2012, ApJ, 751, 158 Hansen, B. M., & Murray, N. 2013, ApJ, 775, 53 Hayashi, C. 1981, Prog. Theor. Phys. Suppl., 70, 35 Inamdar, N. K., & Schlichteng, H. E. 2015, MNRAS, submitted Ikoma, M., & Hori, Y. 2012, ApJ, 753, 66 Ikoma, M., Emori, H., & Nakazawa, K. 2001, ApJ, 553, 999 Johansen, A., Blum, J., Tanaka, H., Ormel, C., Bizzarro, M., & Rickman, H. 2014, in Beuther H., Klessen R., Dullemond C., Henning Th., eds, Protostars and Planets VI. Univ. Arizona Press, Tucson Kokubo, E., & Ida, S. 1998, Icarus, 131, 171 Lee, E. J., Chiang, E., & Ormel, C. W. 2014, ApJ, 797, 95 Lissauer, J. J. et al. 2011, Nature, 470, 53 Lissauer, J. J. et al. 2011, ApJS, 197, 8 Marcy, G. W., Weiss, L. M., Petigura, E. A., Isaacson, H., Howard, A. W., & Buchhave, L. A. 2014, PNAS, 111, 12655 Masset, F. S., D'Angelo, G., & Kley, W. 2006, ApJ, 652, 730 Matsumoto, Y., Nagasawa, M., & Ida, S. 2012, Icarus, 221, 624 Mayor, M., et al. 2009, A&A, 493, 636 Morbidelli, A., Crida, A., Masset, F., & Nelson, R. P. 2008, A&A, 478, 929 Morbidelli, A., Bottke, W. F., Nesvorný, D., & Levison, H. F. 2009, Icarus, 204, 558 Moorhead, A. V., Ford, E. B., Morehead, R. C., et al. 2011, ApJS, 197, 1 Mustill, A. J., & Wyatt, M. C. 2011, MNRAS, 413, 554 O'Brien, D. P., Morbidelli, A., & Levison, H. F. 2006, Icarus, 184, 39 Ogihara, M., & Ida, S. 2009, ApJ, 699, 824 Ogihara, M., & Kobayashi, H. 2013, ApJ, 775, L34 Ogihara, M., Kobayashi, H., & Inutsuka, S. 2014, ApJ, 787, 172 Ogihara, M., Kobayashi, H., Inutsuka, S., & Suzuki, T. K. 2015, A&A, submitted Paardekooper, S. -J., & Mellema, G. 2006, A&A, 459, L17 Paardekooper, S. -J., Baruteau, C., & Kley, W. 2011, MNRAS, 410, 293 Piso, A.-M., & Youdin, A. N. 2014, ApJ, 786, 21 Article number, page 9 of 9
1307.6608
1
1307
2013-07-24T23:33:59
Comparing Dawn, Hubble Space Telescope, and Ground-Based Interpretations of (4) Vesta
[ "astro-ph.EP", "astro-ph.IM", "physics.geo-ph", "physics.ins-det", "physics.space-ph" ]
Observations of asteroid 4 Vesta by NASA's Dawn spacecraft are interesting because its surface has the largest range of albedo, color and composition of any other asteroid visited by spacecraft to date. These hemispherical and rotational variations in surface brightness and composition have been attributed to impact processes since Vesta's formation. Prior to Dawn's arrival at Vesta, its surface properties were the focus of intense telescopic investigations for nearly a hundred years. Ground-based photometric and spectroscopic observations first revealed these variations followed later by those using Hubble Space Telescope. Here we compare interpretations of Vesta's rotation period, pole, albedo, topographic, color, and compositional properties from ground-based telescopes and HST with those from Dawn. Rotational spectral variations observed from ground-based studies are also consistent with those observed by Dawn. While the interpretation of some of these features was tenuous from past data, the interpretations were reasonable given the limitations set by spatial resolution and our knowledge of Vesta and HED meteorites at that time. Our analysis shows that ground-based and HST observations are critical for our understanding of small bodies and provide valuable support for ongoing and future spacecraft missions.
astro-ph.EP
astro-ph
"! #! $! %! &! '! (! )! *! "+! ""! "#! "$! "%! "&! "'! "(! ")! "*! #+! #"! ##! #$! #%! #&! #'! #(! #)! #*! $+! $"! $#! $$! $%! $&! $'! $(! $)! $*! %+! %"! %#! %$! %%! %&! %'! %(! %)! %*! &+! Comparing Dawn, Hubble Space Telescope, and Ground-Based Interpretations of (4) Vesta Vishnu Reddy Max Planck Institute for Solar System Research, Katlenburg-Lindau, Germany Department of Space Studies, University of North Dakota, Grand Forks, USA Planetary Science Institute, Tucson, AZ 85719, USA Email: [email protected] Jian-Yang Li Planetary Science Institute, Tucson, AZ 85719, USA Lucille Le Corre Planetary Science Institute, Tucson, AZ 85719, USA Max Planck Institute for Solar System Research, Katlenburg-Lindau, Germany Jennifer E. C. Scully Institute of Geophysics and Planetary Physics, University of California Los Angeles, Los Angeles California, USA Robert Gaskell Planetary Science Institute, Tucson, Arizona, USA Christopher T. Russell Institute of Geophysics and Planetary Physics, University of California Los Angeles, Los Angeles California, USA Ryan S. Park Jet Propulsion Laboratory, California Institute of Technology, Pasadena, Califronia, USA Andreas Nathues Max Planck Institute for Solar System Research, Katlenburg-Lindau, Germany Carol Raymond Jet Propulsion Laboratory, California Institute of Technology, Pasadena, Califronia, USA Michael J. Gaffey Department of Space Studies, University of North Dakota, Grand Forks, USA Holger Sierks Max Planck Institute for Solar System Research, Katlenburg-Lindau, Germany Kris J. Becker Astrogeology Science Center, U.S. Geological Survey, Flagstaff, Arizona, USA Lucy A. McFadden NASA Goddard Spaceflight Center, Greenbelt, Maryland, USA Pages: 51 ! "! &"! &#! Figures: 9 &$! Tables: 5 &%! &&! Editorial correspondence to: Proposed Running Head: Dawn, HST, Ground-based Studies of Vesta &'! Vishnu Reddy &(! &)! Haleiwa &*! Hawaii 96712 59-495 Hoalike Road '+! '"! '#! (808) 342-8932 (voice) '$! [email protected] '%! '&! ''! '(! ')! '*! (+! ("! (#! ($! (%! (&! ('! ((! ()! (*! ! #! formation. Prior to Dawn’s arrival at Vesta, its surface properties were the focus of rotation period, pole, albedo, topographic, color, and compositional properties from visited by spacecraft to date. These hemispherical and rotational variations in surface brightness and composition have been attributed to impact processes since Vesta’s intense telescopic investigations for nearly a hundred years. Ground-based photometric and spectroscopic observations first revealed these variations followed later by those using Hubble Space Telescope (HST). Here we compare interpretations of Vesta’s ground-based telescopes and HST with those from Dawn. Our goal is to provide ground truth for prior interpretations and help identify the limits of ground-based studies of )+! Observations of asteroid (4) Vesta by NASA’s Dawn spacecraft are interesting because Abstract )"! )#! its surface has the largest range of albedo, color and composition of any other asteroid )$! )%! )&! )'! )(! ))! )*! *+! *"! *#! *$! *%! *&! *'! *(! *)! **! "++! "+"! "+#! spatial resolution of the former. We also present HST and Dawn albedo and color maps ! $! asteroids in general. The improved rotational period measurement from Dawn is determined by Dawn is 309.03º±0.01º, 42.23º ±0.01º and is within the uncertainties of pole orientation determined by Earth-based measurements (Li et al., 2011: 305.8º±3.1º, 41.4º±1.5º). Similarly, the obliquity of Vesta is 27.46º based on the pole measurement 0.222588652 day (Russell et al., 2012), and is consistent with the best ground-based rotation period of 0.22258874 day (Drummond et al., 1998). The pole position for Vesta from Dawn and all previous pole measurements put the obliquity within 3º of this value. Topography range from Dawn shape model is between -22.45 to +19.48 km relative to a 285 km x 285 km x 229 km ellipsoid. The HST range is slightly smaller (-12 km to +12 km relative to a 289 km x 280 km x 229 km ellipsoid) than Dawn, likely due to lower provide valuable support for ongoing and future spacecraft missions. knowledge of Vesta and HED meteorites at that time. Our analysis shows that ground- based and HST observations are critical for our understanding of small bodies and "+$! maps serve to orient observers and identify compositional and albedo features from prior of Vesta in the Claudia (used by the Dawn team) and IAU coordinate systems. These "+%! "+&! studies. We have linked several albedo features identified on HST maps to morphological "+'! features on Vesta using Dawn Framing Camera data. Rotational spectral variations "+(! While the interpretation of some of these features was tenuous from past data, the observed from ground-based studies are also consistent with those observed by Dawn. "+)! "+*! interpretations were reasonable given the limitations set by spatial resolution and our ""+! """! ""#! ""$! ""%! ""&! ""'! ""(! "")! ""*! "#+! "#"! "##! "#$! "#%! "#&! ! %! hemispherical scale albedo variations dominate the lightcurve rather than shape "#'! "#(! Vesta is one of the most frequently observed objects in the Main Asteroid Belt since its 1. Introduction "#)! "#*! discovery by Heinrich Olbers in 1807. Ground-based color observations of Vesta as early "$+! as 1929 (Bobrovnikoff, 1929) revealed surface albedo/color variations that were "$"! attributed to composition. Taylor (1973) noted that Vesta’s lightcurve changed depending "$#! maximum. Degewij (1978) used polarimetry to verify that Vesta’s lightcurve is on viewing geometry with northern hemisphere and equatorial views showing a single "$$! "$%! with the visual wavelength lightcurve suggested that the observed light curve was indeed dominated by albedo. The relationship between albedo and polarization and the overlap "$&! "$'! that controlled by albedo variation. Later photometric observations confirmed "$(! "$)! "$*! "%+! (McCord et al. 1970) revealed a deep absorption band at 0.9 µm attributed to the mineral "%"! pyroxene. Overall spectral shape, and the presence of this pyroxene band, suggested a "%#! meteorites and suggested Vesta as a differentiated object. Rotationally resolved near-IR compositional link between Vesta and the howardites-eucrites-diogenites (HED) "%$! "%%! spectra of Vesta from NASA IRTF suggested that albedo variations might be linked to "%&! "%'! "%(! "%)! surface compositional heterogeneity (e.g., Gaffey, 1997; Vernazza et al. 2005; Reddy et Disk-integrated visible wavelength (0.3 to 1.1 µm) spectral observations of Vesta (Drummond et al. 1988). al. 2010). Subsequent Hubble Space Telescope (HST) observations confirmed the affinity between compositional variations and albedo units thought to be surface morphological features (Thomas et al. 1997a; Binzel et al. 1997; Li et al. 2010). ! &! (Sierks et al. 2011) that imaged the surface of Vesta with an angular resolution of 93 us understand Vesta prior to arrival of Dawn, but also enabled us to verify the validity of these studies. With over 600,000 asteroids discovered so far, and the ever-increasing cost "%*! The Dawn spacecraft entered orbit around Vesta in July 2011 to begin its "&+! mapped the surface of Vesta using its three instruments: Framing Cameras (FC), Visible yearlong mapping mission (Russell et al. 2012). During this period, the spacecraft "&"! "&#! GRaND). The Dawn Framing Cameras (FC) are a pair of identical 1024x1024 pixel and Infrared Mapping Spectrometer (VIR), and Gamma Ray and Neutron Detector "&$! "&%! imagers equipped with seven color filters (0.44-0.98 µm) and one panchromatic filter "&&! "&'! !rad/pixel. Only FC2 was used during Vesta mapping phase with FC1 being a backup. "&(! Table 1 shows the list of filters along with their central wavelength and band pass. The "&)! During approach, three rotational characterization (RC) phases imaged the entire visible FC mapped the surface at varying spatial resolution depending on the orbital phase. "&*! "'+! RC3B) (Table 2). surface at 9.07 km/pixel (RC1), 3.38 km/pixel (RC2), and ~487 meters/pixel (RC3, "'"! "'#! Dawn is the first mission to an asteroid that has been profusely studied by ground- "'$! "'%! "'&! "''! of robotic exploration of small bodies, sending a spacecraft to many of these asteroids is "'(! inconceivable. Dawn presents a unique opportunity to verify and validate ground-based "')! Vesta with those from Dawn FC to verify interpretations of albedo units and rotational observations of Vesta. In this work we aim to compare ground-based and HST data of "'*! "(+! variations. We also provide maps in three coordinate systems: the original Olbers system "("! based on Thomas et al. (1997a); the Claudia system based on Russell et al. (2012) and ! '! based telescopes and HST over many decades. This wealth of knowledge not only helped used by the Dawn science team; and the IAU coordinate system in which all the Dawn data will be archived on the Planetary Data System (PDS). 2. Data Sets and Processing Three data sets were used in this paper: ground-based spectral data from Gaffey (1997) and Reddy et al. (2010); HST data from Thomas et al. (1997a), Binzel et al. (1997), and Li et al. (2010); and Dawn FC data from RC1, RC2 and RC3. The best resolution of HST "(#! "($! "(%! "(&! "('! "((! "()! data (38 km/pixel) is compared to the lowest resolution from Dawn during RC1 (9 "(*! km/pixel). Data reduction procedures for ground-based data are described in Gaffey ")+! Binzel et al., (1997) and Li et al. (2010). A detailed description of Dawn FC data (1997), Reddy et al. (2010); and for HST data processing in Thomas et al. (1997a), ")"! ")#! processing pipeline is presented in the supplementary materials section of Reddy et al. ")$! ")%! ")&! ")'! ")(! "))! ")*! "*+! "*"! "*#! "*$! "*%! interpretive scheme consisting of a single band rainbow color-coded map that uses the ! (! created by using the 0.75-!m filter data which shows greatest albedo contrast among the FC filters and also has least amount of infield stray light residuals after correction (Reddy et al. 2012a). A global 0.75-!m-filter mosaic was extracted from the seven-color mosaic but this is below the wavelength range of Dawn FC filters. The Dawn albedo maps were diogenite (ED) ratio map. The band depth map (0.75/0.92 !m) is a single band color- coded map that helps to quantify the 0.9-!m-pyroxene band depth. The ED ratio is an (2012a). Here we describe processing after the creation of photometrically and spectrally calibrated seven color global mosaics. HST observations of Vesta at shorter wavelengths show higher contrast in albedo using IDL ENVI. In addition to the albedo map, we created a band depth map, and a eucrite- "*&! ratio of 0.98 !m and 0.92 !m filters. Diogenite-rich areas are in red and show a deeper "*'! 0.90-!m pyroxene band whereas eucrite-rich areas are in blue (Reddy et al. 2012a). Due "*(! wavelength and so the eucrite ED ratio is ~1. In contrast diogenites have ED ratio >1 due to higher iron content in eucrites their 0.9-!m pyroxene band is shifted to longer "*)! "**! to their lower iron abundance (Reddy et al. 2012a). #++! 3. Evolution of Coordinate Systems #+"! The evolution of coordinate systems used on Vesta has historically depended on the #+#! Russell et al. 2012). Typically, ground-based rotational spectral studies of asteroids (e.g., spatial resolution of data available at that time (Gaffey, 1997; Thomas et al. 1997b; #+$! Reddy et al. 2010) used a lightcurve-based coordinate system where near-simultaneous #+%! #+&! minima of the lightcurve becoming the prime meridian. Gaffey (1997) observed several lightcurve observations are used to phase spectral observations (arbitrarily) with the #+'! #+(! distinct compositional units on the surface of Vesta based on rotational spectral #+)! meridian on a compositional unit informally called “Leslie Formation,” which was observations. Creating a simple compositional map, Gaffey (1997) centered his prime #+*! #"+! interpreted as olivine-rich. This corresponds approximately to a weak inflection in #""! Gaffey 1997). It is important to note that the original maps published in Gaffey (1997) brightness around the lightcurve maxima of Vesta (0.75 rotational phase in Fig. 3 of #"#! #"$! had an error in latitude range. The correct latitude range of Gaffey (1997) maps is 90°N #"%! #"&! #"'! #"(! (2010). Vesta is in simple rotation with slow precession (Asmar, pers. comm). For this ! )! to 60°S. The general conventions for defining a body fixed coordinate system for Solar System small bodies as adopted by the IAU/WGCCRE are summarized in Archinal et al. case, its body fixed coordinate system follows the right-hand rule that defines a “positive pole” by its direction of angular momentum, and the longitude increases towards the direction of rotation. The positive pole of Vesta is above (or on the north of) the to position the prime meridian at the center of the map, and longitude running between 0 of rotation is referred to as “East”. All coordinate systems defined for Vesta in this work invariable plane of the Solar System (prograde rotation), similar to many planets, further simplifying the case without causing any confusion. Using the geodesy conventions, the follow these conventions. In addition, when displaying maps, we adopted the convention positive pole of Vesta is thereby always referred to as the “North pole”, and the direction #")! #"*! ##+! ##"! ###! ##$! ##%! ##&! ##'! ##(! and 360 (positive values). We adopted one single coordinate system and one single ##)! convention for all the maps and figures used here to avoid further confusion. ##*! were produced before the current conventions were adopted, and used a longitude system Historically, Gaffey (1997), Thomas et al. (1997a), and Binzel et al. (1997) maps #$+! #$"! that increases towards the opposite direction of rotation, i.e., towards west, or a mixed #$#! #$$! #$%! compatible orientation with the IAU conventions, i.e., 0-longitude center with north up. #$&! In addition, the map generated by Gaffey (1997) used incorrect range of latitude, while #$'! maps generated by Li et al. (2010) fully comply with the IAU conventions. his north-south and east-west directions were compatible with the IAU conventions. The #$(! #$)! Thomas et al. (1997b) proposed a coordinate system for Vesta based on the #$*! most prominent visible feature.” This prominent feature was a 200 km wide low albedo rotational axis they derived from HST images and a new prime meridian centered on “the #%+! ! *! east and west longitude to keep longitude values positive between 0º and 180º in the case of Thomas et al. But except for the longitude system, all those maps were displayed in a day derived by Drummond et al. (1988). All subsequent publications (e.g., Binzel et al., al., 2010, 2012b) till the arrival of Dawn at Vesta used the Thomas et al. (1997b) Olbers asteroid’s surface. The system is defined by Claudia crater, which is a ~625 meter coordinate system with the Thomas et al. (1997b) pole measurement. Table 3 shows the values in pre-Dawn coordinate systems and in the Dawn coordinate system for the right ascension ("0) and declination (#0) of the spin pole of Vesta and for the ephemeris position of the prime meridians. Binzel et al. (1997) updated the Vesta lithologic map from Gaffey (1997) into the Olbers coordinate system, and found that the prime meridian 1997; Vernazza et al. 2005; Zellner et al. 2005; Carry et al. 2010; Li et al. 2010; Reddy et #%"! unit, which was the largest and most distinct visible feature on Vesta, and was informally #%#! Olbers. To calculate the location of the “Olbers” prime meridian for subsequent called “Olbers Regio” (Zellner et al. 1997) in honor of Vesta’s discoverer Wilhelm #%$! #%%! observations they developed a simple equation that used a rotation period of 0.2225887 #%&! #%'! #%(! #%)! #%*! #&+! #&"! #&#! #&$! #&%! #&&! #&'! #&(! #&)! #&*! #'+! #'"! #'#! #'$! framework for representing and navigating the surface of Vesta, which is in accordance ! "+! coordinate system that was founded upon the new spacecraft-derived knowledge of the diameter crater located in Vesta’s Oppia quadrangle (Russell et al. 2012). The location of the crater center is 1.6°S, 356.0°E in Claudia system coordinates. Vesta’s prime meridian is 4° east of Claudia’s position. Longitude increases to the east of the prime meridian. The Dawn team has undertaken all mapping and science investigations in the Claudia in Gaffey’s maps was at ~255° east of Olbers. Early in Dawn’s encounter with Vesta the Dawn science team derived a coordinate system. The purpose of the Claudia coordinate system is to provide a convenient with the observations made by the Dawn spacecraft. Consistent with IAU best practices the Dawn Science team chose a small crater near the equator to anchor the coordinate system. The prime meridian was located for the convenience of the mappers, whose processing the data to spot errors by visual inspection. Coordinate system errors were not #'%! #'&! #''! #'(! quadrangles are aligned with the prime meridian. It is traditional in mapping planetary #')! surfaces to have the quadrangles all begin aligned with the prime meridian. The choice of #'*! meridian. Further, the prominent ‘Snowman’ craters, Marcia, Calpurnia and Minucia, the Dawn coordinates ensured that. No significant features were split by the prime #(+! were centrally located. It was believed to be safer for the users of the Dawn observations #("! #(#! to work with a unique coordinate system that is clearly distinct from the early coordinate #($! systems, which are based on various pole locations. These earlier coordinate systems (Li #(%! et al., 2011; Archinal et al., 2011; Seidelmann et al., 2007; Seidelmann et al., 2005; #(&! Claudia coordinate system. With the large difference between the Claudia system and the Seidelmann et al., 2002; Thomas et al., 1997) are nearer to the 180° meridian of the #('! #((! earlier systems, the use of the Claudia coordinate system can be easily visually verified. #()! #(*! #)+! uncommon, but were found promptly with this device (S. Joy, personal comm., 2011). #)"! Claudia was chosen out of the innumerable small craters in the general location of #)#! Also it can be simply located by the use of a number of marker craters. Claudia is the prime meridian, first because according to standard practices, it is close to the equator. #)$! #)%! located roughly midway between Oppia and Gegania craters and to the north of Divalia #)&! #)'! 1a), in which Claudia lies on the bar of the H. Near the midpoint of this bar are two ! ""! In practice, this choice of coordinate systems enabled those operating the spacecraft, and Fossa. Claudia and eight larger craters form a cursive uppercase H (dotted lines in Fig. Taylor, 1973; Chang and Change, 1962; Haupt, 1958; Magnusson, 1986). Although the rotational period of Vesta has been determined to an accuracy of 10-7 days, there was an #)(! craters that are merged together and have a small crater in the center of their eastern rim #))! Claudia was also chosen because it has a distinctive morphology, which makes it (see Fig. 1b). Claudia is roughly 3.5 kilometers to the southeast of this small crater. #)*! #*+! distinguishable from other similarly sized craters (see Fig. 1c). Claudia has a sharp, fresh #*"! rim; there is a ~170 meter diameter crater on the northeastern rim and two ~80 meter #*#! craters, along with many smaller craters, within Claudia’s interior. #*$! All early determinations of the rotational period of Vesta were performed through its 4. Comparison of Rotational Period and Pole Orientation #*%! #*&! rotational lightcurves (e.g., Cuffey, 1953; Groeneveld and Kuiper, 1954; Gehrels, 1967; #*'! #*(! #*)! #**! its lightcurve is single-peaked or double-peaked. Later observations using polarimetry, $++! speckle interferometry, and radar favored the shorter period and showed that the surface $+"! Drummond et al., 1988; Taylor et al., 1985). With the help from speckle interferometry of Vesta is variegated possibly due to compositional heterogeneity (Degewij et al., 1979; $+#! $+$! and adaptive optics (AO), the surface of Vesta was resolved and the lightcurve of Vesta is $+%! Drummond et al. (1998) determined a rotational period of 0.22258874 day with an shown to be single-peaked (Drummond et al., 1988). Using a large dataset of AO images, $+&! $+'! uncertainty of 4 on the last decimal place (3.5 ms). This is the most accurate $+(! $+)! $+*! determination of Vesta’s rotational period before Dawn’s arrival at Vesta. The improved rotational period measurement from Dawn is 0.222588652 day with an uncertainty of 35 ambiguity of whether Vesta’s rotational period is 5.34 hr or 10.68 hr, rising from whether !s (Russell et al., 2012). ! "#! determine its pole orientation based on surface albedo features and the limb profiles of the disk. They reported a pole orientation of (RA=301º, Dec=41º) with an uncertainty of sizes of 54 km and 38 km at Vesta, respectively, to construct a shape model of Vesta and 10º on both RA and Dec. Drummond and Christou (2008) fitted an ellipsoidal model to $"+! HST observations before Dawn arrived at Vesta with different datasets and techniques The pole orientation of Vesta was previously determined from ground-based and $""! $"#! (Table 3). Thomas et al. (1997a) used HST images acquired in 1994 and 1996 with pixel $"$! $"%! $"&! $"'! $"(! the limb of observed disk of Vesta from various observing and illumination conditions $")! using ground-based, disk-resolved observations with speckle interferometry and adaptive $"*! with an uncertainty of 7º. Li et al. (2011) combined all previous determinations of optics techniques over the past 27 years. Their best-fit pole orientation was (306º, 38º) $#+! Vesta’s pole orientation and their newly acquired images from HST in 2007 and 2010 at $#"! $##! similar pixel size of previous HST images. They experimented with two methods, $#$! namely control point stereogrammetry and feature tracking, on all four HST datasets $#%! combined, and performed statistical studies of previous pole determination with ground- $#&! measurements was (305.8º, 41.4º)±(3.1º, 1.5º) by Li et al (2011). While in orbit around based data. The latest measurement of Vesta’s pole orientation before Dawn $#'! Vesta, the Dawn spacecraft returned a substantial amount of imaging and tracking data, $#(! which have been used to precisely determine the pole orientation of Vesta as (309.03º, $#)! $#*! 42.23º) with an uncertainty of 0.01º (Russell et al., 2012). The obliquity of Vesta is $$+! $$"! $$#! 27.46º based on the pole measurement from Dawn. All previous pole measurements put the obliquity within 3º of this value (Table 3). 5. Comparisons between HST and Dawn FC lightcurves ! "$! HST observed Vesta four times in 1994 (Zellner et al., 1996), 1996 (Thomas et al., 1997b), 2007 (Li et al., 2010), and 2010 (Li et al., 2011), all with WFPC2 through the its rotational axis. The difference between its intermediate axis and long axis is less than $$$! $$%! $$&! same set of filters; F439W, F673N, F953N, and F1042M. While the observing $$'! circumstances were all different, the 2007 observations have a sub-solar latitude of -5.3º, $$(! Vesta (south of -23º) among all HST observations. To compare the whole-disk being the closest to the sub-solar latitude of Dawn data collected while approaching $$)! $$*! HST 2007 observations and FC images collected on June 30, 2011 that cover one full lightcurves of Vesta collected by Dawn FC with those obtained by HST, we focus on $%+! $%"! rotation of Vesta. The observing geometries of both observations are listed in Table 4. $%#! $%$! $%%! 3% (Thomas et al., 1997b; Russell et al., 2012). The rotational lightcurve of Vesta is $%&! single-peaked, and dominated by the albedo variations on its surface rather than by cross- $%'! Due to moderate multiple scattering on Vesta (Li et al., 2012), the limb-darkening effect sectional area variation or scattering geometry variation from global-scale topography. $%(! $%)! is stronger on Vesta than that on darker objects such as the Moon and C-type asteroids. $%*! $&+! $&"! $&#! $&$! $&%! $&&! to compare the lightcurves obtained at different geometries. This is especially necessary lightcurve than those near the limb or terminator. For this reason, we plotted the brightness of Vesta with respect to sub-solar longitude rather than sub-observer longitude for our case because the Dawn FC imaged the morning side of Vesta, while HST imaged the slightly afternoon side of Vesta, and the phase angles of both observations are quite different (Table 4). ! "%! The shape of Vesta is close to a tri-axial ellipsoid, with the short axis aligned to Therefore albedo features near sub-solar point contribute slightly more to the shape of its difference. Dawn images have much more southern latitude than that of HST images, and in mid northern latitude (30º-60º) the surface of Vesta appears to be brighter in this range $&'! Dawn FC images (red and orange). At all three wavelengths, the overall agreement Fig. 2 shows the lightcurves of Vesta measured from HST images (green) and $&(! $&)! between two observations is excellent in terms of the shape, amplitude, and phase of $&*! lightcurves. The Dawn FC lightcurve is much smoother than the HST lightcurve, $'+! where Vesta is resolved to 9.2 km/pix, compared to 38 km/pix in 2007 HST images. The presumably due to much higher signal-to-noise ratio in Dawn FC images measurement $'"! $'#! slight difference at sub-solar longitude from 60º to 330º could be due to viewing angle $'$! $'%! $'&! of longitude than in other longitudes. $''! Compared to the albedo maps (Fig. 4), the large bright and dark features in the $'(! minimum is located between longitude 270º and 330º, corresponding to the broad dark lightcurves (Fig. 2) all correspond to the large-scale albedo features. The lightcurve $')! $'*! area on Vesta observed in the same longitude. The lightcurve maximum is located $(+! $("! $(#! $($! $(%! $(&! $('! $((! $()! illuminating its topography under the same viewing conditions as the image and ! "&! The topography of Vesta was constructed using the technique of Gaskell et al. (2008) to the lightcurve minimum, the lightcurve peak appears to be slightly round-shaped or plateaued. This is due to the large, relatively dark area near Oppia near ~100º longitude from over 150000 small 99x99 pixel maplets, each representing the topography and relative albedo of a small patch of Vesta’s surface. Each maplet is located in an image by between longitude 60º to 120º, when the bright eastern hemisphere is visible. Compared surrounded by broad bright areas. 6. Comparison of Vesta shape model from HST and Dawn located in many images in a wide range of illuminations, resolutions and viewing (Gaskell, 2011). The maplets themselves are constructed by adjusting the slope and relative albedo at each of its pixels to minimize the residuals between the image performing a simple correlation. The center of the maplet represents a control point in image space, and due to the maplet’s three-dimensional structure, its control point can be geometries. This leads to an extremely precise solution for the location of the maplet’s center. Similarly, the positions of many control points in a single image yield a precise determination of the Vesta-relative camera position and pointing at the image time $(*! $)+! $)"! $)#! $)$! $)%! $)&! $)'! $)(! brightness and illuminated maplet pixel brightness over a large number of images. The $))! slopes are then integrated to produce the topography distribution within the maplet. $)*! Vesta’s surface (Gaskell, 2012, private communication). Figure 3B shows this The ensemble of maplets can be combined to produce a global topography for $*+! $*"! topography in a frame very close to the original Thomas et al. (1997a) coordinate system $*#! $*$! $*%! $*&! $*'! $*(! $*)! $**! %++! %+"! diameter of 460 km with an average depth below the rim of 13 km. The central peak ! "'! 285 km x 285 km x 229 km ellipsoid. Thomas et al. (1997a) were able to glean the topography of Vesta from images gathered by the Hubble Space Telescope. Their map, relative to a 289 km x 280 km x 229 km ellipsoid is shown in Figure 3A. There are striking similarities between the Thomas Hubble topography and the Dawn determined system. The heights range from -22.45 km (blue/violet) to +19.48 km (red) relative to a results. The Thomas range of heights was slightly smaller, but in general the agreement is very good and the differences can be explained in part by the dramatic difference in resolution between the two data sets. Thomas et al. (1997a) note that the south pole (now called Rheasilvia) basin has a higher than other parts. This region corresponds to Matronalia Rupes in Dawn data and rises higher than the rest of the basin rim as observed in HST data (Schenk et al. 2012). 7. Comparison of HST Observations and Dawn FC Data 7.1 Global View 1997a). In comparison, Dawn data show that the basin has a diameter of ~500±25 km and a depth of 19±6 km (Schenk et al. 2012). While these values are greater, they are consistent with HST model given the resolution of the data and the reference ellipsoids Figure 4a shows the HST map of Vesta in 0.673-!m filter projected in the Thomas et al. (1997a) coordinate system based on observations from 1994, 1996 and 2007 oppositions at a resolution of ~50 km/pixel. The prime meridian here is defined through the dark within the basin is 13 km high compared to its deepest parts in HST map (Thomas et al. used. Thomas et al. (1997a) also note that a small section of the Rheasilvia rim is ~ 5 km %+#! %+$! %+%! %+&! %+'! %+(! %+)! %+*! %"+! %""! %"#! %"$! %"%! feature informally named “Olbers” in Thomas et al. (1997a) located at ~15°N. The pole %"&! position for this data is "0 = 301°± 5°, #0 = 41°± 5°. Figure 4b shows the Dawn FC map %"'! meridian similar to Thomas et al. (1997a) coordinate system in Fig. 4a. The pole position of Vesta in 0.75-µm filter from RC1 at a resolution of 9.06 km/pixel with the prime %"(! %")! ("0 = 309.03° ± 0.01°, 42.23° ± 0.01°) is the updated one from Russell et al. (2012) and %"*! Dawn data does not extend to northern latitudes in this map because of the spacecraft hence features in the Dawn FC map are slightly rotated with respect to HST map. The %#+! %#"! location with respect to Vesta during the RC1 phase. Fig. 4c is similar to Fig. 4b but the %##! %#$! %#%! help ground-based observers orient themselves to features on Vesta from Dawn data. ! "(! prime meridian is defined in the Claudia coordinate system (Russell et al. 2012) used by the Dawn science team in all their publications. The purpose of these three figures is to 7.2 East-West Dichotomy Figure 4A-C also helps identify several hemispherical scale albedo features first observed in HST data by Thomas et al. (1997a). Despite the differences in pole position between Figures 4a and 4b, the Western hemisphere has an overall lower albedo than the Eastern hemisphere. Based on rotationally resolved near-IR spectra, Gaffey (1997) suggested that the dark hemisphere could be a howardite or polymict eucrite regolith that has been darkened by ‘age-related darkening effect.’ Binzel et al. (1997) was the first to %#&! %#'! %#(! %#)! %#*! %$+! %$"! %$#! detect this East-West dichotomy from HST images and interpreted the Western %$$! hemisphere as being “dominated by iron-rich and relatively calcium-rich pyroxene” %$%! Vesta could be impact craters/basins filled with dark material similar to lunar mare. Near- similar to basaltic flows like eucrites. Zellner et al. (2005) suggested albedo features on %$&! %$'! IR observations obtained by Vernazza et al. (2005) indicated that the Western hemisphere %$(! %$)! %$*! %%+! %%"! %%#! %%$! previous studies (16°) (Gaffey 1997 and Binzel et al. 1997). %%%! Reddy et al. (2012a) studied the albedo and color variations on Vesta using Dawn %%&! maps from Binzel et al. (1997) and Li et al. (2010) at hemispherical scale (Fig. 5 and 6). FC color images. These albedo/color maps are in excellent agreement with HST albedo %%'! %%(! Interpreting the composition of the Western hemisphere, they concluded that it is ! ")! hemisphere is dominated by howardite/polymict eucrite. Li et al. (2010) and Reddy et al. ground-based spectral observations, Reddy et al. (2010) concluded that the Western (2010) showed that this hemisphere is predominantly a large eucritic unit using the band ratio relationships associated with spectral characteristics of HED meteorites. Using (2010) observations were also made at southerly sub-Earth latitude (-18.9°) compared to could be dominated by eucrite-type material. More recently, HST observations by Li et al. %%)! dominated by a howardite and/or polymict eucrite, which is in agreement with ground- %%*! However, Reddy et al. (2012a and 2012b) conclude that the cause of the lower albedo in based (Gaffey, 1997; Vernazza et al. 2005) and HST observations (Binzel et al. 1997). %&+! %&"! the Western Hemisphere is not due to an age darkening effect (Binzel et al. 1997) or %&#! lunar mare type material filling impact craters/basins (Zellner et al. 2005) but due to the %&$! in fall of carbonaceous chondritic material. The presence of exogenous carbonaceous %&%! Vesta is well documented (e.g., Buchanan et al., 1993; Zolensky et al., 1996). In HEDs, chondrite meteorite clasts in some howardite-eucrite-diogenite (HED) meteorites from %&&! %&'! carbonaceous chondrite clasts generally make up to 5 vol. % (Zolensky et al., 1996), but %&(! on rare occasions can be 60 vol. % of howardites (Herrin et al. 2011). While some of %&)! these clasts have been heated and dehydrated during impact (~400°C), a majority of them %&*! Carbonaceous chondrite clasts are the largest exogenic material observed in HED are still hydrated (containing H2O- and/or OH-bearing phases) (Zolensky et al., 1996). %'+! meteorites, but surprisingly they were not suggested as analogs for the dark hemisphere %'"! %'#! on Vesta by most ground-based and HST studies. Ground-based telescopic observations %'$! %'%! %'&! suggested contamination from impacting carbonaceous chondrites as possible cause of %''! this feature. %'(! Hasegawa et al. however, the 2006 observations were at a different viewing geometry Rivkin et al. (2006) observations did not find the same feature reported in %')! %'*! than 2003 and did not extend as far North as 2003, and variability of spatial extent of the %(+! 3-!m feature was offered as a possible explanation. A weak absorption band at 2.8 !m ! "*! of Vesta in the mid-IR indicated the possible presence of a 3-!m absorption feature (Hasegawa et al. 2003). Hasegawa et al. (2003) noting the presence of a 3-!m absorption hemisphere as impact excavated plutonic material similar to diogenites. They also identified several compositionally distinct units using band parameter systematics and concluded that several Eastern hemisphere units have substantial olivine component (more discussion on olivine in Section 6.3). Vernazza et al. (2005) and Carry et al. (2010) Eastern hemisphere of Vesta using ground-based spectral and adaptive optics data. Li et also interpreted spectral variations in near-IR to indicate a substantial diogenite in the %("! attributed to OH adsorption is reported with variable expression in the spatial domain %(#! (DeSanctis et al., 2012b). Evidence of hydrogen enrichment in the same region has been %($! water-related feature near the vibrational bands of hydrated species. found by Dawn (Prettyman et al. 2012) and confirms the ground-based evidence for a %(%! %(&! Gaffey (1997) and Binzel et al. (1997) interpreted the higher albedo Eastern %('! %((! %()! %(*! %)+! %)"! %)#! al. (2010) and Reddy et al. (2010) concluded there might be a large diogenite unit %)$! straddling the boundary between the two hemispheres. %)%! material on Vesta using Dawn FC color images and color indicators of HED components. Reddy et al. (2012a) mapped the abundance and distribution of diogenite-rich %)&! Albedo (0.75-!m filter), pyroxene band depth (0.75/0.92-!m filter ratio) and eucrite- %)'! %)(! diogenite (ED) ratio (0.98/0.92-!m filter ratio) maps of Vesta (Fig. 6a-c) from Dawn FC %))! map shows diogenite-rich areas in red and eucrite-rich areas in blue (Fig. 6c). The confirm the presence of diogenite-rich material in the southern hemisphere. The ED ratio %)*! %*+! distribution of diogenite is predominantly restricted to the southern hemisphere except for %*"! %*#! %*$! a small region that stretches north between 120°-180° W longitude that Reddy et al. (2012a) interpreted as ejecta from the Rheasilvia basin. 7.3 North-South Dichotomy ! #+! olivine present within the upper mantle or both.” Reddy et al. (2010) observed the the Southern hemisphere of Vesta. The ED ratio map of Vesta produced from Dawn data (Fig. 6c) shows more Thomas et al. (1997a) studied the southern hemisphere of Vesta using HST. They noted that the band depth and width of the 0.9-!m pyroxene absorption increased with band position that Vesta’s southern hemisphere is more diogenitic than the northern hemisphere. These conclusions were consistent with Li et al. (2010) HST observations of “excavation depth, consistent with exposure of a higher calcium content and coarser grained pyroxene-rich plutonic assemblage within the crust of Vesta or the exposure of diogenite-rich material (red) in the southern hemisphere centered around the Rheasilvia southern hemisphere of Vesta using the IRTF and concluded from near-IR absorption %*%! %*&! %*'! %*(! %*)! %**! &++! &+"! &+#! &+$! &+%! &+&! basin and bluer eucritic areas in the north confirming ground-based (Gaffey 1997, Reddy &+'! et al. 2010) and HST observations (Thomas et al. 1997; Li et al. 2010) of Vesta. There is &+(! map (Fig. 6c). Brighter areas in the albedo map correspond to areas with deeper band also a distinct correlation between the albedo (Fig. 6a), band depth map (Fig. 6b), and ED &+)! &+*! weaker band depth and more eucrite-rich areas. depth and diogenite-rich areas and conversely darker areas correspond to those with &"+! &""! Reddy et al. (2012a) explored the cause of deeper band depth in diogenite-rich &"#! areas and concluded that this could be a particle size effect as suggested by Thomas et al. &"$! (1997a). Based on meteoritic evidence, the average grain size of diogenites in howardite &"%! material is $70 µm (Beck and McSween, 2011). This difference could be due to a breccias ranges from 500 µm to 1.5 millimeters whereas the average grain size of eucritic &"&! &"'! combination of surface composition and regolith processes (Chamberlain et al. 2007). ! #"! could also be explained by shorter surface exposure ages (fresher impacts) compared to eucrites that might have experienced longer exposures to regolith processes on the that included albedo, band depth, and width (ED). Following similar methods, Li et al. (2010) expanded this list to include 15 new features (numerically ordered) in the southern hemisphere. Of these 34 features identified in HST images, a few are prominently visible Diogenites cooled at depth with coarser (pyroxene) grains compared to eucrites, which are typically fine-grained basalts (surface flows). The larger grain size for diogenites &"(! &")! &"*! &#+! &#"! surface (impact gardening) (Reddy et al. 2012a). The latter would indicate Rheasilvia &##! basin forming more recently excavating diogenite rich material. &#$! Binzel et al. (1997) identified 19 different albedo and color units on Vesta using HST 7.4 Cross correlation between HST and Dawn FC Albedo Maps &#%! &#&! observations. These alphabetically ordered units were selected for their spectral diversity &#'! &#(! &#)! &#*! in albedo maps while others stand out in band depth and ED ratio maps. Figure 7A shows &$+! those units in the albedo map of Vesta from HST in 0.673-!m filter and Fig. 7b shows &$"! maps are in Thomas et al. (1997a) coordinate system. the same units as identified in the albedo map from Dawn FC in 0.75-!m filter. Both &$#! &$$! Albedo features from the HST map have been identified and labeled on the Dawn &$%! FC map by correlating albedo patterns visually and confirming them based on the &$&! most prominent albedo features that are easily identifiable in both maps include low description of their spectral properties from Binzel et al. (1997) and Li et al. (2010). The &$'! &$(! Q and #15. High albedo units that are relatively easy to identify in both maps include albedo features such as “Olbers,” features #8, #9, A, and B within the dark unit, features &$)! &$*! features #4, #15, #13, #11, and Z. Given the differences in spatial resolution (50 km/pixel ! ##! longitudinally and helped create the first maps (Gaffey, 1997). Subsequent disk-resolved color filter observations by HST (Thomas et al. 1997) led to the identification of albedo features associated with these units (Binzel et al. 1997) in the northern hemisphere. Li et based on ground-based spectra. vs. 9 km/pixel) and pole orientation, some diffuse features in the HST map (Fig. 7a) spectral parameters (Band Centers and Band Area Ratio or BAR) with rotational cannot be confirmed in the Dawn FC map shown in Fig. 7b. A complete list of HST units lightcurve of Vesta, the location of distinct mineralogical units could be constrained &%+! &%"! &%#! and their corresponding surface features with IAU approved names from Dawn FC &%$! images is in Table 5. &%%! Ground-based rotationally resolved spectroscopy of Vesta led to the identification of 7.5 Interpretations of specific features &%&! &%'! specific compositional units on Vesta (Gaffey, 1997). By correlating variations in &%(! &%)! &%*! &&+! &&"! &&#! &&$! &&%! &&&! Thomas et al. (1997a) observed three distinct regions that had showed variations &&'! in band depth and band width. These include the Rheasilvia basin (discussed in section &&(! Matronalia Rupes feature on Dawn FC maps corresponds to feature #7 (Table 5) on HST 6.1), Matronalia Rupes on the rim of the South Pole crater, and a potential crater. &&)! maps. This 20-km high scarp (Schenk et al. 2012) stretches for nearly 200 km around the &&*! Rheasilvia basin. Thomas et al. (1997a) noted a reverse compositional trend in &'+! Matronalia Rupes “consistent with composition of excavated material that has been &'"! &'#! overturned.” Dawn FC color images show Matronalia Rupes is dominated by diogenite- ! #$! al. (2010) identified additional compositional units in the southern hemisphere of Vesta from HST images and Reddy et al. (2010) linked these to specific mineralogical units rich material (Reddy et al. 2012a) excavated during the formation of the Rheasilvia basin &'$! &'%! consistent with Thomas et al. (1997a) interpretation. Li et al. (2010) also identified &'&! several features (#4, #5, #6) that are interpreted as diogenite-rich areas “located on high &''! within the diogenite-rich ejecta blanket that spreads north between 180° and 120°W rims of the crater.” All these units are located along the rim of the Rheasilvia basin or &'(! &')! longitude in Fig 6A-C. This ejecta blanket and diogenite units have also been observed in &'*! ground-based compositional map from Gaffey (1997) and Reddy et al. (2010). &(+! Gaffey (1997) noted a decrease in ratio of 2-!m pyroxene band (Band II) area to &("! Vesta rotated. A well established cause for decrease in BAR is increase in the abundance that of 1-!m pyroxene band (Band I) area from ground based spectral observations as &(#! &($! of olivine in a mixture of olivine and orthopyroxene (Cloutis et al. 1986). Based on this &(%! &(&! surface of Vesta located between 60-120° E longitude (Fig. 8-9). This olivine-rich unit &('! geographically corresponds to feature #15 from HST map of Li et al. (2010) (Fig. 7A-B). &((! HST (Li et al. 2010) with band depth shallower than the background surface. This feature has the reddest spectra color over the entire surface of Vesta observed by the &()! &(*! In Dawn FC color maps (Fig. 6A-C) this feature corresponds to the Oppia crater &)+! &)"! &)#! &)$! &)%! &)&! composition of Oppia region, Le Corre et al. (2012) estimated the BAR as 1.57, which is ! #%! and its ejecta blanket. Reddy et al. (2012a) observed that Oppia has the reddest visible spectra slope (0.75/0.44 !m) on Vesta consistent with Li et al. (2010) HST observations. The band depth (0.75/0.92 !m) of Oppia region is also shallower than the average surface interpretation Gaffey (1997) suggested the presence of a large olivine-rich unit on the of Vesta (Reddy et al. 2012a). Due to low albedo, red slope and weaker pyroxene bands, Li et al. (2010) interpreted this region to be space weathered (Fig. 8). Investigating the much lower than the average BAR (2.74) for Vestan surface (Gaffey, 1997). Le Corre et feature on Vesta as “Olbers Regio” in honor of its discoverer Wilhelm Olbers. This feature also defined the location of the prime meridian in HST maps. Binzel et al. (1997) by the Oppia impact. Hence the low BAR for Oppia region, which was interpreted as olivine by Gaffey (1997), might not be interpreted so today. The best way to determine al. (2013) attributed this low BAR to impact melt or excavation of cumulate eucrite layer &)'! &)(! &))! &)*! &*+! the presence of olivine on Vesta’s surface is through the interpretation of higher &*"! resolution data from Dawn. We also rule out selective lunar-style space weathering as the &*#! Oppia (Pieters et al. 2012). cause of the red visible slope and lower BAR observed in color and spectral data of &*$! &*%! As noted earlier, Zellner et al. (1997) unofficially named a prominent low albedo &*&! &*'! &*(! interpreted this low albedo feature as “remnants of Vesta’s ancient basaltic crust.” They &*)! also noted that the pyroxene band depth of the Olbers feature was relatively shallow and &**! weathering to explain the low albedo and shallow pyroxene absorption band of Olbers, narrow suggesting the predominance of a single pyroxene. Invoking lunar-style space '++! Binzel et al. (1997) concluded that Olbers is remnant ancient crust. '+"! '+#! Dawn FC observations of Vesta show that the Olbers region in the HST map '+$! with HST observations by Binzel et al. (1997) this region has low albedo and weaker corresponds to the dark area East of Marcia, Calpurnia, and Minucia craters. Consistent '+%! '+&! pyroxene band compared to Vesta global average. Prettyman et al. (2012) noted an '+'! '+(! '+)! chondrite xenoliths. Although our study of this region is still ongoing, the ejecta ! #&! enhancement in hydrogen signature corresponding to Olbers region based on GRaND observations of Vesta and attributed this to the presence of H2O/OH in carbonaceous band of this feature is indistinguishable from the background surface in HST observations. feature within the Rheasilvia basin. The Framing Camera shows that Antonia crater is from Li et al. (2010). They observed that this feature has an extremely blue spectral slope and is located on the steepest slope(s) on Vesta. The depth of the pyroxene absorption In ground based observations of Vesta (Binzel et al. 1997; Gaffey, 1997) this feature corresponds to a diogenite-rich unit located ~45° S latitude. In Dawn FC images, this feature corresponds to ejecta around the Antonia crater, a 17.4–km diameter impact indeed located on a topographically steep slope on Vesta consistent with HST '+*! distribution around Marcia suggests that it is younger than low albedo region east of it. '"+! mixing of carbonaceous chondrite impactor material and ancient howardite regolith. The low albedo and weaker pyroxene bands of Olbers region are probably due to the '""! '"#! One of the brightest albedo spots in HST observations of Vesta is feature #13 '"$! '"%! '"&! '"'! '"(! '")! '"*! '#+! '#"! '##! '#$! '#%! '#&! '#'! '#(! '#)! interpreted as diogenite-rich consistent with HST observations. '#*! 7.6 Space Weathering '$+! modified when exposed to space environment in the absence of an atmosphere. On the Space weathering consists of processes by which the surface optical properties are '$"! ! #'! observations by Li et al. (2010). ED ratio map of Vesta (Fig. 6b-c) showed that the region around Antonia crater is dominated by diogenite-rich material consistent with ground- based observations by Gaffey, (1997). Of the six geologic units identified by Li et al. (2010) in HST images (Fig. 9), area IV (60° E longitude) showed a deep pyroxene band, neutral red slope and was interpreted as diogenite rich. In Dawn FC data (ED map), this region corresponds to 30- km diameter Numisia crater, which lies SE of Olbers region. Numisia has been increased level of space weathering (Murchie et al. 2002). Space weathering on Vesta has been a source for intense scrutiny over the last two decades because of its similar basaltic composition as that of the Moon. Detection of large-scale albedo variations on Vesta using HST suggested that some of them might be ancient ‘space weathered’ units (Binzel et al. 1997). Hiroi and Pieters (1998) conducted grains causing these observed effects (Loeffler et al. 2008). In contrast to the Moon, Moon the primary effect of space weathering is decrease in optical albedo, increase in spectral slope (reddening), and a decrease in absorption band depth (Pieters et al. 2000; Taylor et al. 2001). Nano phase iron created during space weathering coats the lunar soil '$#! '$$! '$%! '$&! '$'! asteroids visited by spacecraft prior to the arrival of Dawn at Vesta showed distinctly '$(! different effects of space weathering (Gaffey, 2010). On (243) Ida, the band depth '$)! with increased level of space weathering (Veverka et al. 1996). On (433) Eros, the overall decreased and the spectral slope increased, while the albedo remained nearly constant '$*! '%+! albedo decreased but the spectral slope and band depth remained nearly constant with '%"! '%#! '%$! '%%! '%&! '%'! simulated space weathering experiments on HED meteorites and compared them to Vesta '%(! and Vestoids. Vestoids are small asteroids (~10 km) that have spectral properties similar '%)! Vestoid spectra showed effects similar to lunar style space weathering, but some Vestoid to Vesta and are part of the Vesta dynamical family. They found that Vesta and some '%*! '&+! spectra did not follow the “HED-lunar space weathering trend.” They concluded that '&"! '&#! '&$! '&%! effects. Meteoritical evidence for space weathering products in Vestan regolith has been ! #(! these Vestoids could be from a different parent body. Addressing the cause of “bluer” visible spectra of Vesta, Vernazza et al. (2006) invoked the possibility of a magnetic field around Vesta that could be protecting the surface against lunar-style space weathering limited. Noble et al. (2011) searched for space weathering products (agglutinates and of space weathering operating on each of these objects (Gaffey, 2010). 8. Future Earth-Based Observations principle cause of lunar space weathering effects, on Vesta’s regolith particles. Gradual fading of freshly exposed material seen on Vesta has been attributed to localized mixing of material with diverse spectral and albedo properties (Pieters et al. 2012, McCord et al. 2012, Li et al. 2012). McCord et al. (2012) have shown that impact mixing of bright '&&! '&'! nano phase iron bearing rims) in howardite Kapoeta and found several “melt products, '&(! including spherules and agglutinates and possible nano phase bearing rim.” They '&)! weathering is active on asteroids like Vesta. concluded that the presence of these products would suggest that lunar style space '&*! ''+! Vesta’s surface shows effects of space weathering but it is unlike that seen on the Moon. Analyzing Dawn FC and VIR spectral data, Pieters et al. (2012) noted that ''"! ''#! They also found no evidence for the accumulation of nano phase iron, which is the ''$! ''%! ''&! '''! ''(! pristine material (Li et al. 2012) with darker exogenous carbonaceous chondrite material '')! (Reddy et al. 2012) gives rise to the background gray material. In light of this new ''*! weathering to all asteroids is invalid. This result is also consistent with space weathering information from Dawn, it appears that general application of lunar style space '(+! '("! trends observed on other asteroids by spacecraft (Ida and Eros) that suggest unique styles '(#! '($! '(%! '(&! '('! '((! only be possible with an orbital mission. Based on the comparison of ground-based and ! #)! better prepared to explore Vesta. While the results from the mission has confirmed many pre-Dawn characterizations of Vesta, a myriad of surprises were revealed, which could Studies of Vesta by ground-based and orbiting telescopes helped the Dawn mission to be based observations have been extremely accurate at constraining rotation period, surface HST data of Vesta with Dawn’s Framing Camera data, we can conclude that ground composition using reflectance spectroscopy, and detecting/mapping rotational spectral variations on small bodies. HST observations have also been successful in accurately limitations in accessing the prevalence and magnitude of space weathering on individual '()! '(*! ')+! ')"! ')#! identifying albedo and color features, surface composition of these features and ')$! constraining the shape, pole orientation and topography. The success of these precursor ')%! HST data, and b) interpretations backed by laboratory studies of HED meteorites that studies can be attributed to two factors: a) availability of high quality ground-based and ')&! were delivered to Earth from Vesta. ')'! ')(! However, ground-based observations and their interpretations have severe '))! ')*! '*+! '*"! '*#! '*$! '*%! '*&! '*'! '*(! '*)! '**! (++! small bodies. As noted earlier, space weathering on Vesta (Pieters et al. 2012) and other small bodies visited by spacecraft has been shown to be unlike that on the Moon (Gaffey, large and small projectiles control the distribution and mixing of materials in Vesta’s regolith. Future interpretations of ground-based spectra of small bodies must take into account these findings from Dawn and other missions before interpreting red spectral slope in ground based spectra of asteroids as an indicator of space weathering. 2010). On Vesta, alternate processes such as presence of exogenous opaques (Reddy et al. 2012b; McCord et al. 2012), observing geometry/phase angle (Reddy et al. 2012c; Sanchez et al. 2012) and particle size (Reddy et al. 2012a; Pieters et al. 2012) have been shown to be the cause of traditional lunar-style space weathering effects like low albedo, red spectral slope, and weaker absorption bands. The effects of multiple impacts from ! #*! determination of Vesta’s rotational period before Dawn’s arrival at Vesta. • Dawn spacecraft has precisely determined pole orientation of Vesta to be (309.03º, 2012) is 0.222588652 days with an uncertainty of 35 !s. This is consistent with Drummond et al. (1998) rotational period of 0.22258874 day, the most accurate 42.23º) with an uncertainty of 0.01º (Russell et al., 2012). This is a substantial improvement over the best ground/HST based pole position (305.8º, 41.4º)±(3.1º, (+"! Ground-based and HST observations of Vesta provided valuable insight into the 9. Summary and Conclusions (+#! (+$! Dawn spacecraft. Our comparative study reveals the following: asteroid’s composition, surface albedo and topography prior to the arrival of NASA’ (+%! (+&! • Rotation period of Vesta as determined by the Dawn spacecraft (Russell et al., (+'! (+(! (+)! (+*! ("+! (""! ("#! ("$! ("%! ("&! ("'! ("(! (")! ("*! (#+! (#"! (##! • While the Dawn FC lightcurve is much smoother than the HST lightcurve due to • Global albedo/color maps of Vesta using HST (Binzel et al. 1997; Li et al. 2010) • HST topographic map from Thomas, et al (1997) is consistent with Dawn shape model. There are striking similarities between HST topography and the Dawn determined results. HST range of heights was slightly smaller (-12 km to +12 km) than Dawn (-22.45 km to +19.48 km), likely due to much lower spatial resolution much higher signal-to-noise ratio, the overall agreement between the two is excellent in terms of the shape, amplitude, and phase of lightcurves. 1.5º) by Li et al. (2010). of the former, but in general the agreement is very good. are in excellent agreement with Dawn FC maps (Reddy et al. 2012). East-West ! $+! (#$! (#%! (#&! (#'! (#(! (#)! (#*! ($+! ($"! ($#! ($$! ($%! ($&! ($'! ($(! ($)! ($*! (%+! (%"! (%#! (%$! (%%! (%&! hemispherical dichotomy observed in ground-based (e.g., Gaffey 1997) and HST images (Thomas et al. 1997a) has been confirmed by Dawn FC color images. • Ground-based telescopic observations indicated the possible presence of a 3-!m absorption feature. Hasegawa et al. (2003) and Rivkin et al. (2006) suggested contamination from impacting carbonaceous chondrites as possible cause of this feature. Dawn observations of Vesta (De Sanctis et al. 2012b; Prettyman et al. 2012; McCord et al. 2012; Reddy et al. 2012b) confirm these ground-based observations. • Predominance of diogenite-rich material in the higher albedo Eastern hemisphere observed by ground-based observers (Gaffey 1997; Binzel et al. 1997; Vernazza et al. 2005; Carry et al. 2010; Li et al. 2010) has been confirmed by Dawn (Reddy et al. 2012a). • A majority of albedo and color units observed in HST images of Vesta (Binzel et al. 1997; Li et al. 2010) have been confirmed by Dawn FC observations. • Dawn data supports an alternative interpretation of “olivine-rich” unit observed by Gaffey (1997). The low BAR for Oppia region, which was interpreted as olivine by Gaffey (1997), would not be so interpreted based on Dawn data (Le Corre et al. (2011). • Ground-based observations have severe limitations in identifying space weathering and constraining its magnitude on individual small bodies. Dawn observations have demonstrated that lunar-style space weathering does not operate on Vesta, despite both objects having a basaltic surface composition (Pieters et al. 2012). ! $"! its physical properties. This can be attributed to two factors: a) availability of high • Precursor studies of Vesta have been very successful in accurately characterizing (%'! (%(! (%)! quality of ground-based and HST data, and b) interpretations backed by (%*! laboratory spectra and studies of samples of HED meteorites (&+! Acknowledgement (&"! We thank the Dawn team for the development, cruise, orbital insertion, and operations of (&#! (&$! the Dawn spacecraft at Vesta. The Framing Camera project is financially supported by (&%! Dawn at Vesta Participating Scientist Program for funding the research. A portion of this the Max Planck Society and the German Space Agency, DLR. We also thank NASA’s (&&! work was performed at the Jet Propulsion Laboratory, California Institute of Technology, (&'! (&(! VR would like to thank Juan Andreas Sanchez and Guneshwar Singh Thangjam for their under contract with NASA. Dawn data is archived with the NASA Planetary Data System. (&)! (&*! help in improving the manuscript. ('+! References ('"! Archinal, B.A., et al., 2011. Report of the IAU/IAG Working Group on cartographic ('#! ('$! ('%! Beck, A.W., McSween, H.Y., Jr., 2010. Diogenites as polymict breccias composed of coordinates and rotational elements: 2009. Celest. Mech. Dynam. Astron. 109, 101-135. ('&! (''! ('(! Binzel, R.P., Gaffey, M.J., Thomas, P.T., Zellner, B.H., Storrs, A.D., Wells, E.N., 1997. orthopyroxenite and harzburgite. Meteorit. Planet. Sci. 45, 5, 850-872. (')! Geologic mapping of Vesta from 1994 Hubble Space Telescope images. Icarus 128, 95- ('*! ((+! (("! Bobrovnikoff, N.T., 1929. The spectra of minor planets. Lick Obs. Bull. 407, 18-27. 103. ((#! (($! Buchanan, P.C., Zolensky, M.E., Reid, A.M., 1993. Carbonaceous chondrite clasts in the ((%! ((&! howardites Bholghati and EET87513. Meteoritics 28, 5, 659-669. ! $#! (('! Carry, B., Vernazza, P., Dumas, C., Fulchignoni, M., 2010. First disk-resolved (((! (()! ((*! Chamberlain, M.A., Lovell, A.J., Sykes, M.V., 2007. Submillimeter lightcurves of Vesta. spectroscopy of (4) Vesta. Icarus 205, 473-482. ()+! ()"! ()#! Chang, Y.C., and Chang, C.-S., 1962. Photometric investigations of seven variable Icarus 192, 448. ()$! ()%! ()&! Cloutis, E.A., Gaffey, M.J., Jackowski, T.L., Reed, K.L., 1986. Calibrations of phase asteroids. Acta Astron. Sin. 10, 101-111. ()'! ()(! ())! abundance, composition and particle size distribution for olivine-orthopyroxene mixtures ()*! Cuffey, J., 1953. Pallas, Vesta, Ceres and Victoria; photoelectric photometry. Astron. J. from reflectance spectra. J. Geophys. Res. 91, 11641-11653. (*+! (*"! (*#! Degewii, J., 1978. Photometry of faint asteroids and satellites. Ph.D. thesis, Leiden 58, 212. (*$! Observatory. (*%! (*&! Degewij, J., Tedesco, E.F., Zellner, B., 1979. Albedo and color contrasts on asteroid (*'! (*(! (*)! De Sanctis, M.C., et al., 2012a. Spectroscopic characterization of mineralogy and its surfaces. Icarus 40, 364-374. (**! )++! )+"! De Sanctis, M.C., et al., 2012b. Detection of Widespread Hydrated Materials on Vesta by diversity across Vesta. Science 336, 697-700. )+#! )+$! )+%! Drummond, J., Christou, J., 2008. Triaxial ellipsoid dimensions and rotational poles of the VIR Imaging Spectrometer On Board the Dawn Mission. Ap. J. Lett. 758. )+&! )+'! )+(! seven asteroids from Lick Observatory adaptive optics images, and of Ceres. Icarus 197, )+)! Drummond, J., Eckart, A., Hege, E.K., 1988. Speckle interferometry of asteroids. Icarus 480-496. )+*! )"+! )""! Drummond, J.D., Fugate, R.Q., Christou, J.C., 1998. Full adaptive optics images of 73, 1-14. )"#! )"$! )"%! asteroids Ceres and Vesta rotational poles and triaxial ellipsoid dimensions. Icarus 132, )"&! Gaffey, M. J., 1997. Surface lithologic heterogeneity of asteroid 4 Vesta. Icarus 127, 130- 80-99. )"'! )"(! )")! Gaffey, M.J., 2010. Space weathering and the interpretation of asteroid reflectance 157. )"*! )#+! )#"! spectra. Icarus 209, 564-574. ! $$! )##! )#$! Gaskell, R. W. and 15 colleagues, 2008. Characterizing and navigating small bodies with )#%! Gaskell, R. W., 2011, Optical Navigation near small bodies AAS paper 11-220, imaging data, Meteoritics and Planetary Science 43:1049-1062. )#&! AAS/AIAA Space Flight Mechanics Meeting, New Orleans, LA, 2011. )#'! )#(! Gehrels, T., 1967. Minor planets. I. The rotation of Vesta. Astro. J. 72, 929-938. )#)! )#*! Groeneveld, I., Kuiper, G.P., 1954. Photometric studies of asteroids. I. Astrophys. J. 120, )$+! )$"! )$#! Hasegawa, S., Murakawa, K., Ishiguro, M., Nonaka, H., Takato, N., Davis, C.J., Ueno, 200-220. )$$! M., Hiroi, T., 2003. Evidence of hydrated and/or hydroxylated minerals on the surface of )$%! )$&! )$'! Haupt, H., 1958. Photoelektrisch-photometrische studies an Vesta. Mitt. Sonnenobs. asteroid 4 Vesta. Geophys. Res. Lett. 30, …… )$(! Kanzelhohe 14, 303. )$)! )$*! Herrin, J.S., Zolensky, M.E., Cartwright, J.A., Mittlefehldt, D.W., Ross, D.K., 2011. )%+! Carbonaceous chondrite-rich howardites, the potential for hydrous lithologies on the )%"! HED parent. Lunar Planet. Sci. XLII, 2806 (abstract#1608). )%#! )%$! Hiroi, T., Pieters, C.M., 1998. Origin of vestoids suggested from the space weathering )%%! )%&! Res. 11, 163-170. )%'! trend in the visible reflectance spectra of HED meteorites and lunar soils. Antarct. Meteor. )%(! )%)! )%*! Le Corre, L., Reddy, V., Nathues, A., Cloutis, E.A., 2011. How to characterize terrains )&+! on 4 Vesta using Dawn Framing Camera color bands? Icarus 216, 2, 376-386. )&"! Vesta from Dawn Framing Camera. Meteorit. Planet. Sci. Supplement 5125 (abstract). )&#! Le Corre, L., et al., 2012. Nature of orange ejecta around Oppia and Octavia craters on )&$! )&%! with Hubble Space Telescope. Icarus 208, 238-251. )&&! Li, J.-Y., et al., 2010. Photometric mapping of Asteroid (4) Vesta’s southern hemisphere )&'! )&(! )&)! Li, J.-Y., et al., 2011. Improved measurement of Asteroid (4) Vesta’s rotational axis )&*! orientation. Icarus 211, 528-534. )'+! )'"! Li, J.-Y., et al., 2012. Investigating the origin of bright materials on Vesta: synthesis, )'#! conclusions, and implications. Lunar Planet. Sci. XLIII, 1659 (abstract). )'$! )'%! Loeffler, M.J., Baragiola, R.A., Murayama, M., 2008. Laboratory simulations of )'&! Magnusson, P., 1986. Distribution of spin axes and senses of rotation for 20 large redeposition of impact ejecta on mineral surfaces. Icarus 196, 285–292. )''! )'(! asteroids. Icarus, 68, 1-39. ! $%! )')! McCord, et al., 2012. Dark material on Vesta: adding carbonaceous volatile-rich )'*! materials to planetary surfaces. Nature 491, 83-86. )(+! )("! Murchie, S., et al., 2002. Color variations on Eros from NEAR multispectral imaging. )(#! )($! )(%! Noble, S.K., Keller, L.P., and Pieters, C.M., 2011. Evidence of space weathering in Icarus 155, 145-168. )(&! )('! )((! regolith breccias II: Asteroidal regolith breccias. Meteorit. Planet. Sci. 45, 2007-2015. )()! )(*! Pieters, C.M., et al., 2000. Space weathering on airless bodies: Resolving a mystery with ))+! lunar samples. Meteorit. Planet. Sci. 35, 1101-1107 ))"! ))#! Pieters, C.M., et al., 2012. The distinctive space weathering on Vesta. Nature 491, 79-82. ))$! ))%! Prettyman, T.H., et al.. 2012. Elemental mapping by Dawn reveals exogenic H in Vesta's ))&! Reddy, V., Gaffey, M.J., Kelley, M.S., Nathues, A., Li, J.-Y., Yarbrough, R., 2010. howarditic regolith. Science 338, 6104, 242-246. ))'! Compositional heterogeneity of Asteroid 4 Vesta’s Southern Hemisphere: Implications ))(! )))! ))*! Reddy, V., et al.,, 2012a. Albedo and color heterogeneity of Vesta from Dawn. Science for the Dawn Mission. Icarus 210, 2, 693-706. )*+! )*"! )*#! Reddy, V., et al., 2012b. Delivery of dark material to Vesta via carbonaceous chondritic 336, 700-704. )*$! )*%! )*&! Reddy, V., et al., 2012c. Photometric, spectral phase and temperature effects on Vesta impacts. Icarus 221, 2, 544-559. )*'! )*(! )*)! Rivkin, A.S., McFadden, L.A., Binzel, R.P., Sykes, M., 2006. Rotationally-resolved and HED meteorites: Implications for Dawn mission. Icarus 217, 153-168. )**! *++! *+"! Russell, C.T., et al., 2012. Dawn at Vesta: Testing the protoplanetary paradigm. Science spectroscopy of Vesta I: 2-4 lm region. Icarus 180, 2,, 464-472. *+#! *+$! *+%! 336, 684-686. *+&! *+'! weathering and taxonomic classification. Icarus 220, 36-50. Sanchez, J.A., Reddy, V., Nathues, A., Cloutis, E.A., Mann, P., Hiesinger, H., 2012. *+(! Phase reddening on near-Earth asteroids: Implications for mineralogical analysis, space *+)! *+*! *"+! Schenk, P., et al., 2012. The geologically recent giant impact basins at Vesta’s south pole. *""! Science 336, 6082, 694-696. ! $&! *"#! *"$! Dynam. Astron. 82, 83-110. Seidelmann, P.K., et al., 2002. Report of the IAU/IAG Working Group on cartographic *"%! coordinates and rotational elements of the planets and satellites: 2000. Celest. Mech. *"&! *"'! *"(! Seidelmann, P.K., et al., 2005. Report of the IAU/IAG Working Group on cartographic *")! coordinates and rotational elements: 2003. Celest. Mech. Dynam. Astron. 91, 203-215. *"*! *#+! Seidelmann, P.K., et al., 2007. Report of the IAU/IAG Working Group on cartographic *#"! coordinates and rotational elements: 2006. Celest. Mech. Dynam. Astron. 98, 155-180. *##! *#$! Sierks, H., et al., 2011. The Dawn Framing Camera. Space Sci. Rev. 163, 263-327. *#%! *#&! Taylor, R.C., 1973. Minor planets and related objects. XIV. asteroid (4) Vesta. Astron. J. *#'! 78, 1131-1139. *#(! *#)! Taylor, R.C., Tapia, S., Tedesco, E.F., 1985. The rotation period and pole orientation of *#*! asteroid 4 Vesta. Icarus 62, 298-304. *$+! Wentworth, S., 2001. The effects of space weathering on Apollo 17 mare soils: *$"! Taylor, L.A., Pieters, C.M., Keller, L.P., Morris, R.V., McKay, D.S., Patchen, A., *$#! *$$! Petrographic and chemical characterization. Meteorit. Planet. Sci. 36, 285-300. *$%! Vesta: Spin pole, size, and shape from HST images. Icarus 128, 88-94. *$&! Thomas, P.C., Binzel, R.P., Gaffey, M.J., Zellner, B.H., Storrs, A.D., Wells, E., 1997a. *$'! *$(! *$)! Thomas, P.C., et al., 1997b. Impact excavation on asteroid 4 Vesta: Hubble Space *$*! Vernazza, P., Mothé-Diniz, T., Barucci, M.A., Birlan, M., Carvano, J.M., Strazzulla, G., Telescope results. Science 277, 1492-1495. *%+! *%"! *%#! Fulchignoni, M., Migliorini, A., 2005. Analysis of near-IR spectra of 1 Ceres and 4 Vesta, *%$! Vernazza, P., Brunetto, R., Strazzulla, G., Fulchignoni, M., Rochette, P., Meyer-Vernet, targets of the Dawn mission. Astron. Astrophys. 436, 3, 1113-1121. *%%! N., Zouganelis, I., 2006. Asteroid colors: a novel tool for magnetic field detection? The *%&! *%'! *%(! Veverka, J., et al., 1996. Ida and Dactyl: Spectral reflectance and color variations. Icarus case of Vesta. Astron. Astrophys. 451, L43-L46. *%)! *%*! *&+! 120, 66-76. *&"! *&#! Zellner, B.H., et al., 1997. Hubble Space Telescope images of asteroid 4 Vesta in 1994. *&$! Icarus 128, 83-87. *&%! *&&! Zellner, N.E.B., Gibbard, S., de Pater, I., Marchis, F., Gaffey, M.J., 2005. Near-IR *&'! imaging of asteroid 4 Vesta. Icarus 177, 190-195. ! $'! *&(! *&)! Zolensky, M.E., Weisberg, M.K., Buchanan, P.C., Mittlefehldt, D.W., 1996. Mineralogy *&*! of carbonaceous chondrite clasts in HED achondrites and the Moon. Meteoritics 31, 518- ! *'+! 537. *'"! *'#! *'$! *'%! *'&! *''! *'(! *')! *'*! Table 1. List of FC filters (except clear filter) with their respective band passes width and *(+! *("! *(#! peak. FWHM (!m) Filter name Wavelength center (!m) F8 0.438 0.040 0.043 0.555 F2 0.042 0.653 F7 0.044 0.749 F3 F6 0.829 0.036 0.045 0.917 F4 *($! F5 0.965 0.086 *(%! Table 2. Observational circumstances for Dawn data collected at Vesta. *(&! *('! Best Resolution Sub-Spacecraft Distance To Orbital Phase Vesta (km) Latitude (meters/pixel) 100,000 -32° 9067 RC1* RC2 3382 -54° 37,000 5,200 -25° 487 RC3 5,200 487 RC3B -25° 2,700 50° to -90° 252 Survey HAMO** (1) 61 66° to -87° 660 to 730 190 to 240 85° to -90° 16 LAMO*** HAMO (2) 60 85° to -85° 640 to 730 *Rotational Characterization **High Altitude Mapping Orbit ***Low Altitude Mapping Orbit ! *((! *()! *(*! *)+! *)"! *)#! *)$! $(! *)%! *)&! Regio, longitude of Claudia and long axis longitude in pre-Dawn coordinate systems (I to Table 3. Values for right ascension, declination of the spin pole, ephemeris position of *)'! VI) and Claudia coordinate system (VII). Obliquity is the angular distance between the the prime meridian, obliquity, right ascension of vernal equinox, longitude of Olbers *)(! *))! *)*! orbital and rotational poles and vernal equinox is the descending intersection of Vesta’s **+! orbital and its equatorial planes. Note that the IIb values were later used by Seidelmann et ! **"! al., 2001, Seidelmann et al., 2005 and Seidelmann et al., 2007. **#! **$! **%! **&! **'! **(! **)! ***! "+++! "++"! "++#! "++$! "++%! "++&! "++'! "++(! "++)! "++*! "+"+! "+""! "+"#! ! $)! Table 4. The observing geometries of Dawn FC data and HST 2007 data we used to compare the lightcurves of Vesta. The last column lists the difference between sub-solar longitude and sub-observer longitude. A positive value indicates that the morning side is imaged, while a negative value indicates that the afternoon side is imaged. "+"$! "+"%! "+"&! "+"'! "+"(! Lat_sun – Phase Sub-observer Sub-Solar Date Data Set Lat_obs angle latitude Latitude 29.1º 26.1º -32.2º -25.7º 2011-06-30 Dawn FC "+")! Table 5. List of albedo and spectral units identified in HST maps of Vesta and their ! 2007-05-14/16 HST 2007 -5.3º -13.4º 9.5º -5.4º "+"*! "+#+! "+#"! corresponding IAU approved names from Dawn FC data. A complete list of named "+##! features on Vesta from the Dawn mission is available at "+#$! ! http://planetarynames.wr.usgs.gov/Page/VESTA/target "+#%! ! HST Albedo/Spectral Unit “Olbers” #13 #11 Q Z #15 #3 Y K A B #8 #6 #4 #5 #7 "+#&! "+#'! ! ! ! IAU Approved Feature from Dawn FC Map Dark ejecta E of Marcia, Calpurnia, Minucia Eusebia Vibidia Numisia Teia Oppia Canuleia, Justinia area Feralia Marcia Octavia ejecta Octavia Lucaria Tholus Pinaria area Rubria, Occia area Aquilia Matronalia Rupes Type of Feature Coordinates (Claudia System) Latitude Longitude Albedo unit -10° to +20° 197° to 215° -42.2° -26.9° -7.0° -3.4° -8° -33.7° to -34.4° +4° +10° -3.3° -13° -29° -7.4° to -15.4° -49.7° -49.5° 204.2° 220.1° 247° 271° 309° 294.5° to 317.9 312° 190° 147° 104° 32° 18.4° 41° 82.7° Crater Crater Crater Crater Crater Craters Planitia Crater Ejecta Crater Hill Crater Craters Crater Scarp $*! by HST in May 2007. The observing geometries are listed in Table 3. In all three panels, red and orange symbols are measured from Dawn FC images, and green symbols are from HST images. Dawn FC observations and HST observations are arbitrarily shifted to align with each other in magnitude at each wavelength. The top panel shows lightcurves Figure 2. Rotational lightcurves as observed by Dawn FC during approach to Vesta and "+#(! !"#$%&'()*+",-.' "+#)! "+#*! Claudia’s location southeast of merged craters and a detailed view of Claudia crater itself Figure 1. Location of Claudia crater in Vesta’s western hemisphere with a close up (b) of "+$+! "+$"! (c). "+$#! "+$$! "+$%! "+$&! "+$'! "+$(! "+$)! "+$*! "+%+! "+%"! "+%#! "+%$! "+%%! "+%&! "+%'! "+%(! "+%)! Figure 3. (A) Global topography of Vesta from Hubble imaging data (Thomas, et al, from Dawn FC clear filter (F1, effective wavelength 699 nm) in red and F7 (652 nm) in the absolute scale ranges from -12 to +12 km in HST data and -22 to +19 km in Dawn 1997) and from Dawn Framing Camera (B) in IAU ‘Olbers’ coordinate system. Although data, Vesta’s relative topography from HST is consistent with Dawn. orange, and from HST WFPC F673N filter (673 nm); middle panel from Dawn FC F8 filter (438 nm) and HST F439W filter (431 nm); and lower panel Dawn FC F5 filter (961 nm) and HST F953N filter (953 nm). All lightcurves are plotted with respect to sub-solar longitude in IAU ‘Olbers’ coordinate system. ! %+! the Dawn science team in all their publications. showing the intensity of the pyroxene 0.90-µm absorption feature in the Thomas et al. Figure 5. Band depth ratio (0.75/0.92 µm) map of Vesta from HST (A) and Dawn (B) "+%*! Figure 4. (A) HST map of Vesta in 0.673-!m filter projected in the Thomas et al. "+&+! (1997a) coordinate system based on observations from 1994, 1996 and 2007 oppositions "+&"! RC1 at a resolution of 9.06 km/pixel with prime meridian similar to Thomas et al. at a resolution of ~50 km/pixel. (B) Dawn FC map of Vesta in 0.75-micron filter from "+&#! "+&$! (1997a) coordinate system in Fig. 4a. The pole position ("0 = 309.03° ± 0.01°, 42.23° ± "+&%! map are slightly rotated with respect to HST map. Fig. 4c is similar to Fig. 4b but the 0.01°) is the updated one from Russell et al. (2012) and hence features in the Dawn FC "+&&! "+&'! prime meridian is defined in the Claudia coordinate system (Russell et al. 2012) used by "+&(! "+&)! "+&*! "+'+! "+'"! "+'#! "+'$! "+'%! "+'&! "+''! "+'(! "+')! "+'*! "+(+! "+("! bright features and dark features identified by Binzel et al. (1997) and Li et al. (2010). ! %"! Figure 6. Dawn FC color maps of Vesta using data obtained during the RC3B approach Figure 7. Albedo map of Vesta from HST (A) and Dawn (B) showing corresponding (1997a) coordinate system. Areas in red have deeper pyroxene band and areas in blue have weaker bands. HST and Dawn band depth maps are consistent with each other. phase of the mission in the Thomas et al. (1997a) coordinate system. (A) Albedo map in 0.75 !m filter, (B) Band depth (0.75/0.92 !m) ratio map with red areas depicting deeper 0.90-!m pyroxene band, and (C) eucrite-diogenite ratio (0.98/0.92 !m) map with red areas indicating more diogenitic material and blue areas more eucritic. http://planetarynames.wr.usgs.gov/images/vesta.pdf al. (2010) and band depth ratio map from Dawn. Both maps are in the Thomas et al. Both maps are in the Thomas et al. (1997a) coordinate system. A global map of Vesta obtained by the Dawn FC camera in Claudia coordinate system is available online at Figure 8. Comparison of compositional units observed in HST map of Vesta from Li et "+(#! "+($! "+(%! "+(&! "+('! "+((! "+()! "+(*! "+)+! Figure 9. Comparison of compositional units observed in ground-based map of Vesta "+)"! Dawn. Both maps are in the Thomas et al. (1997a) coordinate system. Note that this map from Gaffey (1997) modified by Binzel et al. (1997) and band depth ratio map from "+)#! "+)$! wrong latitude due to incorrect calculation of sub-Earth latitude. While the longitudinal originally presented in both Gaffey (1997) and modified by Binzel et al. (1997) has "+)%! "+)&! location of the compositional units in ground-based maps is well constrained (limited by "+)'! lightcurve resolution), latitudinal location is weakly constrained as the maps were created "+)(! using disk integrated data. Due to weak constrains on the latitudinal location of "+))! Dawn maps remains valid despite the latitude error. compositional units in Gaffey (1997) and Binzel et al. (1997) maps, comparison with "+)*! "+*+! "+*"! "+*#! "+*$! "+*%! %#! (1997a) coordinate system. ! "+*&! "+*'! "+*(! "+*)! "+**! ""++! ""+"! ""+#! Figure 1. Dawn, HST, Ground-based Studies of Vesta ! %$! ""+$! ""+%! ""+&! ""+'! Figure 2. Dawn, HST, Ground-based Studies of Vesta ! %%! ""+(! ""+)! ""+*! """+! """"! Figure 3. Dawn, HST, Ground-based Studies of Vesta ! %&! """#! """$! """%! Figure 4. Dawn, HST, Ground-based Studies of Vesta ! %'! """&! """'! """(! """)! """*! ""#+! ""#"! ""##! ""#$! Figure 5. Dawn, HST, Ground-based Studies of Vesta ! %(! ""#%! Figure 6. Dawn, HST, Ground-based Studies of Vesta ""#&! ""#'! ! %)! ""#(! ""#)! ""#*! ""$+! ""$"! ""$#! ""$$! ""$%! Figure 7. Dawn, HST, Ground-based Studies of Vesta ! %*! ""$&! ""$'! ""$(! ""$)! ""$*! ""%+! ""%"! Figure 8. Dawn, HST, Ground-based Studies of Vesta ! &+! ""%#! ""%$! ""%%! ""%&! ""%'! ""%(! Figure 9. Dawn, HST, Ground-based Studies of Vesta ! &"!
1709.03118
2
1709
2017-10-26T12:17:12
The Physics of Protoplanetesimal Dust Agglomerates. X. Mechanical properties of dust aggregates probed by a solid-projectile impact
[ "astro-ph.EP", "cond-mat.soft" ]
Dynamic characterization of mechanical properties of dust aggregates has been one of the most important problems to quantitatively discuss the dust growth in protoplanetary disks. We experimentally investigate the dynamic properties of dust aggregates by low-speed ($\lesssim 3.2$ m s$^{-1}$) impacts of solid projectiles. Spherical impactors made of glass, steel, or lead are dropped onto a dust aggregate of packing fraction $\phi=0.35$ under vacuum conditions. The impact results in cratering or fragmentation of the dust aggregate, depending on the impact energy. The crater shape can be approximated by a spherical segment and no ejecta are observed. To understand the underlying physics of impacts into dust aggregates, the motion of the solid projectile is acquired by a high-speed camera. Using the obtained position data of the impactor, we analyze the drag-force law and dynamic pressure induced by the impact. We find that there are two characteristic strengths. One is defined by the ratio between impact energy and crater volume and is $\simeq 120$ kPa. The other strength indicates the fragmentation threshold of dynamic pressure and is $\simeq 10$ kPa. The former characterizes the apparent plastic deformation and is consistent with the drag force responsible for impactor deceleration. The latter corresponds to the dynamic tensile strength to create cracks. Using these results, a simple model for the compaction and fragmentation threshold of dust aggregates is proposed. In addition, the comparison of drag-force laws for dust aggregates and loose granular matter reveals the similarities and differences between the two materials.
astro-ph.EP
astro-ph
Draft version June 24, 2021 Preprint typeset using LATEX style emulateapj v. 01/23/15 7 1 0 2 t c O 6 2 . ] P E h p - o r t s a [ 2 v 8 1 1 3 0 . 9 0 7 1 : v i X r a THE PHYSICS OF PROTOPLANETESIMAL DUST AGGLOMERATES. X. MECHANICAL PROPERTIES OF DUST AGGREGATES PROBED BY A SOLID-PROJECTILE IMPACT Institut fur Geophysik und extraterrestrische Physik, Technische Universitat zu Braunschweig, Mendelssohnstr. 3, D-38106 Department of Earth and Environmental Sciences, Nagoya University, Furocho, Chikusa, Nagoya, Aichi 464-8601, Japan Braunschweig, Germany and Hiroaki Katsuragi Institut fur Geophysik und extraterrestrische Physik, Technische Universitat zu Braunschweig, Mendelssohnstr. 3, D-38106 Jurgen Blum Braunschweig, Germany Draft version June 24, 2021 ABSTRACT Dynamic characterization of mechanical properties of dust aggregates has been one of the most impor- tant problems to quantitatively discuss the dust growth in protoplanetary disks. We experimentally investigate the dynamic properties of dust aggregates by low-speed (. 3.2 m s−1) impacts of solid projectiles. Spherical impactors made of glass, steel, or lead are dropped onto a dust aggregate of packing fraction φ = 0.35 under vacuum conditions. The impact results in cratering or fragmentation of the dust aggregate, depending on the impact energy. The crater shape can be approximated by a spherical segment and no ejecta are observed. To understand the underlying physics of impacts into dust aggregates, the motion of the solid projectile is acquired by a high-speed camera. Using the obtained position data of the impactor, we analyze the drag-force law and dynamic pressure in- duced by the impact. We find that there are two characteristic strengths. One is defined by the ratio between impact energy and crater volume and is ≃ 120 kPa. The other strength indicates the fragmentation threshold of dynamic pressure and is ≃ 10 kPa. The former characterizes the apparent plastic deformation and is consistent with the drag force responsible for impactor deceleration. The latter corresponds to the dynamic tensile strength to create cracks. Using these results, a simple model for the compaction and fragmentation threshold of dust aggregates is proposed. In addition, the comparison of drag-force laws for dust aggregates and loose granular matter reveals the similarities and differences between the two materials. Keywords: methods: laboratory -- planetary system: formation -- planets and satellites: formation 1. INTRODUCTION In protoplanetary disks, dust growth from sub- micrometer-sized monomer grains to at least kilometer- sized planetesimals has to occur in order to initiate planet formation. Once planetesimals have formed, their own gravity enables them to grow towards planets in mutual collisions. However, the scenario of planetesimal forma- tion is not so straightforward. According to Guttler et al. (2010) and Zsom et al. (2010), the growth of dust ag- gregates by mutual adhesive collisions is limited to cen- timeter size, due to bouncing rather than the sticking collisions. Although their model was the first that com- piled the results of numerous laboratory experiments (see Blum & Wurm 2008, for an earlier compilation), it was far from completion. For example, the influence of the size ratio of colliding aggregates (Bukhari Syed et al. 2017) or erosion (Schrapler et al. 2017) have been exper- imentally studied only recently and revealed new aspects of dust-aggregate collisions. In addition to the direct collision experiments, mechan- ical characterizations of bulk dust aggregates have also been performed. Compressive and tensile strengths of dust aggregates have been measured by static labora- tory experiments (Blum & Schrapler 2004; Blum et al. 2006). The projectile-penetration test as well as com- pressive tests has also been performed (Guttler et al. 2009). Tensile strengths of dust aggregates were mea- sured also by the Brazilian-disk test (Meisner et al. 2012). In Meisner et al. (2012), the packing-fraction dependence of the tensile strength and the elastic modulus based on sound-speed measurements were obtained as well. These mechanical properties are requisites for continuum-based numerical modeling of dust-aggregate collisions (Sirono 2004; Guttler et al. 2009). Furthermore, these mechan- ical properties provide guidelines to our understanding what happens in various complex phenomena caused by collisions of dust aggregates. Typical values of tensile strengths obtained by the above-mentioned experiments are on the order of 1 kPa. This strength value is actually close to that for cometary meteoroids (Trigo-Rodriguez & Llorca 2006). Be- sides, the typical packing-fraction value estimated from the bulk density of cometary nuclei is approximately 0.4 (Blum et al. 2006). This value is also in the range of typical packing-fraction values of dust aggregates used in laboratory experiments. Therefore, model studies with dust aggregates could be useful to discuss physical phe- nomena occurring on comets. To model cometary pro- cesses by using dust aggregates, proper understanding of their mechanical properties is necessary. Although wet granular matter, which can be a model for cohe- sive grains, shows similar tensile strengths on the or- der of 1 kPa, its constituent-grain size is usually macro- scopic (Scheel et al. 2008; Herminghaus 2013), with typ- 2 Katsuragi & Blum ical values around millimeter scale. Therefore, as long as we consider comets as more or less pristine objects (Fulle & Blum 2017), dust aggregates can be regarded as suit- able analogues of pristine materials to study cometary processes. Namely, the mechanical characterization of dust aggregates provides helpful information for both, planetesimal-formation and cometary-surface processes. Since the dust aggregates used in experiments usu- ally consist of micrometer-sized solid (SiO2) monomers, they can be regarded as a certain class of granular mat- ter. Granular matter is defined as a collection of dissi- pative solid grains. The main difference between dust aggregates and granular matter is the role of gravity. In macroscopic granular systems, gravity (body force) plays an essential role. However, monomers in dust ag- gregates are too small and too cohesive to be affected by gravity. Due to their mutual van der Waals attrac- tion, dust aggregates can sustain their bulk shape even for small packing fractions. Contrastively, it is difficult to produce aggregates with small packing fractions us- ing macroscopic granular matter. Under small packing- fraction conditions, macroscopic granular matter exhibits the gaseous rather than the solid state, even under mi- crogravity conditions. Thus, dust aggregates are intrinsi- cally different from conventional granular matter. What is the difference between granular matter and dust ag- gregates in terms of mechanical properties? To answer this question, impact tests might be useful. Impact drag-force modeling has been used to charac- terize mechanical properties of granular matter (Kat- suragi 2016). For instance, the equation of motion of a solid projectile impacting onto a granular matter, m d2z dt2 = mg − kz − m v2 d1 , (1) has been established based on experimental results (Kat- suragi & Durian 2007). Here, m, t, z, and v are the mass of the projectile, the elapsed time from the im- pact moment, the instantaneous penetration depth, and the instantaneous velocity of the projectile, respectively; g represents the gravitational acceleration. The relation between the two parameters k, d1 and the system proper- ties (e.g., projectile density, projectile size, and granular friction) has also been experimentally revealed by Kat- suragi & Durian (2013). For granular impacts, the depth- dependent term kz in Eq. (1) comes from the hydrostatic pressure of the granular target. For dust-aggregate im- pacts, the same form can be applied as long as the pene- tration depth is shallow (see Sect. 3.5). The term mv2/d1 relates to the momentum transfer in both cases. Note that the vertically downward direction corresponds to the positive direction of z in Eq. (1). Actually, the granular impact dynamics depends on various conditions, such as gravity (Nakamura et al. 2013; Altshuler et al. 2014) and packing fraction of the granular matter (Umbanhowar & Goldman 2010; Royer et al. 2011). Equation (1) has to be improved to cover all these broader conditions. However, the accessible range of packing fractions for macroscopic granular matter is very limited as mentioned above. The impact drag force for very small packing fractions has not yet been studied well. By measuring the impact drag force of dust aggregates with small packing fraction and comparing that result with the granular impact case, we can further characterize the mechanical properties of both dust aggregates and granular matter. In this study, we are going to examine the mechan- ical properties of dust aggregates by impact tests with solid projectiles. Specifically, the maximum penetra- tion depth and the dynamic pressure during each im- pact event are computed and analyzed to characterize the dynamic strength of the dust aggregates. In partic- ular, we discuss two kinds of strengths that characterize the mechanical properties of dust aggregates in the dy- namic regime. Using our experimental results, a simple model to compute the compaction induced by the impact and the fragmentation threshold will be derived. At the same time, the drag force model of Eq. (1) is used to explain the deceleration dynamics of penetrating projec- tiles. Then, similarities and differences between macro- scopic granular matter and dust aggregates are discussed on the basis of our experimental data. 2. EXPERIMENT Our experimental setup is very simple and is shown in Fig. 1(a). A cylindrical dust aggregate of packing fraction φ = 0.35, diameter 20 mm, and height 20 mm is prepared as a target (see Blum et al. 2014, for a detailed description of the preparation method). The cylindrical sample is made by using a mold and piston. A fixed amount of monodisperse spherical SiO2 grains of radius 0.76 µm is poured into a mold of 20 mm in diameter. Then, it is compressed by a piston to 20 mm in height and φ = 0.35. We employ monodisperse spherical monomers since they are ideally suited for understanding the phys- ical mechanism and future comparison to particle codes. Although the exact mechanical properties of dust aggre- gates depend on the shape and size distribution of the constituent grains, their qualitative behavior is similar between aggregates consisting of monodisperse spherical grains and polydisperse irregular grains. The dependen- cies of mechanical and morphological properties of dust aggregates on monomer morphologies are basically less than one order of magnitude (Blum et al. 2006; Bertini et al. 2009). Further details on the material properties of monomers and dust aggregates can be found in Blum & Schrapler (2004); Blum et al. (2006). Then, the solid projectile is dropped onto the dust- aggregate target in a vacuum chamber in the laboratory, i.e., under the influence of gravity. The residual pressure in the vacuum chamber is kept at 0.10 Pa. We employ three kinds of projectiles made of glass, steel, and lead. A thin thread is glued on the top of the projectile to be held by a release mechanism, which guarantees a re- producible free fall with zero initial velocity. Diameter D and density ρp (including thread-mass effect) of the projectiles are (D, ρp)=(4.0 mm, 2.6 × 103 kg m−3) for glass, (4.0 mm, 7.7 × 103 kg m−3) for steel, and (4.5 mm, 11 × 103 kg m−3) for lead. The impact velocity ranges from v0 = 0.19 m s−1 to v0 = 3.2 m s−1 and is controlled by the free-fall height of the projectile. In this study, 21 impact experiments (7 with glass, 8 with steel, and 6 with lead) were carried out with various impact veloci- ties. The impact kinetic energy E = mv2 0/2 ranges from E = 3.6×10−6 J to E = 1.1×10−3 J. The impact of the projectile is recorded by a high- speed camera (Photron SA-5) using a macro lens. The Mechanical properties of dust aggregates probed by a solid-projectile impact 3 (cid:9)(cid:66)(cid:10) (cid:84)(cid:80)(cid:77)(cid:74)(cid:69)(cid:1)(cid:81)(cid:83)(cid:80)(cid:75)(cid:70)(cid:68)(cid:85)(cid:74)(cid:77)(cid:70) (cid:88)(cid:74)(cid:85)(cid:73)(cid:1)(cid:85)(cid:73)(cid:83)(cid:70)(cid:66)(cid:69) (cid:9)(cid:67)(cid:10) (cid:91) (cid:68)(cid:66)(cid:78)(cid:70)(cid:83)(cid:66) (cid:68)(cid:66) (cid:69)(cid:86)(cid:84)(cid:85)(cid:1)(cid:66)(cid:72)(cid:72)(cid:83)(cid:70)(cid:72)(cid:66)(cid:85)(cid:70) (cid:9)(cid:68)(cid:10) (cid:9)(cid:69)(cid:10) (cid:9)(cid:70)(cid:10) Figure 1. Experimental setup and raw data. (a) A solid projectile hung by a thread is released from a pre-deterimed height onto a dust aggregate in a vacuum chamber. (b) Side view images of the impact of a glass sphere of diameter D = 4.0 mm with v0 = 2.0 m s−1. (c) Top view of the crater created by the glass-sphere impact shown in (b). (d) Fragmentation caused by an impact of a steel sphere of D = 4.0 mm with v0 = 2.6 m s−1. (e) Comparison between the measured crater cross section (solid curve) and the spherical shape of the projectile (dashed curve). The crater is produced by an impact of a steel sphere with D = 4.0 mm diameter and an impact velocity of v0 = 1.7 m s−1. Mind the different scales of the two axes. images of size 512 × 320 pixels with spatial resolution of 20 µm pixel−1 are acquired with a rate of 42, 000 frames per second. Trimmed example images of the impacting projectile taken by the high-speed camera are shown in Fig. 1(b). The position of the projectile is determined by image analysis. The image correlation of the upper hemisphere of the projectile is used to precisely identify the position of the projectile. Using the kinematic data obtained by the image analysis, we derive the mechanical properties of the target dust aggregate and drag-force law experienced by the impinging solid projectile, as will be shown below. 3. RESULTS AND ANALYSES 3.1. Cratering and fragmentation When the kinetic energy of the impactor is small, only a shallow crater is formed on the target surface. An example of a crater is shown in Fig. 1(c). To evaluate the crater morphology, laser profilometry is used, like in de Vet & de Bruyn (2007). However, the accuracy of our simple system is not sufficient to derive the de- tailed properties of the crater, in particular its depth. Thus, we adopt the laser-profilometry method only to qualitatively confirm that the crater shape can be ap- proximated by a spherical segment with the projectile's diameter. The comparison between the measured crater shape and the spherical shape of the projectile is shown in Fig. 1(e). The good agreement of two profiles means that the crater is just an imprint of the impact. Nei- ther excavation of target material nor ejecta splashing by the impact was observed. The crater is formed solely by the compression of the target dust aggregate. This type of cratering has not been found in other low-speed granular impact studies (see Walsh et al. 2003; Katsuragi 2010; Pacheco-V´azquez & Ruiz-Su´arez 2011). When the kinetic energy of the impact exceeds a certain threshold value, fragmentation of the target dust aggregate occurs. An example of a fragmented dust aggregate is shown in Fig. 1(d). 3.2. Kinematics of projectile motion Example data of the projectile motion during impact obtained by image analysis of the raw video data are shown in Fig. 2. The left column exhibits a high-speed impact (v0 = 3.2 m s−1), whereas the right column dis- plays a low-speed case (v0 = 0.29 m s−1). A glass-sphere projectile of D = 4.0 mm is used in both cases. In the top row (Fig. 2(a,e)), the penetration depth z, which is directly measured from the image data, is shown as a function of time t. z = 0 and t = 0 correspond to the initial surface level of the dust aggregate and the im- pact moment, respectively. In the insets of Fig. 2(a,e), the penetrating regimes are magnified so that individual data points can be discerned. The impact moment is identified by using the ve- locity data v(t) shown in the second row, Fig. 2(b,f). The velocity data are computed by differentiating z(t). Then, free-fall lines are fitted to the obviously free-falling regime. The free-fall portions of the trajectory with slopes g = 9.8 m s−2 are shown as red dashed lines in Fig. 2(b,f). The moment at which the velocity curve deviates from the free-fall line is defined as the impact moment (t = 0). As shown in Fig. 2(b,f), the velocity data exhibit a very sudden decrease after the impact. Thus, it is not difficult to identify the impact moment. The measurement uncertainty in this experiment is esti- mated by the uncertainty of each free-fall fitting. In the later analyses, this fitting uncertainty is used to calcu- late the corresponding errors with the error propagation method. In many impact events, a rebound of the projectile is observed. The rebound moment and its velocity can also be identified by a free-fall fitting. The rebound velocity is defined by the velocity at which the projectile motion follows the free-fall line again. The rebounding free-fall lines are shown as blue dashed lines in Fig. 2(b,f). 4 Katsuragi & Blum The straight line in Fig. 3 shows the power-law relation zmax∝H 1/3. While the data show considerable scatter- ing, they roughly agree with the power-one-third scaling. A similar power-law relation was found in experiments in which solid projectiles impacted into loose granular mat- ter (Uehara et al. 2003; Ambroso et al. 2005). Actually, this power-law relation simply implies that the penetra- tion is in the strength regime (Katsuragi 2016), which is atypical for coarse granular matter, but typical for cohesion-dominated material. Figure 2. Example data of projectile motion during impact. The left and right column corresponds to glass-sphere impacts with im- pact velocities of v0 = 3.2 m s−1 and v0 = 0.29 m s−1, respectively. From top to bottom, penetration depth z, velocity v, acceleration a, and dynamic pressure pdy are shown as a function of time t. In the insets of (a,e), z(t) data during the penetration are magnified to display the individual data point. Units in the insets are iden- tical to those used in the main plots. As normalization, we chose t = 0 and z = 0 for the impact moment and target surface, re- spectively. Although the difference of impact velocity between the left and right case is more than one order of magnitude, the pen- etration dynamics looks similar. Mind the units of km s−2 in the acceleration data. By differentiating v(t) once more, the acceleration a(t) can be obtained (Fig. 2(c,g)). The acceleration data are noisy, because the second derivative of the z(t) data was used without any smoothing. Note that the units of the vertical axes in Fig. 2(c,g) are km s−2. Thus, the fluc- tuation of a(t) is much larger than g. However, we can still confirm a clear peak in the deceleration curves. Although the impact velocities vary by one order of magnitude between the left and right columns in Fig. 2, the qualitative behavior of both impacts is quite similar. In both cases, a shallow cratering and rebound can be observed. Using steel and lead projectiles, fragmenta- tion of dust aggregate was observed in 4 impacts, while the remaining 17 impacts resulted in the shallow crater- ing. In the fragmentation cases, the breaking target and deep penetration of the projectile prevented a reliable measurement by image analysis. Therefore, we mainly focus on the cratering regime in the following analyses. 3.3. Penetration depth scaling From the peak value of z(t), we can extract the max- imum penetration depth. The maximum penetration depth corresponds to the crater depth since the crater is merely an imprint of the penetrating projectile as men- tioned in Sect. 3.1 (see Fig. 1(e)). The relation between the maximum penetration depth zmax and the total drop distance of the projectile H = h + zmax, where h is the free-fall height computed from the impact velocity through h = v2 0/2g, is plotted in log-log style in Fig. 3. Figure 3. Scaling relation between the maximum penetration depth zmax and the total drop distance H = h + zmax, where h is the free-fall height. The relation zmax∝H 1/3 (straight line) can roughly be confirmed. Error bars are computed by error prop- agation from the uncertainties of the free-fall fitting. The scaling relation shown in Fig. 3 also implies that the impact energy governs the impact dynamics, since H relates to the released potential energy. This tendency is different from the previous study by Guttler et al. (2009), in which the impact between a glass sphere and a very porous dust aggregate (φ = 0.15) was experimentally investigated. Guttler et al. (2009) found that the pene- tration depth can be scaled rather by momentum than by energy. In general, the momentum-dominant scaling explains the dissipative situation better. However, the current experimental result can be explained by energy scaling. In Guttler et al. (2009), deep penetration of the projectile was observed. On the contrary, shallow pene- tration followed by rebound is the dominant outcome in our experiment. While both experiments are dissipative, the degree of dissipation might be different, which could result in the different scaling relations. 3.4. Dynamic pressure Here, we compute another quantity to characterize the dynamic properties of dust aggregates. To derive the dy- namic strength of the dust aggregates, we estimated the dynamic pressure pdy. For shallow penetrations, the ef- fect of contact between projectile and target must be taken into account to compute the instantaneous dy- namic pressure. As mentioned in Sect. 3.3, the penetration dynamics is governed by the impact energy. Thus, let us begin with the energy conservation, i.e., ρpVpamax(cid:18) V A(cid:19)max = 1 2 ρpVpv2 0, (2) Mechanical properties of dust aggregates probed by a solid-projectile impact 5 where Vp is the projectile volume and amax is the maxi- mum deceleration; amax = max(a). V and A are the in- stantaneous penetration (crater) volume V = π(Dz2/2 − z3/3) and the contacting cross-sectional area projected onto the initial target surface A = πz(D − 2z), respec- tively. Thus, (V /A)max denotes the maximum value of the effective penetration depth. By assuming a spherical crater shape in shallow penetration, (V /A)max can be ap- proximated as zmax/2 by neglecting higher order terms. Thus, zmaxamax = v2 0, (3) is obtained. This simple kinematic relation can be checked directly against the experimental data and is shown in Fig. 4. As expected, all data agree very well with Eq. (3), which is free from any fitting parameters. Figure 4. Relation between zmaxamax and v2 0. The black line indicates zmaxamax = v2 0 (Eq. (3)). Error bars are computed by the error propagation from the uncertainties of the free-fall fitting. To evaluate the dynamic pressure caused by the im- pact, we have to estimate the volume supporting the ap- plied pressure. Actually, Guttler et al. (2009) performed similar penetration experiments to ours and measured the compressed zone in the target dust aggregate by us- ing the X-ray micro-CT method. According to their re- sults, the volume of the compressed zone is equivalent to 0.8 - 1.2 of the projectile-sphere volume. Therefore, we assume that the impact-affected volume corresponds to Vp. Then, the impact energy written in Eq. (2) is equiv- alent to max(pdy)Vp. From this assumption and Eqs. (2) and (3), we arrive at max(pdy) = 1 2 ρpzmaxamax = 1 2 ρpv2 0. (4) Equation (4) reveals the relations between the mechan- ical properties, kinematic values, and the initial condi- tions, by means of the dynamic pressure. To compute the instantaneous value of the dynamic pressure, pdy, we extend this relation to every instant by, pdy = 1 2 ρpaz. (5) This relation is geometrically equivalent to pdy = ρpV a/A in shallow penetration. The computed values of pdy are shown in Fig. 2(d,h). Since both a and z vary by about one order of magnitude between the left and right columns in Fig. 2, pdy differs by about two orders of magnitude. 3.5. Drag force characterization To analyze the penetration dynamics, we tried to fit the experimental data to the simple drag-force model written in Eq. (1). Equation (1) has an analytic solution in v-z space (Eq. (2) in Katsuragi & Durian 2013), v2 v2 0 = e− 2z d1 − kd1z mv2 0 +(cid:18) gd1 v2 0 + kd2 1 2mv2 0(cid:19)(cid:16)1 − e− 2z d1(cid:17) . (6) Using this analytic solution, the experimentally obtained v(z) data can be fitted. The fitting results are shown in Fig. 5(a-c). The colored curves are the measured data and the gray dashed curves are the fit results. Here, the two parameters k and d1 were taken as free fit parame- ters. As shown in Fig. 5(a-c), all experimental data can be very well fitted by Eq. (6). Note that the values of k and d1 do not depend on time t, because they are de- termined by fitting the entire shape of each curve shown in Fig. 5(a-c). The values of the two fit parameters k and d1 provide helpful information to discuss the physi- cal situation of the impact. Since the first drag-force term Fk = kz is proportional to the penetration depth z, it relates to the deforma- tion. For the impacts into granular matter, Fk is re- lated to the frictional force by the depth-proportional hydrostatic pressure and Coulomb friction of the gran- ular target. However, the hydrostatic pressure is irrele- vant for impacts into cohesive dust aggregates. Rather, elastic and plastic deformations should determine Fk in this case. Assuming that the impact energy is dissipated by plastic deformation forming the cratering imprint, the deformation-based force Fk = kz should be related to the strength. To compute the relevant strength, Fk must be divided by the contact area. The contact area of a shal- low penetration of a sphere can approximately be written as A≃πzD by neglecting the z2 term. Then, the analogue of the strength can be estimated as Fk/A≃k/(πD). In Fig. 6(a), k/(πD) is plotted as a function of the impact velocity v0. The second fit parameter d1 characterizes the inertial- drag force Fin = m v2/d1, which originates from the momentum transfer between projectile and target. The momentum transfer can be written as ρt v2 D2 dt ∼ ρp D3 dv, where ρt and ρp are the bulk densities of tar- get and projectile, and ρp D3 dv/dt = Fin is the inertial- drag force based on the infinitesimal deceleration dv dur- ing the infinitesimal duration dt. From this momentum balance, a relation d1∼(ρp/(ρt) D) can be derived. To characterize the inertial drag by a dimensionless number corresponding to a drag coefficient, we plot d1ρt/(ρp D) in Fig. 6(b). Here, the contact area is approximated by D2 for the sake of simplicity. Although the actual contact area is proportional to πzD, the factor πz is replaced by a constant characteristic length scale D. Since the Fin- dominant regime is limited to very shallow indentations as discussed later (Fig. 5(d-f)), z can be approximated with a constant. Thus, we replace πz with D in this case. Although D is much greater than πz, there is no other relevant constant length scale in the system. While the specific value of normalized d1 depends on the choice of the length scale, its qualitative behavior is similar as long as the constant length scale is used. In Fig. 6, both quantities k/(πD) and d1 ρt/(ρp D) be- 6 Katsuragi & Blum Figure 5. Top row: instantaneous velocity v as a function of the penetration depth z of (a) glass, (b) steel, and (c) lead projectiles with various impact energies. Colored curves indicate the experimental results while the gray dashed curves show fitting results. Bottom row: ratio between inertial drag Fin and deformation-based drag Fk for each impact ((d) glass, (e) steel, and (f) lead) computed from the fitting results. Except for the very early stage of impact, Fin is much smaller than Fk. appropriately characterize the inertial drag by means of momentum transfer. Moreover, the value of d1 ρt/(ρp D) is similar to that for the granular impact case (Katsuragi & Durian 2007, 2013). Values slightly smaller than unity originate in using the length scale D in the normalization of d1. However, D is the only relevant length scale in the localized impact compression. The ratio of inertial and deformation-based drag forces, Fin/Fk, can also be computed from the fitting result. The computed ratio is shown in Fig. 5(d-f). Obviously, the ratio exceeds unity only in the very early stages of im- pact. In fact, only the first data point shows large Fin/Fk values in all cases. Therefore, we can safely say that the inertial drag is efficient only within the initial 24 µs (= 1/42, 000 s; temporal resolution of the measurement). After the very initial inertial-dominant regime, the dy- namics is mainly governed by Fk. Therefore, Fk must be the relevant parameter to estimate the penetration depth and resultant crater shape. This tendency is consistent with the strength-dominant cratering (penetration) pro- cess if we consider that Fk characterizes plastic deforma- tion. 3.6. Scaling of crater volume, restitution coefficient, and dynamic pressure In the strength regime, the crater volume should be proportional to the kinetic energy E of the impact. As- suming a spherical crater shape, we can compute the crater volume Vc = π(Dz2 max/3). The relation between Vc and E is plotted in Fig. 7(a) and shows a clear linear relation with the slope of unity in the log-log plot. Thus, the data suggest the relation max/2 − z3 Vc = Y −1 k E, (7) where Yk is an effective strength (yield strength) char- acterizing the cratering dynamics. We find that the quality of scaling by Eq. (7) shown in Fig. 7(a) is bet- ter than the scaling by zmax∝H 1/3 shown in Fig. 3. This improvement of scaling quality suggests that the assumption of a spherical crater is reasonable. From the least-square fitting, the strength value is obtained Normalized fit parameter values characterizing Figure 6. (a) deformation-based drag k/(πD) and (b) inertial drag d1 ρt/(ρp D). For relatively high impact velocities (v0 & 0.4 m s−1), these quantities become practically constant. Horizon- tal dashed lines are intended to guide the eye. Error bars for k/(πD) and d1 ρt/(ρp D) indicate the uncertainty of fitting shown in Fig. 5(a-c). Error bars for v0 are estimated from the uncertain- ties of the free-fall fitting. come practically constant for relatively high impact ve- locites (v0 & 0.4 m s−1). The horizontal dashed lines in Fig. 6 are displayed to guide the eye. The saturated con- stant values are k/(πD)≃200 kPa and d1 ρt/(ρp D)≃0.3. The former value is significantly larger than the typical value of the dynamic strength shown in Fig. 2(d,h). The reason for this difference is an important point to charac- terize dust-aggregate impacts as discuss later in Sect. 4.1. The latter value is close to unity. This means that we Mechanical properties of dust aggregates probed by a solid-projectile impact 7 as Yk = 120 kPa. This value agrees well with the drag- based strength value, k/(πD) = 200 kPa (in the satu- rated regime of Fig. 6(a)). function of E. The straight line in Fig. 7(b) indicates a power-law relation of ǫ ∝ E−1/4. (8) This relation actually conflicts with the classical theory for the restitution coefficient associated with plastic de- formation (Johnson 1985). While Johnson (1985) derived the relation ǫ ∝ v−1/4 , the current experimental result suggests ǫ ∝ E−1/4 ∝ v−1/2 , considerably steeper as for classical elastic-plastic materials. The reason for this dis- crepancy is not clear at present. The complex rheological properties of dust aggregates may be responsible for this restitution behavior shown in Fig. 7(b) and described by Eq. (8). 0 0 Finally, we analyze the dynamic pressure pdy. The maximum value of the dynamic pressure max(pdy) could play a crucial role to characterize the mechanical prop- erties of dust aggregates in a dynamic situation. In Fig. 7(c), we plot max(pdy) as a function of E. Since the data suggest a linear dependence between the two values, we can write a relation max(pdy) = V −1 s E, (9) where Vs is a constant having the dimension of a volume. From the least-squares fitting, we get Vs = 4.6×10−8 m3. This volume is much greater than the typical crater vol- ume ≃10−9 m3 and much smaller than the entire target volume ≃6.3×10−6 m3. Rather, Vs is comparable to the typical projectile volume Vp≃3.3×10−8 m3. This rela- tion Vs ≃ Vp is natural since we have assumed that the volume supporting pdy corresponds to Vp (see Sect. 3.4). The threshold for fragmentation can also be derived by considering max(pdy). The estimated max(pdy) val- ues for the fragmentation cases distribute in the top- right corner of Fig. 7(c) bounded by the dashed lines. This implies that, once max(pdy) exceeds a certain threshold value, fragmentation occurs due to macro- scopic crack propagation in the target dust aggregate. The threshold value for our dust-aggregate targets is p∗ dy≃10 kPa (Fig. 7(c)). This value is approximately one order of magnitude less than k/(πD) and Yk. Note that this fragmentation threshold is not necessar- ily identical to the catastrophic disruption limit which is usually characterized by a largest fragment mass smaller than one half of the original mass. Although we do not measure the mass of the largest fragment, it is seemingly not always less than one half of the original mass so that p∗ dy denotes the onset of fragmentation. 4. DISCUSSION 4.1. Physical meaning of the strength values The tensile strength of dust aggregates has been de- termined by considering the stress required to open a crack (Blum et al. 2006). By counting the number of monomer-monomer contacts per unit area in a dust ag- gregate, Blum et al. (2006) obtained a form for the tensile strength Y (Eq. (6) in Blum et al. 2006), Y = 3φFstick 2πs2 0 , (10) where Fstick and s0 are the adhesion force and ra- dius of monomers composing the dust aggregate, respec- Figure 7. Crater volume Vc (a), restitution coefficient ǫ (b), and maximum dynamic pressure max(pdyn) (c) as a function of kinetic impact energy E. (a) The linear relation Vc∝E can be confirmed and the fitting quality is better than in Fig. 3. (b) The restitu- tion coefficient ǫ can be scaled as ǫ ∝ E−1/4. (c) The relation max(pdyn)∝E is held in the cratering regime. Fragmentation of dust aggregates occurs at the highest impact energies (or equiv- alently at the largest dynamic pressures) beyond the dashed line levels. Error bars are computed by error propagation from the uncertainties of the free-fall fitting. If the collision is completely inelastic without any re- bound, these strength values originate only from the pure plastic deformation and all the impact energy is dissi- pated by the cratering. However, we do observe rebound of the projectile in many impacts. For instance, rebounds can clearly be confirmed in Fig. 2. As mentioned in Sect. 3.2, the impact and rebound velocities v0 and v′ can be determined by the data fitting. Then, the resti- tution coefficient ǫ = v′/v0 can be computed. The de- rived values of ǫ as a function of E are shown in Fig. 7(b). In some cases, detection of rebound is impossible due to very small v′ or other limitations of image analysis. All impacts for which we could identify a rebound are plot- ted in Fig. 7(b). As can be seen, ǫ is clearly a decreasing 8 Katsuragi & Blum tively. Inserting typical values of φ = 0.35, Fstick = 67×10−9 N (Heim et al. 1999), and s0 = 0.76 µm (Poppe & Schrapler 2005) into Eq. (10), we obtain Y = 19 kPa. This value is quite close to the fragmentation threshold derived by the dynamic pressure p∗ dy ≃ 10 kPa. This cor- respondence is reasonable, because both quantities Y and p∗ dy characterize the local criterion to open a crack. Once the macroscopic crack is opened in a dust aggregate, it will be unstable in a dynamic state and further propa- gate until the aggregate splits. Therefore, fragmentation takes place when the dynamic pressure exceeds the crit- ical value p∗ dy in localized impacts. However, measured values of the static tensile strength for dust aggregates are less than the above-estimated value (Blum et al. 2006; Meisner et al. 2012). To explain this difference, Blum et al. (2006) have used the energy balance as well, which then yields a much smaller value for the tensile strength. However, the current experimental results suggest that the force balance is essential for estimating the tensile strength in the dynamic impact situation. From the linear relation between max(pdy) and the im- pact kinetic energy E, a volume Vs(≃ Vp) can be esti- mated as written in Eq. (9), which corresponds to the volume that is supporting the applied stress max(pdy). This agreement of Vs and Vp is physically reasonable from the viewpoint of dimensional analysis. If the size of the target is large compared to the projectile size, the rel- evant length scale, which determines the extent of the impact, can only be the size of projectile. Then, the volume of the impact-affected (compressed) zone natu- rally corresponds to the projectile volume. A similarly affected zone can actually be observed in a horizontally- dragged two-dimensional granular layer (Takehara et al. 2010). In our experiments, the target volume is much greater than the projectile volume. As already mentioned, Guttler et al. (2009) reported that the impact-induced compression is limited to a region approximately the vol- ume of the projectile. That is, only the vicinity of the impact point is compressed. In addition, it has been shown that the boundary effect even in granular-impact- cratering experiment is very limited (Nelson et al. 2008). Thus, we consider the impact compression is localized to a volume much smaller than our target and the open- boundary effect is negligible at least in the cratering regime. However, the fragmentation limit certainly de- pends on the size of the target. Systematic experiments like those by Bukhari Syed et al. (2017) are necessary to reveal the size dependence of the fragmentation limit. In this study, we rather focus on the fragmentation limit using identical-size impactors in terms of the maximum dynamic pressure. Systematic studies with target-size variations are interesting future problems. Recently, a periodic propagation of a localized com- paction band has been found in a compressed crush- able granular material (Vald`es et al. 2011; Guillard et al. 2015). The occurrence of a periodic compaction depends on the competition among elastic, local braking, and vis- cous timescales. Thus, if we can observe a similar oscilla- tion with dust aggregates by controlling the compression rate, its analysis might provide further mechanical in- formation. Numerical and experimental studies on the behavior of compressed zones are interesting future top- Strength values obtained or estimated in this study Table 1 Symbol Value (kPa) From Yk k/πD p∗ dy Y 120 200 10 19 Crater shape, Eq. (7) Penetration dynamics, Fig. 6(a) Fragmentation threshold of pdy, Fig. 7(c) Stress to open a crack, Eq. (10) ics. In this study, we also found another type of strength by the analyses of deformation, i.e., Eq. (7) and the drag force fitting to Eq. (6). The two obtained values Yk = 120 kPa and k/(πD) = 200 kPa are close to each other and about one order of magnitude greater than the dynamic pressure threshold p∗ dy = 10 kPa. This is be- cause the former two strengths are computed based only on the apparent deformation, whereas the latter takes into account the compressed zone beneath the crater as well. In other words, one must consider not only the di- rect deformation, like cratering, but also the compression of the surrounding zone to properly model the strength of very soft target materials, such as dust aggregates. However, the projectile motion and the resultant crater shapes are much easier to measure than the internal com- pression caused by the impact. To estimate the impact condition from the resultant crater shape, the effective strength Yk is convenient. As long as the scaling of Eq. (7) is held, the cratering process can be understood effectively by the strength-dominant dynamics governed by the impact energy. Strength values obtained in this study are summarized in Table 1. The relation between max(pdy) and Yk can be de- rived from Eqs. (7), (9) and Vs≃Vp = πD3/6, Yk, and max(pdy) in the following way Yk = 1 6 (cid:18) D zmax(cid:19)2 ρpv2 0 = 1 3 (cid:18) D zmax(cid:19)2 max(pdy). (11) Using this form and Eq. (4), we can caluclate the strength value from the crater depth zmax and the impact velocity v0. To apply these forms, the projectile diameter D and its density ρp must be constant, i.e., the projectile has to be much harder than the target. Then, the value of Yk can be estimated even from a single penetration test. To derive the fragmentation threshold p∗ dy, however, one has to search for the threshold impact velocity v∗ 0 at which the fragmentation begins to occur. Once we obtain v∗ 0 , the strength can be estimated by Eq. (4). To build a model unifying these strength values, sys- tematic studies by experiments and numerical simula- tions are still needed. We have fixed the packing fraction of the dust aggregates in this experiment. In addition, our experiments were performed only under the influence of gravity. Variations of packing fraction and gravita- tional effect are crucial research topics to improve the model. 4.2. Possible applications in astrophysics In our experiments, we used dense solid projectiles, whereas in previous studies, mutual collisions of similar dust aggregates have mainly been investigated to directly Mechanical properties of dust aggregates probed by a solid-projectile impact 9 simulate the planetesimal formation process. Collisions between objects of very different densities in protoplane- tary disks might be not be the rule, they still might hap- pen, e.g., when dust aggregates collide with chondrules (Beitz et al. 2013) or CAIs. Recently, Bukhari Syed et al. (2017) have experimentally shown that the size ratio of colliding dust aggregates is an important factor for the collisional outcome. In addition to the size ratio, the density ratio of two colliding bodies might also affect the result of mutual collisions. For example, in more ener- getic impact-cratering experiments with a dust-aggregate projectile, ejecta coming from all over the target sur- face were observed (Wurm et al. 2005). However, we did not recognize such ejection in our experiments. In the high impact-energy regime, fragmentation of the tar- get is induced rather than the enhanced ejection of dust. To understand this difference, the influence of projectile density and structure have to be studied. Since this work aims at discussing the fundamental me- chanical properties of dust aggregates in the dynamical state, we employed a solid-projectile impact as a kind of the simplest test case. Although the collision outcomes could depend on the size and/or density ratio of the col- liding bodies, we believe that the energy-based descrip- tions, like Fig. 7, and the correct strength values should basically be universal in terms of local dynamics around the impact point. Using our current result, we can esti- mate the deformation and volume of the compressed zone around the impact point. The specific model to calculate the compression and the onset of fragmentation will be discussed in the next section (Sect. 4.3). Repeated impacts of the type of impacts studied here may cause surface compaction and would render the tar- get aggregate into a soft core with a hard shell around, as studied by Weidling et al. (2009). Such a process could induce a history dependence in dust-aggregate growth. Further systematic studies are necessary to discuss the importance or unimportance of surface compaction due to the low-energy impacts. As for cometary, asteroidal or lunar surface processes, it is known that their surfaces experience various impacts with objects of different size, density and impact velocity, as well as the fall-back of excavated material that leads to the formation of loose regolith layers. To properly sim- ulate the formation and compaction of regolith surfaces, a good knowledge of the impact physics into granular matter is required. Examples of applications of experi- mental impact studies such as the one described in this paper are high-velocity impacts into asteroidal surfaces with and without regolith (see, e.g., Beitz et al. 2016) or dust-aggregate impacts into granular matter (see, e.g., Planes et al. 2017). 4.3. Modeling of compaction and fragmentation of dust aggregates By using the results obtained so far, we propose a sim- ple model to estimate the fragmentation threshold and surface compaction induced by an impact into a dust ag- gregate. First, the fragmentation-threshold strength of dust aggregates can be estimated by Eq. (10). If the maximum dynamic pressure computed by Eq. (4) ex- ceeds this fragmentation threshold, fragmentation occurs for homogeneous dust aggregates. From the current re- sults, we cannot estimate the largest fragment mass and fragment-mass distribution. However, these properties were derived by Bukhari Syed et al. (2017) for aggregate- aggregate collisions. If the maximum dynamic pressure is less than the fragmentation threshold, compaction is induced. In these impact cases, a volume corresponding to the projectile volume Vp is compressed to Vp − Vc, the compaction proceeds by a factor Vp/(Vp − Vc). Equa- tion (7) provides the recipe to compute Vc. Although the value Yk = 120 kPa can be used for now, it is truly applicable for φ = 0.35 only. The dependence of Yk on the volume filling factor must be revealed in future stud- ies to complete the compaction modeling. The simplest approximation is Yk(φ)≃120φ/0.35 (kPa). If the geomet- rical condition at the fragmentation threshold D/z ∗ max in Eq. (11) is independent of φ, this approximation is plau- sible, because p∗ is proportional to φ in Eq. (10). Then, Vc, which is a key factor to calculate the compaction, can be estimated with Eq. (7). By assuming many random impacts, patchy-compacted surfaces result in an inhomo- geneity of the dust aggregates. To discuss the collision outcomes of inhomogeneous dust aggregates, more com- plex (history-dependent) impact tests are required both by experiments and numerical simulations. Our study provides only the first-step of modeling for such com- plex modes of dust-aggregate growth in the planetesimal formation stage. 4.4. Drag-force comparison with granular impact The projectile motion in this study can be fitted well by Eq. (6), which has been proposed to explain the gran- ular impact drag force. The two normalized fit parame- ters roughly become constant in the high-speed regime of our impacts (Fig. 6), but are velocity dependent in the low-speed regime. This variation probably comes from the elastic effect of the dust aggregates. As shown in Fig. 7(b), most of the impacts result in the rebound of the projectile, which is the elastic response of the com- pressed dust aggregate. The restitution coefficient de- pends on the impact energy (see Fig. 7(b)). When the impact energy is small, the restitution coefficient is rel- atively large. Therefore, the impacts possess an elastic as well as a plastic component. However, k/(πD) can basically be understood by considering the plastic de- formation of cratering, as discussed in Sect. 3.5. For small impact energies, the effect of elastic deformation is enhanced and added to plasticity-based k/(πD) value. This is a possible reason for the variation of k/(πD) in the small v0 regime. This partially elastic effect might also strengthen the inertial drag. The pseudo decrease of d1ρt/ρpD could stem from this effect. From the experi- mental data, we consider ǫ.0.15 is sufficient to apply the simple drag-force model with constant parameter values. In impact into granular matter, Fk originates from the In this study, however, the principal frictional effect. effect to cause Fk corresponds to plastic deformation. Both frictional and plastic effects are dissipative. How- ever, the form Fk = kz is similar to the elastic behavior in terms of the equation of motion. In this sense, the drag force model Eq. (1) should be improved. In fact, the value of k in the granular impact cannot be fully ex- plained by the simple Coulomb friction and hydrostatic pressure (Katsuragi & Durian 2007, 2013). Usually, the dissipative effect can be modeled by a rate-dependent term in the equation of motion. While we have a v2- 10 Katsuragi & Blum dependent term in Eq. (1), the v-dependent term is ab- sent. Although the absence of the v-dependent term re- lates to the rate-dependent granular viscosity (Katsuragi 2016), a v-dependent term improves the model in some cases (Nakamura et al. 2013). At present, Eq. (1) can explain the motion of projectile very well both in gran- ular and dust-aggregate impacts, at least as an empir- ical model. Furthermore, the form of Fk is consistent with the energy scaling in the strength-dominant regime, Eq. (7). To build more precise drag-force models, rate- dependent rheological properties should be studied in de- tail. 5. CONCLUSION A simple experiment to characterize the dynamic me- chanical properties of dust aggregates was performed. We dropped a solid spherical projectile onto a dust aggre- gate and measured the motion of the impinging projec- tile by using a high-speed camera. The experiment was performed in a vacuum chamber under the influence of gravity. From the acquired high-speed images, temporal developments of position z, velocity v, acceleration a of the projectile, as well as the dynamic pressure pdy were computed using image analysis. By varying the impact velocity and the density of the projectiles, the impact energy E was varied over two orders of magnitude. In the relatively low impact-energy regime, the impact re- sults in a spherical-indent cratering without any ejection of dust. In this cratering regime, rebound of the pro- jectile can also be observed. When the impact energy is large enough, fragmentation of the target dust aggregate is caused. To break the target, the dynamic pressure must exceed the threshold value of p∗ dy≃10 kPa. This critical dynamic pressure can be explained by the stress to open a crack. From the relation between crater vol- ume Vc and impact energy E, an effective yield strength Yk = E/Vc = 120 kPa can be obtained, which relates to the plastic deformation of the target aggregate and whose value is consistent with the drag-based strength k/(πD) = 200 kPa. Here, the value of k is computed from the fitting of kinematic data to Eq. (6). Another fit parameter in Eq. (6), d1, indicates that the inertial drag is less important for dust-aggregate impacts, except for the very initial stage of the impact. From the relation between the maximum dynamic pressure max(pdyn) and impact energy E, it turned out that the volume com- pressed by the impact Vs = E/max(pdyn)≃4.6×10−8 m3 is close to the projectile volume Vp. These results provide fundamental understandings about mechanical proper- ties of dust aggregates in the dynamic state. A simple model estimating the fragmentation threshold and the compaction degree of the impacted dust aggregate is pro- posed on the basis of our experimental result. HK thanks JSPS KAKENHI Grant Nos. 15KK0158 and 15H03707 for financial JB thanks the Deutsche Forschungsgemeinschaft (DFG, grant Bl 298/24-1) and the Deutsches Zentrum fur Luft- und Raumfahrt (DLR, grant 50WM1536) for financial sup- support. port. REFERENCES Altshuler, E., Torres, H., Gonz´alez-Pita, A., et al. 2014, Geophys. Res. Lett., 41, 3032 Ambroso, M. A., Santore, C. R., Abate, A. R., & Durian, D. J. 2005, Phys. Rev. E, 71, 051305 Beitz, E., Blum, J., Parisi, M. G., & Trigo-Rodriguez, J. 2016, ApJ, 824, 12 Beitz, E., Guttler, C., Nakamura, A. M., Tsuchiyama, A., & Blum, J. 2013, Icarus, 225, 558 Bertini, I., Gutierrez, P. J., & Sabolo, W. 2009, A&A, 504, 625 Blum, J., & Schrapler, R. 2004, Phys. Rev. Lett., 93, 115503 Blum, J., Schrapler, R., Davidsson, B. J. R., & Trigo-Rodriguez, J. M. 2006, ApJ, 652, 1768 Blum, J., & Wurm, G. 2008, Annu. Rev. Astro. Astrophys., 46, 21 Blum, J., Beitz, E., Bukhari, M., et al. 2014, JoVE (Journal of Visualized Experiments), e51541 Bukhari Syed, M., Blum, J., Jansson, K. W., & Johansen, A. 2017, ApJ, 834, 1 de Vet, S. J., & de Bruyn, J. R. 2007, Phys. Rev. E, 76, 041306 Fulle, M., & Blum, J. 2017, MNRAS, 469, S39 Guillard, F., Golshan, P., Shen, L., Vald`es, J. R., & Einav, I. 2015, Nat. Phys., 11, 835 Guttler, C., Blum, J., Zsom, A., Ormel, C. W., & Dullemond, C. P. 2010, A&A, 513, A56 Guttler, C., Krause, M., Geretshauser, R. J., Speith, R., & Blum, J. 2009, ApJ, 701, 130 Heim, L.-O., Blum, J., Preuss, M., & Butt, H.-J. 1999, Phys. Rev. Lett., 83, 3328 Herminghaus, S. 2013, Wet granular matter, A Truly Complex Fluid (World Scientific) Johnson, K. L. 1985, Contact Mechanics (Cambridge: Cambridge University Press) Katsuragi, H. 2010, Phys. Rev. Lett., 104, 218001 -- . 2016, Physics of Soft Impact and Cratering, Vol. LNP 910 (Springer) Katsuragi, H., & Durian, D. J. 2007, Nat. Phys., 3, 420 -- . 2013, Phys. Rev. E, 87, 052208 Meisner, T., Wurm, G., & Teiser, J. 2012, A&A, 544, A138 Nakamura, A. M., Setoh, M., Wada, K., Yamashita, Y., & Sangen, K. 2013, Icarus, 223, 222 Nelson, E. L., Katsuragi, H., Mayor, P., & Durian, D. J. 2008, Phys. Rev. Lett., 101, 068001 Pacheco-V´azquez, F., & Ruiz-Su´arez, J. C. 2011, Phys. Rev. Lett., 107, 218001 Planes, M. B., Mill´an, E. N., Urbassek, H. M., & Bringa, E. M. 2017, A&A, doi:https://doi.org/10.1051/0004-6361/201730954, in press Poppe, T., & Schrapler, R. 2005, A&A, 438, 1 Royer, J. R., Conyers, B., Corwin, E. I., Eng, P. J., & Jaeger, H. M. 2011, EPL, 93, 28008 Scheel, M., Seemann, R., Brinkmann, M., et al. 2008, J. Phys.: Condens. Matter, 20, 494236 Schrapler, R., Blum, J., Krijt, S., & Raabe, J.-H. 2017, submitted to ApJ Sirono, S.-i. 2004, Icarus, 167, 431 Takehara, Y., Fujimoto, S., & Okumura, K. 2010, EPL, 92, 44003 Trigo-Rodriguez, J. M., & Llorca, J. 2006, Monthly Notices of the Royal Astronomical Society, 372, 655 Uehara, J. S., Ambroso, M. A., Ojha, R. P., & Durian, D. J. 2003, Phys. Rev. Lett., 90, 194301 Umbanhowar, P., & Goldman, D. I. 2010, Phys. Rev. E, 82, 010301 Vald`es, J. R., Fernandes, F. L., & Einav, I. 2011, Granular Matter, 14, 71 Walsh, A. M., Holloway, K. E., Habdas, P., & de Bruyn, J. R. 2003, Phys. Rev. Lett., 91, 104301 Weidling, R., Guttler, C., Blum, J., & Brauer, F. 2009, ApJ, 696, 2036 Wurm, G., Paraskov, G., & Krauss, O. 2005, Phys. Rev. E, 71, 021304 Zsom, A., Ormel, C. W., Guttler, C., Blum, J., & Dullemond, C. P. 2010, A&A, 513, A57
1606.00848
3
1606
2016-08-15T15:35:59
HATS-18 b: An Extreme Short--Period Massive Transiting Planet Spinning Up Its Star
[ "astro-ph.EP", "astro-ph.SR" ]
We report the discovery by the HATSouth network of HATS-18 b: a 1.980 +/- 0.077 Mj, 1.337 +0.102 -0.049 Rj planet in a 0.8378 day orbit, around a solar analog star (mass 1.037 +/- 0.047 Msun, and radius 1.020 +0.057 -0.031 Rsun) with V=14.067 +/- 0.040 mag. The high planet mass, combined with its short orbital period, implies strong tidal coupling between the planetary orbit and the star. In fact, given its inferred age, HATS-18 shows evidence of significant tidal spin up, which together with WASP-19 (a very similar system) allows us to constrain the tidal quality factor for Sun-like stars to be in the range 6.5 <= lg(Q*/k_2) <= 7 even after allowing for extremely pessimistic model uncertainties. In addition, the HATS-18 system is among the best systems (and often the best system) for testing a multitude of star--planet interactions, be they gravitational, magnetic or radiative, as well as planet formation and migration theories.
astro-ph.EP
astro-ph
Draft version 00 August 16, 2016 Preprint typeset using LATEX style emulateapj v. 2/16/10 HATS-18b: AN EXTREME SHORT–PERIOD MASSIVE TRANSITING PLANET SPINNING UP ITS STAR † K. Penev1, J. D. Hartman1, G. ´A. Bakos1,⋆,⋆⋆, S. Ciceri6, R. Brahm4,5, D. Bayliss2,3, J. Bento2, A. Jord´an4,5, Z. Csubry1, W. Bhatti1, M. de Val-Borro1, N. Espinoza4,5, G. Zhou1, L. Mancini6, M. Rabus4,6, V. Suc4, T. Henning6, B. Schmidt2, R. W. Noyes9, J. L´az´ar12, I. Papp12, P. S´ari12, 6 1 0 2 g u A 5 1 . ] P E h p - o r t s a [ 3 v 8 4 8 0 0 . 6 0 6 1 : v i X r a Draft version 00 August 16, 2016 ABSTRACT We report the discovery by the HATSouth network of HATS-18b: a 1.980±0.077 MJ, 1.337+0.102 −0.049 RJ planet in a 0.8378 day orbit, around a solar analog star (mass 1.037 ± 0.047 M⊙, and radius 1.020+0.057 −0.031 R⊙) with V = 14.067 ± 0.040 mag. The high planet mass, combined with its short or- bital period, implies strong tidal coupling between the planetary orbit and the star. In fact, given its inferred age, HATS-18 shows evidence of significant tidal spin up, which together with WASP-19 (a very similar system) allows us to constrain the tidal quality factor for Sun–like stars to be in the range 6.5 . log10(Q∗/k2) . 7 even after allowing for extremely pessimistic model uncertainties. In addition, the HATS-18 system is among the best systems (and often the best system) for testing a multitude of star–planet interactions, be they gravitational, magnetic or radiative, as well as planet formation and migration theories. Subject headings: planetary systems - stars: individual (HATS-18) - techniques: spectroscopic, photometric 1. INTRODUCTION Hot Jupiters, gas giant planets with orbital periods shorter than a few days, are among the easiest extra- solar planets to detect through either transit or radial velocity (RV) searches (to date, the two most productive methods). In spite of that, the sample of these planets is rather small, showing that they are intrinsically rare. Among those, giant planets with extreme short-period orbits, say under one day, are the easiest to detect yet the most scarce. In fact, out of the 4696 candidate planet Kepler objects of interest (KOI) on the NASA exoplanet archive12, only 229 have a radius of at least 6 Earth radii 1 Department of Astrophysical Sciences, Princeton University, NJ 08544, USA ⋆ Alfred P. Sloan Research Fellow ⋆⋆ Packard Fellow 2 Research School of Astronomy and Astrophysics, Australian National University, Canberra, ACT 2611, Australia 3 Observatoire Astronomique de l'Universit´e de Gen`eve, 51 ch. des Maillettes, 1290 Versoix, Switzerland 4 Instituto de Astrof´ısica, Facultad de F´ısica, Pontificia Uni- versidad Cat´olica de Chile, Av. Vicuna Mackenna 4860, 7820436 Macul, Santiago, Chile; [email protected] 5 Millennium Institute of Astrophysics, Av. Vicuna Mackenna 4860, 7820436 Macul, Santiago, Chile 6 Max Planck Institute for Astronomy, Heidelberg, Germany 9 Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA 12 Hungarian Astronomical Association, Budapest, Hungary † The HATSouth network is operated by a collaboration con- sisting of Princeton University (PU), the Max Planck Insti- tute fur Astronomie (MPIA), the Australian National Univer- sity (ANU), and the Pontificia Universidad Cat´olica de Chile (PUC). The station at Las Campanas Observatory (LCO) of the Carnegie Institute is operated by PU in conjunction with PUC, the station at the High Energy Spectroscopic Survey (H.E.S.S.) site is operated in conjunction with MPIA, and the station at Siding Spring Observatory (SSO) is operated jointly with ANU. This paper includes data gathered with the MPG 2.2 m telescope at the ESO Observatory in La Silla. This paper uses observa- tions obtained with facilities of the Las Cumbres Observatory Global Telescope. 12 http://exoplanetarchive.ipac.caltech.edu/ (approximately half the radius of Jupiter) and orbital periods shorter than 5 days, and of those, only 41 have a periods shorter than 1 day. This, combined with the fact, that these are expected to be the KOIs with the highest chance of being false positives (c.f. Fressin et al. 2013), have the highest probability to transit and that none of the transiting ones should be missed by Kepler, demonstrates how unusual these planetary systems are. On the other hand, this exotic population of planets, especially the ones transiting their star, is very valuable, since it pushes theories of planet formation, structure and evolution, as well as planet–star interactions to the limit (c.f. Ida & Lin 2008; Dawson & Murray-Clay 2013; Albrecht et al. 2012; Ginzburg & Sari 2015; Penev et al. 2012). In addition, the deep and frequent transits and large RV signals of these objects make them the easiest to carry follow–up studies on, thus enhancing their power to constrain theories even further. We report the discovery by the HATSouth transit sur- vey (Bakos et al. 2013) of HATS-18b: a very short pe- riod (0.8378 day) massive (1.980 ± 0.077 MJ) extraso- lar planet around a star very similar to our Sun (mass 1.037 ± 0.047 M⊙, radius 1.020+0.057 −0.031 R⊙ and effective temperature 5600 ± 120 K). Due to the proximity of the planet to its host star, this system provides one of the best laboratories for testing theories of star–planet in- teractions and planet formation. In fact, we argue that HATS-18 shows signs of being tidally spun–up by the planet, and that modelling this effect for this system alone constrains the tidal dissipation efficiency of the host star to better than an order of magnitude even with very generous assumptions on possible formation scenarios or model parameter uncertainties. Further, we show that expanding such models to the few other very short period systems, should drastically improve that constraint. Further, such modelling may begin to dis- entangle some of the very poorly understood physics be- hind tidal dissipation by measuring its dependence on 2 Penev et al. -0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 g a m ∆ g a m ∆ 0.06 -0.1 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.05 0 Orbital phase 0.05 0.1 Fig. 1.- Unbinned instrumental r band light curve of HATS- 18 folded with the period P = 0.8378434 days resulting from the global fit described in Section 3. The solid line shows the best-fit transit model (see Section 3). In the lower panel we zoom–in on the transit; the dark filled points here show the light curve binned in phase using a bin-size of 0.002. various system properties. The layout of the paper is as follows: in § 2 we de- scribe the discovery and follow–up observations used to confirm HATS-18b as a planet; in § 3 we outline the com- bined photometric and spectroscopic analysis performed and give the inferred system properties; in § 4 we place HATS-18 in the context of other extremely short period exoplanet systems; in § 5 we derive constraints on the tidal quality factor for stars similar to the Sun by mod- elling HATS-18 and WASP-19's orbital and stellar spin evolution; and we conclude with a discussion in § 6. 2. OBSERVATIONS 2.1. Photometry 2.1.1. Photometric detection The star HATS-18 (Table 3) was observed by HAT- South instruments between UT 2011 April 18 and UT 2013 July 21 using the HS-2, HS-4, and HS-6 units at the Las Campanas Observatory in Chile, the High Energy Spectroscopic Survey site in Namibia, and Siding Spring Observatory in Australia, respectively. A total of 5372, 3758 and 4008 images of HATS-18 were obtained with HS-2, HS-4 and HS-6, respectively. The observations were obtained through a Sloan r filter with an exposure time of 240 s. The data were reduced to trend-filtered light curves using the aperture photometry pipeline de- scribed by Penev et al. (2013) and making use of Exter- nal Parameter Decorrelation (EPD; Bakos et al. 2010) and the Trend Filtering Algorithm (TFA; Kov´acs et al. 2005) to remove systematic variations. We searched for transits using the Box Least Squares (BLS; Kov´acs et al. 2002) algorithm, and detected a P = 0.8378 day peri- odic transit signal in the light curve of HATS-18 (Fig- ure 1; the data are available in Table 1). After detect- ing the signal we re-applied the TFA filter, this time in signal reconstruction mode, so as to obtain an undis- torted trend-filtered light curve. The per–point root mean square (RMS) residual combined filtered HAT- South light curve (after subtracting the best-fit model transit) is 0.015 mag, which is typical for a star of this magnitude. 2.1.2. Photometric follow-up We obtained follow-up light curves of HATS-18 using the LCOGT 1 m telescope network. An ingress was ob- served on UT 2015 July 18 with the SBIG camera and a Sloan i filter on the 1 m at the South African Astro- nomical Observatory (SAAO). A total of 33 images were collected at a median cadence of 201 s. A full transit was observed on UT 2016 Jan 22 with the sinistro camera and a Sloan i filter on the 1 m at Cerro Tololo Inter-American Observatory (CTIO). A total of 61 images were collected at a median cadence of 219 s. For the record we also note that a full transit was observed on UT 2016 January 3 with the SBIG camera on the 1 m at SAAO, however due to tracking and weather problems we were unable to extract high precision photometry from these images, and therefore do not include these data in our analysis. For details of the reduction procedure used to extract light curves from the raw images see Penev et al. (2013). The follow-up light curves are shown, together with our best-fit model, in Figure 2, while the data are available in Table 1. The per-point precision of the SBIG obser- vations is 2.5 mmag, while the per-point precision of the sinistro observations is 1.7 mmag. 2015 July 18 (LCOGT 1m) 2016 Jan 22 (LCOGT 1m) i-band 0 0.02 i s t e s f f o y r a r t i b r A - ) g a m ( ∆ 0.04 0.06 0.08 0.1 -0.1 -0.05 0 0.05 0.1 Time from transit center (days) Fig. 2.- Unbinned follow-up transit light curve of HATS-18 obtained with telescopes from the LCOGT 1 m network. Our best fit is shown by the solid lines. The residuals from the best-fit model are shown below in the same order. 2.2. Spectroscopy Spectroscopic follow-up observations of HATS-18 were carried out with WiFeS on the ANU 2.3 m telescope (Dopita et al. 2007) and with FEROS on the MPG 2.2 m (Kaufer & Pasquini 1998). A total of three spectra were obtained with WiFeS be- tween UT 2015 Feb 28 and UT 2015 Mar 2, two at a resolution of R ≡ ∆ λ / λ = 7000, and one at R = 3000. These data were reduced and analyzed following the pro- cedure described by Bayliss et al. (2013). The R = 3000 spectrum was used to estimate the spectral type and sur- face gravity of HATS-18 (we find that it is a G dwarf), while the R = 7000 spectra were used to rule out an RV variation greater than 5 km s−1. HATS-18b TABLE 1 3 BJD (2 400 000+) 56442.67216 56411.67208 56343.80691 56444.34817 56392.40213 56395.75361 56416.69970 56469.48392 56446.86219 56458.59202 Differential photometry of HATS-18 Maga σMag Mag(orig)b Filter Instrument 0.03291 −0.01713 −0.00841 0.01720 −0.00247 0.02231 −0.02166 −0.02100 −0.00054 −0.03131 0.00789 0.00723 0.00655 0.00712 0.00646 0.00737 0.00680 0.00717 0.00641 0.00744 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · r r r r r r r r r r HS HS HS HS HS HS HS HS HS HS Note. - This table is available in a machine-readable form in the online jour- nal. A portion is shown here for guidance regarding its form and content. The data are also available on the HATSouth website at http://www.hatsouth.org. a The out-of-transit level has been subtracted. For the HATSouth light curve (rows with "HS" in the Instrument column), these magnitudes have been de- trended using the EPD and TFA procedures prior to fitting a transit model to the light curve. We apply the TFA in signal-reconstruction mode so as to preserve the transit depth. For the follow-up light curves (rows with an Instru- ment other than "HS") these magnitudes have been detrended with the EPD procedure, carried out simultaneously with the transit fit. b Raw magnitude values without application of the EPD procedure. This is only reported for the follow-up light curves. We obtained six R = 48000 spectra with FEROS be- tween UT 2015 Jun 12 and UT 2015 Jun 20. These were reduced to high precision RV and spectral line bisector span (BS) measurements following Jord´an et al. (2014), and were also used to determine high precision atmo- spheric parameters (Section 3). The RVs show a clear K = 415.2±10.0 m s−1 sinusoidal variation in phase with the transit ephemeris (Figure 3; the data are provided in Table 2), confirming this object as a transiting planet system. The BSs exhibit significant scatter, as is typical for a faint V = 14.067 ± 0.040 mag star, but are uncor- related with the RVs. The scatter is also well below the level expected if this were a blended stellar eclipsing bi- nary system (Section 3). 3. ANALYSIS We analyzed the photometric and spectroscopic ob- servations of HATS-18 to determine the parameters of the system using the standard procedures developed for HATNet and HATSouth (see Bakos et al. 2010, with modifications described by Hartman et al. 2012). High-precision stellar atmospheric parameters were measured from the FEROS spectra using ZASPE (Brahm et. al. 2016). The resulting Teff⋆ and [Fe/H] measurements were combined with the stellar density ρ⋆ determined through our joint light curve and RV curve analysis, to determine the stellar mass, radius, age, lu- minosity, and other physical parameters, by comparison with the Yonsei-Yale (Y2; Yi et al. 2001) stellar evolution models (see Figure 4). This provided a revised estimate of log g⋆ which was fixed in a second iteration of ZA- SPE. Our final adopted stellar parameters are listed in Table 3. We find that the star HATS-18 has a mass of 1.037 ± 0.047 M⊙, a radius of 1.020+0.057 −0.031 R⊙, and is at a reddening-corrected distance of 645+36 −25 pc. We simultaneously carried out a joint analysis of the High-precision FEROS RVs (fit using a Keplerian orbit) and the HS and LCOGT 1 m light curves (fit using a Mandel & Agol 2002 transit model with fixed quadratic 400 200 ) 1 - s m ( V R 0 -200 -400 ) 1 - s m ( - C O 50 40 30 20 10 0 -10 -20 -30 -40 200 ) 1 - s m ( S B 150 100 50 0 -50 0.0 0.2 0.4 0.6 0.8 1.0 Phase with respect to Tc Fig. 3.- Top panel: High-precision RV measurements from MPG 2.2 m/FEROS together with our best-fit orbit model. Zero phase corresponds to the time of mid-transit. The center-of-mass velocity has been subtracted. Second panel: Velocity O−C residuals from the best-fit model. The error bars include the jitter which is varied in the fit. Third panel: Bisector spans (BS). Note the different vertical scales of the panels. limb darkening coefficients taken from Claret 2004) to measure the stellar density, as well as the orbital and planetary parameters. This analysis makes use of a dif- ferential evolution Markov Chain Monte Carlo procedure (DEMCMC; ter Braak 2006) to estimate the posterior parameter distributions, which we use to determine the median parameter values and their 1σ uncertainties. The results are listed in Table 4. We find that the planet Penev et al. TABLE 2 Relative radial velocities and bisector span measurements of HATS-18. BJD (2 457 100+) RVa (m s−1) b σRV (m s−1) BS (m s−1) σBS Phase Instrument 85.64999 86.50430 88.59324 90.51136 91.49572 93.58965 −389.04 −391.04 424.96 −207.04 −420.04 407.96 21.00 20.00 20.00 15.00 17.00 17.00 66.0 19.0 −24.0 41.0 −10.0 149.0 24.0 21.0 22.0 16.0 18.0 18.0 0.274 0.294 0.787 0.076 0.251 0.750 FEROS FEROS FEROS FEROS FEROS FEROS a Relative RVs, with γRV (see table 3) subtracted. b jitter considered in Section 3. Internal errors excluding the component of astrophysical/instrumental 4 ] 3 m c / g [ * ρ 0.0 0.5 1.0 1.5 2.0 2.5 r e w o P ] g a m [ r ∆ 0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 10-1 0.06 0.04 0.02 0.00 −0.02 −0.04 −0.06 0.0 100 101 Period [days] 102 0.5 1.0 1.5 2.0 Rotational Phase Fig. 5.- Top: The Lomb–Scargle periodogram of HATS-18 light curves (in signal reconstruction mode for the transits but not the rotational modulation) with transits removed. Bottom: the same lightcurve folded with the best–fit stellar spin period (the points in the second half of the plot are duplicates of those in the first half). The lightcurve of HATS-18 shows a clear signature of stellar spin variability. In Fig. 5 we show the Lomb–Scargle periodogram of the HATSouth discovery lightcurve of HATS-18, with observations during tran- sits removed, as well as the lightcurve as a function of the best fit spin period (9.8 days) phase. In order to get a handle on the uncertainty in the stellar spin period we split the lightcurve into 9 segments, each containing three spin periods and fit for the rotation period in each segment separately and adopt the standard deviation of the individual measurements as the period uncertainty. The resulting spin period estimate is Prot⋆ = 9.8 ± 0.4 days. 6200 6000 5800 5600 Effective temperature [K] 5400 Fig. 4.- Comparison between the measured values of Teff⋆ and ρ⋆ (from ZASPE applied to the FEROS spectra, and from our modeling of the light curves and RV data, respectively), and the Y2 model isochrones from Yi et al. (2001). The best-fit values (dark filled circle), and approximate 1σ and 2σ confidence ellipsoids are shown. The values from our initial ZASPE iteration are shown with the open triangle. The Y2 isochrones are shown for ages of 0.2 Gyr, and 1.0 to 14.0 Gyr in 1 Gyr increments. HATS-18b has a mass of 1.980 ± 0.077 MJ, and a radius of 1.337+0.102 −0.049 RJ. We fit the data both assuming a cir- cular orbit, and allowing for a non-zero eccentricity. We find that the observations are consistent with a circular orbit: e = 0.063 ± 0.049, with a 95% confidence upper- limit of e < 0.166, and therefore adopt the parameters that come from assuming a circular orbit (we also find that the Bayesian evidence for the circular orbit model is higher than the evidence for the free-eccentricity model). 3.1. Ruling Out Blended Models In order to rule out the possibility that HATS-18 is a blended stellar eclipsing binary system, we carried out a blend analysis of the photometric data following Hartman et al. (2012). We find that all blend models tested can be rejected based on the photometry alone with 3.5σ confidence. Moreover, the blend models which come closest to fitting the photometry (those which can- not be rejected with greater than 5σ confidence) yield simulated RVs that are not at all similar to what we ob- serve (i.e., the simulated blend-model RVs do not show a sinusoidal variation in phase with the photometric ephemeris). We conclude that HATS-18 is not a blended stellar eclipsing binary system, and is instead a transiting planet system. 3.2. Photometric Rotation Period HATS-18b 5 Stellar Parameters for HATS-18 TABLE 3 Parameter Value Source Identifying Information R.A. (h:m:s) Dec. (d:m:s) R.A.p.m. (mas/yr) Dec.p.m. (mas/yr) GSC ID 2MASS ID Spectroscopic properties Teff⋆ (K) . . . . . . . . . Spectral type . . . . . [Fe/H] . . . . . . . . . . . . v sin i (km s−1) . . . γRV (m s−1) . . . . . . Photometric properties B (mag) . . . . . . . . . . V (mag) . . . . . . . . . . g (mag) . . . . . . . . . . r (mag) . . . . . . . . . . i (mag) . . . . . . . . . . . J (mag) . . . . . . . . . . H (mag). . . . . . . . . . Ks (mag). . . . . . . . . Derived properties M⋆ (M⊙). . . . . . . . . R⋆ (R⊙) . . . . . . . . . log g⋆ (cgs) . . . . . . . ρ⋆ (g cm−3) c . . . . . ρ⋆ (g cm−3) c . . . . . L⋆ (L⊙) . . . . . . . . . . MV (mag) . . . . . . . . MK (mag,ESO) Age (Gyr) . . . . . . . . AV (mag) d . . . . . . Distance (pc) . . . . . Prot⋆ (days) . . . . . . 11h35m49.92s −29◦09′21.6′′ 2.7 ± 1.2 −4.4 ± 1.2 GSC 6664-00410 2MASS 11354977-2909216 5600 ± 120 G 0.280 ± 0.080 6.23 ± 0.47 7663.3 ± 7.7 14.870 ± 0.060 14.067 ± 0.040 14.407 ± 0.020 13.854 ± 0.030 13.77 ± 0.15 12.736 ± 0.026 12.382 ± 0.028 12.289 ± 0.028 1.037 ± 0.047 1.020+0.057 −0.031 4.436 ± 0.034 1.38+0.13 −0.21 1.37+0.12 −0.23 0.93 ± 0.13 4.94 ± 0.17 3.281 ± 0.099 4.2 ± 2.2 0.076+0.114 −0.076 645+36 −25 9.8 ± 0.4 2MASS 2MASS 2MASS 2MASS GSC 2MASS ZASPE a ZASPE ZASPE ZASPE FEROS APASS APASS APASS APASS APASS 2MASS 2MASS 2MASS Y2+ρ⋆+ZASPE b Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE Light curves Y2+Light curves+ZASPE Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE Y2+ρ⋆+ZASPE HATSouth light curve a ZASPE = "Zonal Atmospherical Stellar Parameter Estimator" method for the analysis of high-resolution spectra applied to the FEROS spectra of HATS-18. These parameters rely primarily on ZASPE, but have a small dependence also on the iterative analysis incorporating the isochrone search and global modeling of the data, as described in the text. b Isochrones+ρ⋆+ZASPE = Based on the Y2 isochrones (Yi et al. 2001), the stellar density used as a luminosity indicator, and the ZASPE results. c We list two values for ρ⋆. The first value is determined from the global fit to the light curves and RV data, without imposing a constraint that the parameters match the stellar evolution models. The second value results from restricting the posterior distribution to combinations of ρ⋆+Teff ⋆+[Fe/H] that match to a Y2 stellar model. d Total V band extinction to the star determined by comparing the catalog broad-band photometry listed in the table to the expected magnitudes from the Isochrones+ρ⋆+ZASPE model for the star. We use the Cardelli et al. (1989) extinc- tion law. 4. COMPARISON TO OTHER SHORT PERIOD SYSTEMS Due to its very short orbital period and relatively high planetary mass, the HATS-18 system is ideal for test- ing theories of star–planet interactions, whether those occur through radiation, gravity or magnetic fields. Fig- ures 6-8 show a comparison between the present sample of giant planets (mass at least 0.1 MJ) in orbital peri- ods shorter than two days and the HATS-18 system in a number of parameters related to the strength of vari- ous star–planet interactions that have been suggested to occur. The possible magnetic interactions (and hence their observable effects) are expected to grow in strength the deeper the planet is in its star's magnetic field and the stronger the field is. In general, stars with surface convec- tive zones are expected to have much stronger magnetic fields than stars with surface radiative zones, since in the former case some form of convectively driven dynamo is expected to operate in the stellar envelope. Further, the dynamo is expected to generate a larger field for faster rotating stars, hence the two readily observable quan- tities to compare in order to gauge the observability of magnetic star–planet interactions are the size of the orbit relative to the stellar radius (a/R⋆) and the stellar spin 6 Penev et al. Parameters for the transiting planet HATS-18b. TABLE 4 Parameter Value a Light curve parameters P (days) . . . . . . . . . . . . . . . . . . Tc (BJD) b . . . . . . . . . . . . . . . T14 (days) b . . . . . . . . . . . . . . T12 = T34 (days) b . . . . . . . . a/R⋆ . . . . . . . . . . . . . . . . . . . . . c . . . . . . . . . . . . . . . . . . . . ζ/R⋆ Rp/R⋆ . . . . . . . . . . . . . . . . . . . . b2 . . . . . . . . . . . . . . . . . . . . . . . . . b ≡ a cos i/R⋆ . . . . . . . . . . . . . i (deg) . . . . . . . . . . . . . . . . . . . . 0.83784340 ± 0.00000047 2457089.90598 ± 0.00026 0.07886 ± 0.00083 0.0101 ± 0.0010 3.71+0.11 −0.22 29.09+0.26 −0.19 0.1347 ± 0.0019 0.085+0.110 −0.054 0.29+0.15 −0.11 85.5+1.9 −2.8 Limb-darkening coefficients d c1, i (linear term) . . . . . . . . . c2, i (quadratic term) . . . . . c1, r . . . . . . . . . . . . . . . . . . . . . . c2, r . . . . . . . . . . . . . . . . . . . . . . RV parameters K (m s−1) . . . . . . . . . . . . . . . . e e . . . . . . . . . . . . . . . . . . . . . . . . FEROS RV jitter (m s−1) f Planetary parameters Mp (MJ) . . . . . . . . . . . . . . . . . Rp (RJ) . . . . . . . . . . . . . . . . . . C(Mp, Rp) g . . . . . . . . . . . . . . ρp (g cm−3) . . . . . . . . . . . . . . . log gp (cgs) . . . . . . . . . . . . . . . . a (AU) . . . . . . . . . . . . . . . . . . . . Teq (K) h . . . . . . . . . . . . . . . . . Θ i . . . . . . . . . . . . . . . . . . . . . . . hF i (109erg s−1 cm−2) i . . . 0.3097 0.3143 0.4124 0.2959 415.2 ± 10.0 < 0.166 < 11.4 1.980 ± 0.077 1.337+0.102 −0.049 0.36 1.02+0.13 −0.20 3.435+0.035 −0.063 0.01761 ± 0.00027 2060 ± 59 0.0498+0.0025 −0.0033 4.07 ± 0.48 a For each parameter we give the median value and 68.3% (1σ) confidence intervals from the posterior distribution. Reported results assume a circular orbit. b Reported times are in Barycentric Julian Date calculated directly from UTC, without correction for leap seconds. Tc: Reference epoch of mid transit that minimizes the correlation with the orbital period. T14: total transit duration, time be- tween first to last contact; T12 = T34: ingress/egress time, time between first and second, or third and fourth contact. c Reciprocal of the half duration of the transit used as a jump parameter in our MCMC analysis in place of a/R⋆. It is related to a/R⋆ by the expression ζ/R⋆ = a/R⋆(2π(1 + e sin ω))/(P√1 − b2√1 − e2) (Bakos et al. 2010). d Values for a quadratic law, adopted from the tabulations by Claret (2004) according to the spectroscopic (ZASPE) parame- ters listed in Table 3. e The 95% confidence upper-limit on the eccentricity. All other parameters listed are determined assuming a circular orbit. f Error term, either astrophysical or instrumental in origin, added in quadrature to the formal RV errors. This term is varied in the fit assuming a prior that is inversely proportional to the jitter. We find that the jitter is consistent with zero, and thus give the 95% confidence upper limit. g Correlation coefficient between the planetary mass Mp and radius Rp determined from the parameter posterior distribution via C(Mp, Rp) = h(Mp−hMpi)(Rp−hRpi)i/(σMp σRp )i, where h·i is the expectation value operator, and σx is the standard deviation of parameter x. h Planet equilibrium temperature averaged over the orbit, cal- culated assuming a Bond albedo of zero, and that flux is re– radiated from the full planet surface. i The Safronov number is given by Θ = 1 (a/Rp)(Mp/M⋆) (see Hansen & Barman 2007). j Incoming flux per unit surface area, averaged over the orbit. 2 (Vesc/Vorb)2 = HATS-18b 7 8 7 6 ⋆ R / a 5 4 KELT-1 3 WASP-18 2 0 10 WASP-19 OGLE-TR-056 WASP-103 20 30 Prot ⋆ [days] WASP-12 40 50 103 102 ] c e s [ (cid:1) 6 0 1 / 101 ∗ Q (cid:0) t f i h s T 100 10-1 0.5 KELT-1 WASP-18 WASP-103 WASP-33 WASP-12 WASP-19 OGLE-TR-056 WASP-43 1.0 1.5 2.0 2.5 3.0 Transit Depth [%] Fig. 6.- Size of the planetary orbit relative to the stellar ra- dius as a function of the stellar rotation period, estimated using the measured projected equatorial velocity of the stars and their estimated radii. Planets other than HATS-18(big star symbol) are all transiting planets from the NASA Exoplanet Archive with or- bital periods shorter than 2 days and masses at least 0.1 MJ. Filled symbols: host star effective temperature is below 6250 K (surface convective zone stars); empty symbols: host star effective tem- perature is above 6250 K (surface radiative zone stars); half–filled symbols: host star effective temperature is consistent with 6250 K within quoted error bars. WASP-33 KELT-1 WASP-12 WASP-103 WASP-18 OGLE-TR-056 CoRoT-14 WASP-19 2.8 2.6 2.4 2.2 2.0 1.8 1.6 1.4 ] K 0 0 0 1 [ f f l e p T 1.2 0 2 4 6 8 10 12 Rroche/Rpl Fig. 7.- The equilibrium temperature of the planet, assuming ideal black body against the fraction of the ratio of the Roche radius to planet radius for the same systems plotted in Fig. 6. period. From Fig. 6, we see that HATS-18 is among the three surface convective zone systems (HATS-18, WASP- 19 and OGLE-TR-56) whose error bars are consistent with having the smallest a/R⋆ and among those it has the shortest stellar spin period (inferred either from its projected spin velocity, or the observed rotational mod- ulation in its lightcurve). Another rather dramatic effect of star–planet interac- tions is for the stellar irradiation/wind to drive outflows from the planet. Clearly this process will occur more readily for planets closer to filling their Roche radius and for hotter planets. Fig. 7 plots the ratio of the plan- etary to the Roche radius for each system against the equilibrium effective temperature for the planet (assum- ing a perfect black body) for the same sample of plan- ets as in Fig. 6. Again, HATS-18 is among the plan- ets with most favourable parameters, although in this case there is a cluster of very–hot, very small Roche ra- tio planets around surface radiative zone stars, for one of which (WASP-12b) outflows have been claimed (c.f. Fig. 8.- The shift in mid–transit time ephemeris after a decade for a tidal quality factor of Q⋆ = 106. Fossati et al. 2010; Haswell et al. 2012). . The most direct way of detecting tidal interactions be- tween a star and its companion planet is to see the or- bital decay due to tidal dissipation in the star. This is most readily accomplished through observing the result- ing deviation from a linear mid–transit time ephemeris. Detecting this effect will provide a direct measurement of the tidal dissipation efficiency of the parent star: the least constrained parameter in tidal interactions involv- ing stars and giant planets. Fig. 8 shows that HATS-18b is the planet around a convective envelope star with the largest expected shift in mid–transit time after a decade. 5. HOST STAR SPIN–UP AND A MEASUREMENT OF Q⋆ Given that HATS-18 has an age consistent with the age of the Sun, and that it is very close to solar mass, its spin period should be close to that of the Sun or to the re- cently measured rotation periods in the 4.2 Gyr old open cluster M 67 (Barnes et al. 2016): Prot⋆ ≈ 30 days, even if the stellar age were at the lower end of the estimated er- ror bar (2.2 Gyrs), the expected spin period is Prot⋆ ≈ 20. Instead, in § 3 we found v sin i and stellar radius corre- sponding to a spin period of Prot⋆ = 8.3±0.8 days, which is consistent with the photometrically determined rota- tion period of Prot⋆ = 9.8 ± 0.4 days. This much faster spin rate is close to what is observed for Solar mass stars in clusters with ages around 600 Myr: the 550 Myr old M 37 (Hartman et al. 2009), the 580 Myr old Praesepe (Agueros et al. 2011; Delorme et al. 2011; Kov´acs et al. 2014), and the 625 Myr old Hyades (Delorme et al. 2011). A natural explanation for this apparent discrepancy is suggested by the fact that the HATS-18 system contains a very short period giant planet, which should have expe- rienced some orbital decay due to tidal dissipation in the star. The angular momentum taken out of the planetary orbit as it shrinks is deposited in the star and hence the star is spun–up. The fact that we see evidence for this tidal spin–up, means that we can use it to measure the tidal dissipation properties of the star. In this section we describe a method for carrying out such a measurement and show the resulting constraints. 5.1. The Tidal and Stellar Spin Model Stars like HATS-18 continuously lose angular momen- tum throughout their lifetime by magnetically impart- ing angular momentum to the wind of charged particles 8 Penev et al. launched from their surfaces. As a result, in order to re- late the stellar tidal dissipation efficiency to the observed stellar spin, we need to model this angular momentum loss simultaneously with the tidal spin–up. There are a number of options for modelling the tidal evolution, and the angular momentum loss. However, in an effort to keep the number of model parameters small while constructing a consistent model we will use the tidal evolution formulation of (Lai 2012) and assume a constant value for Q′⋆ ≡ Q⋆/k2, where Q⋆ is the fraction of tidal energy lost in one orbital period, and k2 is the Love number of the star. Note that, while tidal dissi- pation in the planet may be more efficient than in the star, it will quickly result in a circular orbit and plane- tary spin synchronized with the orbit, which will make the tidal deformation of the planet static and hence not subject to dissipation. Further, assuming constant dis- sipation efficiency is clearly not physical. In particular, the dissipation should vanish (Q′⋆ = ∞) when the tidal frequency approaches zero and increase gradually as the frequency moves away from zero. However, for tidal fre- quencies near that observed for HATS-18, the dissipation is expected to become less efficient as the frequency in- creases. Since there is currently no agreement on the ex- pected dependence of Q′⋆ on frequency and other param- eters, we don't have a choice but to assume Q′⋆ = const. In practice, the way to interpret the results is that the Q′⋆ measured by our analysis is appropriate for the currently observed state of the system analyzed, since the observed spin–up of the host star is overwhelmingly dominated by the very recent tidal evolution (see Fig. 9). We will model the star as consisting of two distinct zones: the surface convective envelope and the radia- tive core, and all tidal dissipation will be assumed to occur in the envelope. As a result, any angular mo- mentum lost by the orbit will be deposited exclusively in the convective zone of HATS-18. This will tend to drive differential rotation between the core and the enve- lope, which will in turn be suppressed by at present not well understood coupling processes, but its efficiency is reasonably constrained by observations (c.f. Irwin et al. 2007; Gallet & Bouvier 2015; Amard et al. 2016). The model for the evolution of the stellar spin tracks a sin- gle value for the spin of each zone, allows for angular momentum exchange between the core and the envelope and for angular momentum loss due to the stellar wind. The particular formulation we will use is given in detail in Irwin et al. (2007). The loss of angular momentum from the convective envelope due to the wind is given by: d ~Jconv dt !wind ≡ K~ωconv min(~ωconv2 , ω2 sat) R⊙(cid:19)1/2(cid:18) M⋆ (cid:18) R⋆ M⊙(cid:19)−1/2 (1) Where K and ωsat are parameters for the efficiency of the coupling of the convective zone rotation to the wind, ~Jconv is the angular momentum of the convective zone, ~ωconv is the angular velocity of the convective zone and R⋆ are the radius and mass of the parent star R⊙ in solar units respectively. and M⋆ M⊙ In addition, angular momentum is exchanged between the radiative core and convective envelope by mass ex- change and by a torque driving the two zones toward solid body rotation: d ~Jconv dt !coup = − d ~Jrad dt !coup = = 1 τc Iconv ~Jrad − Irad ~Jconv Irad + Iconv − 2R2 rad 3Iconv (cid:18) dMrad dt (cid:19) ~ωconv (2) where ~Jrad is the angular momentum of the radiative core, Iconv and Irad are the moments of inertia of the con- vective and radiative zones respectively, Mrad and Rrad are the mass and outer radius of the radiative zone, and τc is a model parameter giving the timescale on which the core and the envelope converge to solid body rotation. Finally, we will use YREC tracks (Demarque et al. 2008) for the evolution of the stellar quantities (Iconv, Irad, R⋆, Mrad and Rrad). The combined orbital and stellar spin evolution de- scribed above was computed using a more general ver- sion of the POET code (Penev et al. 2014), which, among other things, allows following the evolution for systems in which the stellar spin is misaligned with the orbit. 5.2. Method Given values for all model parameters, in order to fully specify the evolution of the system, we need to choose appropriate boundary conditions. Clearly, the observed state of the system provides those, but if we wish to use the observed stellar spin to constrain Q′⋆ we must find independent spin boundary conditions. Fortunately, ro- tation periods for stars in young open clusters have been widely measured. Conveniently, as long as the stellar spin–down parameters are chosen to reproduce the ob- served evolution of stellar spin with age in open clusters, it makes very little difference which particular cluster we choose to start the evolution from. This is because for reasonable tidal dissipation rates, only a very tiny fraction of the orbital evolution occurs in the first few hundred Myrs, and as a result, the stellar spin evolu- tion hardly differs from that of an isolated star. This is very fortunate, since our results will not depend on the formation mechanism of HJs. Whether they form very early through disk migration, or much later through high–eccentricity migration, will have only a negligible effect on the final stellar spin. Example evolutions of HATS-18, using the nominal parameters from tables 3 and 4, adding the planet at ages 10 Myrs, 133 Myrs and 1 Gyr are shown in Fig. 9. In all cases, the evolution was started with the spin the star would have if it evolved only under the influence of angular momentum loss to stellar wind, and the initial orbital period of the planet was selected to reproduce the currently observed orbital period at the current age. We can see that, as expected, the effect of the age at which the planet migrates to its short period orbit on the stellar spin is utterly negligi- ble compared to the uncertainty of the measurement. In addition, Fig. 9 also shows that effect of assuming a frequency dependent tidal dissipation is relatively small, with even quite steep dependence on period (Q′⋆ ∝ P 2 HATS-18b 9 50 10 5 ] s y a d [ d o i r e P 1 0.5 no planet early planet nominal planet late planet Q′⋆ ∝ P 2 Q′⋆ ∝ P−2 10-2 10-1 Age [Gyr] 100 Fig. 9.- Example evolution of HATS-18 spin (solid lines) and HATS-18b orbital period (dashed lines) using the nominal mea- sured parameters for the system and log10(Q′ ⋆) = 7.3 at the ob- served tidal frequency of 0.46 days. The different lines correspond to adding the planet at ages 10 Myr (early planet), 133 Myrs (nom- inal planet) and 1 Gyr (late planet) as well as two additional as- sumptions for the frequency dependence of Q′ ⋆. The initial orbital period is chosen such that the present orbital period is reproduced at the present age of 4.2 Gyrs (open black circle). The initial stel- lar spin at the time the planet is added is that of a star evolving only under the influence of angular momentum loss (line labeled no planet) due to stellar wind. Regardless of the assumed planet migration age, the presently observed stellar spin period is repro- duced at the present system age (filled black circle), to much better than the measurement uncertainty. The different frequency scaling of Q′ ⋆ also have a relatively minor effect on the predicted stellar spin (both land within 2-sigma of the measured spin period). or Q′⋆ ∝ P −2) reproducing the currently observed stellar spin to within 2-sigma of the measured value, as long as log10(Q′⋆) = 7.3 at the observed tidal period for HATS-18 (0.46 days). For the constraint derived below, we used the combined spin periods for M 50 (Irwin et al. 2009) and the Pleiades (Hartman et al. 2010), since the two clusters are very close in age, have consistent period distributions, and together provide a large sample of stars for which the spin period has been measured. We assumed a starting age of 133 Myrs for all evolutions, close to the one estimated for the above clusters. In order to constrain the value of the tidal dissipa- tion parameter Q′⋆ defined above, fully accounting for the posterior distributions of the measured HATS-18 system properties, we will follow the following procedure: 1. select a random step from the converged DEMCMC chain, thus getting values for the present age of the HATS-18 system as well as the stellar and plane- tary masses and the stellar radius. 2. Randomly select one of from the Pleiades/M 50 with measured rotation period that has a mass within 0.1M⊙ of the randomly selected stellar mass above and use its spin period as the initial spin for the calculated evolution. the stars 3. Select a random value for log10(Q′⋆) from a uniform distribution in the range (5, 9). 4. Find an initial orbital period, such that starting the evolution at an age of 133 Myrs with the above parameters and evolving to the randomly selected present system age, results in the observed orbital period (the comparatively tiny uncertainty in the current orbital period is ignored). 5. Assume a normal distribution for the measured stellar spin period at the present age and evaluate the distribution at the resulting stellar spin period with the above evolution to get p(Q′⋆). Repeating the above steps multiple times allows us to build a cumulative distribution function (CDF) for log10(Q′⋆) by summing up all p(Q′⋆) values up to a par- ticular log10(Q′⋆). The number of iterations was chosen such that doubling their number did not result in signif- icant changes in the CDF. Finally, the entire procedure was repeated for a number of assumptions about the parameters of the spin model in order to investigate the sensitivity of the constraint to these parameters. In addition, even though planets around stars with surface convective zones appear to be well aligned with their host star's spin, it is possible that they form with a wide range of obliquities, which then de- cay on a timescale short compared to the tidal orbital de- cay for typical planets, but it may not be short compared to the orbital decay for HATS-18. In order to investigate the impact this could have on the results, we also con- sidered the most extreme possible case, of starting the star spinning in exactly the opposite direction to the or- bit and evolving to a presently assumed prograde state. The particular set of parameters considered is given in table 5. The "nominal" and "retrograde" models use the parameters for the stellar spin evolution which best fit the observed spin periods of open clusters (Irwin et al. 2007). An important point to note is that the change in parameter values away from the nominal model, for the other cases considered, do not represent actual un- certainties. In fact, all of these changes are in dramatic conflict with observations, demonstrating that very large changes in the models are required to make appreciable changes to the inferred Q′⋆ constraint. A more appro- priate treatment, which accounts properly for the shifts in the model parameters allowed by the cluster data, is beyond the scope of this paper, but the range of mod- els considered demonstrates the robustness of the results presented here. In addition to HATS-18, we carried out the steps out- lined in the previous section for WASP-19. This is an- other one of the three planetary systems whose measured semimajor axis to stellar radius ratio is consistent with being the smallest, and hence can be expected to have its host star spun up due to tidal dissipation. Indeed, it also seems to be spinning faster than expected for its age. In fact, Tregloan-Reed et al. (2013) observed the planet transiting in front of, what appears to be the same star spot, on two consecutive nights, which allowed them to measure WASP-19's spin period to be 11.76 ± 0.09 days, while the discovery paper (Hebb et al. 2010) quoted a photometrically detected rotation period of 10.5 ± 0.2 days. Neither of these periods is consistent with the isochronal constraint that the system is older than 1 Gyr (Hebb et al. 2010). In order to make the results from HATS-18 and WASP- 19 as comparable as possible, we used the same set of isochrones and the same fitting procedure to derive an isochronal age for WASP-19 of 8 ± 3 Gyr. Further, both 10 Penev et al. TABLE 5 The sets of assumptions for which constraints on log10(Q′ ⋆) were derived and the results for each system. Name K ⊙day2 M⊙R (rad2Gyr) ) 2 ( τc ωsat Initial spin HATS-18 Constraint WASP-19 Constraint (Myr) (rad/day) 68.2% Confidence Interval 68.2% Confidence Interval nominala retroa K = 0.11333b K = 0.22666b τc = 1b τc = 25b ωsat = 1.225b ωsat = 4.9b 0.17 0.17 0.11333 0.22666 0.17 0.17 0.17 0.17 10 10 10 10 1 25 10 10 2.45 2.45 2.45 2.45 2.45 2.45 1.225 4.9 prograde retrograde prograde prograde prograde prograde prograde prograde 7.2 - 7.4 6.8 - 7.1 7.3 - 7.6 7.1 - 7.3 7.0 - 7.3 7.3 - 7.6 7.2 - 7.4 7.2 - 7.4 6.5 - 6.9 6.2 - 6.6 6.6 - 7.1 6.4 - 6.8 6.3 - 6.8 6.6 - 7.0 6.5 - 6.9 6.5 - 6.8 a The values of K, τc and ωsat used for these models are best fit values to observations of stellar spin in open clusters of various ages. b The changes in stellar angular momentum loss parameters used in these models do not represent actual uncertainties, but are in fact much larger. All of these models are in clear conflict with observations. The particular values used were chosen to demonstrate the (lack of) sensitivity of the results to each parameter separately. stars have masses very close to solar, which means we do not need to worry about dependences of the vari- ous model parameters on the stellar mass. Finally, a proper DEMCMC fit to the WASP-19 observations is not available, so unlike for HATS-18, we simply assume the relevant parameters for WASP-19 from the liter- ature and use a Normal distribution with the quoted uncertainties. The particular values we employed were taken from Tregloan-Reed et al. (2013), and are con- sistent with the rest of the literature: M⋆ = 0.904 ± 0.045M⊙, R⋆ = 1.004 ± 0.018R⊙, Mpl = 1.114 ± 0.04MJ, and we adopted the Tregloan-Reed et al. (2013) stellar spin period of 11.76 ± 0.09 days and orbital period of P = 0.788840 ± 0.0000003 days. 5.3. Results In order to generate plots of the probability density functions (PDF) derived by the procedure described above, we fit a smoothing bicubic spline to the cumu- lative distribution with a tiny amount of smoothing in order to suppress numerical oscillations when taking the derivative. Fig. 10 shows the PDF derived for log10(Q′⋆) for HATS-18 and WASP-19 with the various models of table 5. The constraints obtained for Q′⋆ are given in the last column of that table. The confidence interval was derived by evaluating the inverse cumulative distri- bution function for log10(Q′⋆) at 15.87% and at 84.13%. Since most of the orbital decay happens at late times when the star is evolving only very slowly on the main sequence, it is a very good approximation to assume a non–evolving star with the present properties in the last Gyr or so of the evolution. As a result, as long as the star is started with the spin predicted by angular mo- mentum loss in the absence of a planet, the results are only very slightly sensitive to the exact stellar evolution models used. In particular this means that the exact stellar age determined by matching the evolution models to the present star, has only a very small effect on the results. Clearly, parametrizing tidal dissipation by a single number (Q′⋆ in our case) is a gross oversimplification of In reality Q′⋆ should depend on the physics involved. the stellar mass, the tidal frequency, and the stellar spin. This can affect the results in two ways: first, it could be one way to explain the different results obtained for the two systems, and second, even for a single system, the spin of the star and the tidal frequency evolve, thus different tidal dissipation will operate at different times during the system's past. However, for the two planetary systems considered all these parameters are currently al- most identical. Further, due to the strong dependence of the rate of orbital decay on the planet–star separa- tion, and the fact that angular momentum loss is faster for faster spinning stars, only the most recent part of the evolution of the systems matters (as demonstrated in Fig. 9). So even though the past spin histories of the two stars may have been somewhat different (due to the different planetary masses), this has a relatively small impact on the results. In addition, since the evolution is dominated by the latest stages, strictly speaking, the constraints derived here give the tidal quality factor for parameters close to the currently observed ones (a stellar mass of ap- proximately a 1M⊙, for orbital periods of approximately 0.8 days and for stellar spin periods of about 10 days). Finally, this also means that the formation mechanism for the planets is irrelevant for the derived constraints. While it is true that starting the orbital evolution later, if planet migration is delayed, can decrease the amount of angular momentum added to the star, this is a totally negligible effect (see Fig. 9). Disentangling the dependence of the tidal dissipation on some of these quantities may be possible by perform- ing similar analysis on a larger number of exoplanet sys- tems, ideally all currently known extremely short period ones. In addition, orbital circularization and spin syn- chronization in open cluster binaries is able to probe much longer orbital periods than is feasible with extra- solar planets. 6. DISCUSSION HATS-18 is an extreme short–period planet which is among the best targets for testing theories of planet– star interactions. like a num- ber of other extremely short period giant–planet hosts (e.g. WASP-19 above, WASP-103 (Gillon et al. 2014), OGLE-TR-113 (Bouchy et al. 2004)) appears to be spin- ning too fast for its age. HATS-18 is the best system In fact, the host star, HATS-18b 11 3.5 3.0 2.5 2.0 1.5 1.0 0.5 F D P 0.0 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 log10(Q′⋆ ) 2.0 1.5 F D P 1.0 0.5 0.0 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 log10(Q′⋆ ) 3.5 3.0 2.5 2.0 1.5 1.0 0.5 F D P 0.0 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 log10(Q′⋆ ) ωsat = 4. 9 ωsat = 1. 225 Kw = 0. 11333 Kw = 0. 22666 τc = 25Myr τc = 1Myr retro nominal Fig. 10.- Top: The PDF of log10(Q′ ⋆) from the HATS-18 sys- tem parameters. The various lines correspond to the models from table 5 with the models for which the angular momentum loss pa- rameters match the best fit of the stellar rotation rates in open clusters plotted with thicker lines. Middle: The same as the first panel, but using WASP-19 system parameters. Bottom: All the curves from the previous two plots together with HATS-18 plotted with solid lines, and WAPS-19 with dashed. to–date for constraining the stellar tidal dissipation by assuming that the extra stellar angular momentum was delivered by tidal decay of the orbit. In fact, we applied this method to the two exoplanet systems whose host stars should have been spun up the most, and which have very similar properties, to derive tight constraints on the stellar tidal quality factor at least in the regime appli- cable to those systems. In fact, if both of these planets are assumed to have formed in orbits well aligned with their parent star's spin, there is only a very narrow range around log10(Q′⋆) = 7 for which the present spin period of both stars is at least marginally consistent with the expected degree of spin–up. This tight constraint will also apply if planets form with a wide range of initial obliquities, but are quickly re-aligned by some process which operates on timescales short compared to the or- bital decay. On the other hand, if planets are assumed to form with a wide range of obliquities, and if at least for the extremely short periods of HATS-18 and WASP- 19, the timescale for orbital decay is shorter than any processes which tend to align the orbit with the stellar equator, it is plausible that WASP-19 started out in a well aligned orbit, while HATS-18 was significantly mis- aligned in which case, 6.5 < log10(Q′⋆) < 7. Clearly, a more systematic effort to analyze all suitable exoplanet systems and properly account for the stellar angular mo- mentum loss uncertainties is bound to yield very mean- ingful constraints on the stellar tidal dissipation, as well how it changes with various system properties. constraints do not match the recently suggested detection of orbital decay in WASP-12 (Maciejewski et al. 2016), which would correspond to a tidal quality factor of Q′⋆ = 2.5 × 105. However, the authors of that study point out that at present the ob- served period change is still marginally consistent with apsidal precession. Further, as we pointed out above, the tidal quality factor is not expected to be the same across different systems, and WASP-12 differs from both HATS-18 and WASP-19 in several important respects: it has a hotter star, with only a minimal surface convective zone, and it appears to be spinning significantly slower. Both of these properties are expected to impact the tidal dissipation. The same measurement is also within reach for HATS-18b. For example, after 28 years, the time of arrival of HATS-18b transits will have shifted by 60 s if Q′⋆ = 107 due to tidal orbital decay, thus making it feasible to measure. These As we argued in § 4, extremely short period planets like HATS-18 provide a fantastic laboratory to test a range of interactions between the planet and the star, and hence, expanding this sample is extremely valuable for the study of extrasolar planets. support Acknowledgements - Development of the HATSouth project was funded by NSF MRI grant NSF/AST- 0723074, operations have been supported by NASA grants NNX09AB29G and NNX12AH91H, and follow- up observations received partial support from grant NSF/AST-1108686. K.P. acknowledges support from NASA grants NNX13AQ62G and NNG14FC03C. G.B. acknowledges from the David and Lucile Packard Foundation, from NASA grants NNX13AJ15G, NNX14AF87G and NNX13AQ62G. J.H. acknowl- edges support from NASA grants NNX13AJ15G and NNX14AF87G. R.B. and N.E. are supported by CONICYT-PCHA/Doctorado Nacional. A.J. acknowl- edges from FONDECYT project 1130857, BASAL CATA PFB-06, and from the Ministry of Econ- omy, Development, and Tourism's Millenium Science Ini- tiative through grant IC120009, awarded to the Mille- nium Institute of Astrophysics, MAS. R.B. and N.E. ac- knowledge additional support from the Ministry of Econ- support 12 Penev et al. omy, Development, and Tourism's Millenium Science Ini- tiative through grant IC120009, awarded to the Mille- nium Institute of Astrophysics, MAS. V.S. acknowledges support from BASAL CATA PFB-06. This paper uses observations obtained with facilities of the Las Cum- bres Observatory Global Telescope. Work at the Aus- tralian National University is supported by ARC Lau- reate Fellowship Grant FL0992131. We acknowledge the use of the AAVSO Photometric All-Sky Survey (APASS), funded by the Robert Martin Ayers Sciences Fund, and the SIMBAD database, operated at CDS, Strasbourg, France. Operations at the MPG 2.2 m Telescope are jointly performed by the Max Planck Gesellschaft and the European Southern Observatory. G. B. wishes to thank the warm hospitality of Ad´ele and Joachim Cranz at the farm Isabis, supporting the operations and service missions of HATSouth. Agueros, M. A., Covey, K. R., Lemonias, J. J., et al. 2011, ApJ, Hartman, J. D., Bakos, G. ´A., Kov´acs, G., & Noyes, R. W. 2010, HATS-18b REFERENCES 13 740, 110 Albrecht, S., Winn, J. N., Johnson, J. A., et al. 2012, ApJ, 757, 18 Amard, L., Palacios, A., Charbonnel, C., Gallet, F., & Bouvier, J. 2016, A&A, 587, A105 Bakos, G. ´A., Torres, G., P´al, A., et al. 2010, ApJ, 710, 1724 Bakos, G. ´A., Csubry, Z., Penev, K., et al. 2013, PASP, 125, 154 Barnes, S. A., Weingrill, J., Fritzewski, D., & Strassmeier, K. G. 2016, ArXiv e-prints, 1603.09179 Bayliss, D., Zhou, G., Penev, K., et al. 2013, AJ, 146, 113 Bouchy, F., Pont, F., Santos, N. C., et al. 2004, A&A, 421, L13 Brahm et. al., . 2016 Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Claret, A. 2004, A&A, 428, 1001 Dawson, R. I., & Murray-Clay, R. A. 2013, ApJ, 767, L24 Delorme, P., Collier Cameron, A., Hebb, L., et al. 2011, MNRAS, 413, 2218 Demarque, P., Guenther, D. B., Li, L. H., Mazumdar, A., & Straka, C. W. 2008, Ap&SS, 316, 31 Dopita, M., Hart, J., McGregor, P., et al. 2007, Ap&SS, 310, 255 Fossati, L., Haswell, C. A., Froning, C. S., et al. 2010, ApJ, 714, L222 Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, 81 Gallet, F., & Bouvier, J. 2015, A&A, 577, A98 Gillon, M., Anderson, D. R., Collier-Cameron, A., et al. 2014, A&A, 562, L3 Ginzburg, S., & Sari, R. 2015, ApJ, 803, 111 Hansen, B. M. S., & Barman, T. 2007, ApJ, 671, 861 MNRAS, 408, 475 Hartman, J. D., Gaudi, B. S., Pinsonneault, M. H., et al. 2009, ApJ, 691, 342 Hartman, J. D., Bakos, G. ´A., B´eky, B., et al. 2012, AJ, 144, 139 Haswell, C. A., Fossati, L., Ayres, T., et al. 2012, ApJ, 760, 79 Hebb, L., Collier-Cameron, A., Triaud, A. H. M. J., et al. 2010, ApJ, 708, 224 Ida, S., & Lin, D. N. C. 2008, ApJ, 673, 487 Irwin, J., Aigrain, S., Bouvier, J., et al. 2009, MNRAS, 392, 1456 Irwin, J., Hodgkin, S., Aigrain, S., et al. 2007, MNRAS, 377, 741 Jord´an, A., Brahm, R., Bakos, G. ´A., et al. 2014, AJ, 148, 29 Kaufer, A., & Pasquini, L. 1998, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 3355, Optical Astronomical Instrumentation, ed. S. D'Odorico, 844–854 Kov´acs, G., Bakos, G., & Noyes, R. W. 2005, MNRAS, 356, 557 Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369 Kov´acs, G., Hartman, J. D., Bakos, G. ´A., et al. 2014, MNRAS, 442, 2081 Lai, D. 2012, MNRAS, 423, 486 Maciejewski, G., Dimitrov, D., Fern´andez, M., et al. 2016, A&A, 588, L6 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Penev, K., Jackson, B., Spada, F., & Thom, N. 2012, ApJ, 751, 96 Penev, K., Zhang, M., & Jackson, B. 2014, PASP, 126, 553 Penev, K., Bakos, G. ´A., Bayliss, D., et al. 2013, AJ, 145, 5 ter Braak, C. J. F. 2006, Statistics and Computing, 16, 239 Tregloan-Reed, J., Southworth, J., & Tappert, C. 2013, MNRAS, 428, 3671 Yi, S., Demarque, P., Kim, Y.-C., et al. 2001, ApJS, 136, 417
1510.03144
1
1510
2015-10-12T05:25:29
The color-magnitude distribution of small Jupiter Trojans
[ "astro-ph.EP" ]
We present an analysis of survey observations targeting the leading L4 Jupiter Trojan cloud near opposition using the wide-field Suprime-Cam CCD camera on the 8.2 m Subaru Telescope. The survey covered about 38 deg$^2$ of sky and imaged 147 fields spread across a wide region of the L4 cloud. Each field was imaged in both the $g'$ and the $i'$ band, allowing for the measurement of $g-i$ color. We detected 557 Trojans in the observed fields, ranging in absolute magnitude from $H=10.0$ to $H = 20.3$. We fit the total magnitude distribution to a broken power law and show that the power-law slope rolls over from $0.45\pm 0.05$ to $0.36^{+0.05}_{-0.09}$ at a break magnitude of $H_{b}=14.93^{+0.73}_{-0.88}$. Combining the best-fit magnitude distribution of faint objects from our survey with an analysis of the magnitude distribution of bright objects listed in the Minor Planet Center catalog, we obtain the absolute magnitude distribution of Trojans over the entire range from $H=7.2$ to $H=16.4$. We show that the $g-i$ color of Trojans decreases with increasing magnitude. In the context of the less-red and red color populations, as classified in Wong et al. 2014 using photometric and spectroscopic data, we demonstrate that the observed trend in color for the faint Trojans is consistent with the expected trend derived from extrapolation of the best-fit color population magnitude distributions for bright catalogued Trojans. This indicates a steady increase in the relative number of less-red objects with decreasing size. Finally, we interpret our results using collisional modeling and propose several hypotheses for the color evolution of the Jupiter Trojan population.
astro-ph.EP
astro-ph
Draft version October 18, 2018 Preprint typeset using LATEX style emulateapj v. 5/2/11 Keywords: minor planets, asteroids: general THE COLOR-MAGNITUDE DISTRIBUTION OF SMALL JUPITER TROJANS * Ian Wong and Michael E. Brown Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA; [email protected] Draft version October 18, 2018 ABSTRACT We present an analysis of survey observations targeting the leading L4 Jupiter Trojan cloud near opposition using the wide-field Suprime-Cam CCD camera on the 8.2 m Subaru Telescope. The survey covered about 38 deg2 of sky and imaged 147 fields spread across a wide region of the L4 cloud. Each field was imaged in both the g(cid:48) and the i(cid:48) band, allowing for the measurement of g − i color. We detected 557 Trojans in the observed fields, ranging in absolute magnitude from H = 10.0 to H = 20.3. We fit the total magnitude distribution to a broken power law and show that the power-law slope rolls over from 0.45 ± 0.05 to 0.36+0.05−0.09 at a break magnitude of Hb = 14.93+0.73−0.88. Combining the best-fit magnitude distribution of faint objects from our survey with an analysis of the magnitude distribution of bright objects listed in the Minor Planet Center catalog, we obtain the absolute magnitude distribution of Trojans over the entire range from H = 7.2 to H = 16.4. We show that the g − i color of Trojans decreases with increasing magnitude. In the context of the less-red and red color populations, as classified in Wong et al. (2014) using photometric and spectroscopic data, we demonstrate that the observed trend in color for the faint Trojans is consistent with the expected trend derived from extrapolation of the best-fit color population magnitude distributions for bright catalogued Trojans. This indicates a steady increase in the relative number of less-red objects with decreasing size. Finally, we interpret our results using collisional modeling and propose several hypotheses for the color evolution of the Jupiter Trojan population. 1. INTRODUCTION Residing at a mean heliocentric distance of 5.2 AU, the Jupiter Trojans are asteroids that share Jupiter's or- bit around the Sun and are grouped into two extended swarms centered around the stable L4 and L5 Lagrangian points. Estimates of the size of this population indi- cate that the Trojans are comparable in number to main belt asteroids of similar size (Szab´o et al. 2007; Naka- mura & Yoshida 2008). Explaining the origin and evo- lution of this significant population of minor bodies is crucial for understanding the formation and dynamical history of the Solar System. While early theories posited that the Trojans could have formed out of the body of planetesimals and dust in the immediate vicinity of a growing Jupiter (Marzari & Scholl 1998), later studies re- vealed that such in situ formation is not consistent with the observed total mass and broad inclination distribu- tion. An alternative theory suggests that the Trojans formed at large heliocentric distances out of the same body of material that produced the Kuiper Belt (Mor- bidelli et al. 2005). Subsequent migration of the gas gi- ants triggered a period of chaotic dynamical alterations in the outer Solar System, during which the primordial trans-Neptunian planetesimals were disrupted (Tsiganis et al. 2005; Nesvorn´y & Morbidelli 2012). It is hypoth- esized that a fraction of these objects were scattered in- wards and captured by Jupiter as Trojan asteroids. A detailed study of the size distribution of Trojans promises to shed light on the relationships between the Trojans and other minor body populations in the outer Solar System, and more broadly, constrain models of late Solar System evolution. The size distribution, or * Based on data collected at Subaru Telescope, which is oper- ated by the National Astronomical Observatory of Japan. as a proxy, the magnitude distribution, offers signifi- cant insight into the nature of the Trojan population, as it contains information about the conditions in which the objects were formed as well as the processes that have shaped the population since its formation. Previ- ous studies of the Trojan magnitude distribution have largely focused on objects larger than ∼10-20 km in di- ameter (e.g., Jewitt et al. 2000; Szab´o et al. 2007), al- though the advent of larger telescopes and improved in- struments has presented the opportunity to carry out surveys of smaller Trojans. Yoshida & Nakamura (2005) and Yoshida & Nakamura (2008) presented the first mag- nitude distribution for small L4 and L5 Trojans as part of a small survey, with several dozen objects detected down to sizes of ∼2 km. The detection of many more faint objects promises to expand our understanding of the small Trojan population. Little is known about the composition and surface properties of Trojans. Both large-scale and targeted ob- servational studies over the past few decades have re- vealed a population that is notably more homogeneous than the main belt asteroids, with low albedos and spec- tral slopes ranging form neutral to moderately red (e.g., Szab´o et al. 2007; Roig et al. 2008; Fern´andez et al. 2009). Meanwhile, visible and near-infrared spectroscopy has been unable to detect any incontrovertible spectral features (e.g., Dotto et al. 2006; Fornasier et al. 2007; Yang & Jewitt 2007; Melita et al. 2008; Emery et al. 2011). However, recent work has uncovered bimodalities in the distribution of various spectral properties, such as spectral slope in the visible (Szab´o et al. 2007; Roig et al. 2008; Melita et al. 2008) and the near-infrared (Emery et al. 2011). In Wong et al. (2014), the data from these previous studies were compiled and shown to be indicative of the existence of two color populations -- 2 the so-called red and less-red populations. The magni- tude distributions of these two populations are distinct, differing especially in the power-law distribution slopes of objects smaller than ∼50 km. Several hypotheses for the origin of this discrepancy have been posited, includ- ing different source regions for red and less-red Trojans, conversion of red objects to less-red fragments upon col- lision, and space weathering effects (Melita et al. 2008; Wong et al. 2014). By extending the analysis of Trojan colors to smaller objects, we hope to better understand the underlying processes behind the color dichotomy and the differing magnitude distributions of the color popu- lations. In this paper, we present the results of our survey of small Trojans in the leading L4 cloud. We detected over 550 Trojans and measured their brightness in two filters, from which their magnitudes and colors were computed. We calculate the best-fit curve describing the total mag- nitude distribution down to a limiting absolute magni- tude of H = 16.4. In addition, we present the first anal- ysis of the color distribution of faint Trojans to date and compare the measured trends with previously-published results for brighter Trojans (Wong et al. 2014). Lastly, we use collisional modeling to interpret the derived mag- nitude and color distributions. 2. OBSERVATIONS Observations of the L4 Trojan cloud were carried out on UT 2014 February 27 and 28 at the 8.2 m Subaru Telescope situated atop Mauna Kea, Hawaii. Using the Suprime-Cam instrument -- a mosaic CCD camera con- sisting of ten 2048 × 4096 pixel CCD chips that covers a 34(cid:48)×27(cid:48) field of view with a pixel scale of 0.20(cid:48)(cid:48) (Miyazaki et al. 2002) -- we observed 147 fields, corresponding to a total survey area of 37.5 deg2. To detect moving objects as well as obtain color photometry, we imaged each field four times -- twice in the g(cid:48) filter (λeff = 480.9 nm) and twice in the i(cid:48) filter (λeff = 770.9 nm). The average time interval between epochs is about 20 minutes for images in the same filter, and about 30 minutes for images in dif- ferent filters. We chose an exposure time of 60 seconds for all images to optimize survey depth and coverage. The resulting average observational arc for each moving object is roughly 70 minutes. The on-sky positions of the 147 observed Suprime- Cam fields are shown in Figure 1. The surveyed re- gion was divided into blocks of 10 − 12 observing fields, which we imaged in the filter order g(cid:48) − g(cid:48) − i(cid:48) − i(cid:48), or in reverse. Blocks observed toward the beginning of each night targeted the trailing edge of the L4 Tro- jan cloud, while blocks observed later in the night are concentrated closer to the peak of the spatial distribu- tion. To place the observed field locations in the con- text of the L4 point and the spatial extent of the lead- ing Trojan cloud, we use the empirical L4 Trojan num- J ) ∼ ber density model from Szab´o et al. (2007): n(λ(cid:48) J − 60◦)2/2σ2 exp [−(λ(cid:48) β] with σλ = 14◦ and σβ = 9◦, where λ(cid:48) J are respectively the helio- centric ecliptic longitude and latitude relative to Jupiter. In Figure 1, the 50%, 10%, and 5% relative number den- sity contours are shown in geocentric longitude-latitude space (λ, β); the position of the L4 point during the time of our observations is at around (λ, β) = (170◦, 0◦). We λ] × exp [−β(cid:48)2 J and β(cid:48) J , β(cid:48) J /2σ2 see that the majority of the observed fields lie in regions with predicted Trojan number densities at least 50% of the peak value. 2.1. Moving object detection After bias-subtracting the images, we flat-fielded them using the twilight flats taken at the start of the first night of observation. For every chip image (10 per exposure, 40 per observed field, 5880 for the entire survey), we cal- culated the astrometric solution by first creating a pixel position catalog of all the bright sources in the image. This was done using Version 2.11 of SExtractor (Bertin & Arnouts 1996), with the threshold for source detection set at a high value (typically 30 or higher, depending on the seeing, in units of the estimated background stan- dard deviation). The source catalog was then passed to Version 1.7.0 of SCAMP (Bertin 2006), which matches objects in the source catalog with those in a reference catalog of stars and computes the astrometric projection parameters for each chip image; we used the 9th Data Release of the Sloan Digital Sky Survey (SDSS DR9) as our reference catalog. In order to assess the quality of the astrometric solution from SCAMP, we compared the corrected on-sky position of stars in each image with the position of matched reference stars and found typi- cal residual RMS values much less than 0.1(cid:48)(cid:48). Likewise, we calculated the position scatter between matched stars from pairs of images taken in the same observing field and found typical RMS values less than 0.05(cid:48)(cid:48). Next, we passed the distortion-corrected images through SExtractor again, this time setting the detec- tion threshold at 1.2 times the background standard de- viation; we also required detected sources to consist of at least two adjacent pixels with pixel values above the detection threshold. These conditions were chosen to minimize the detection of faint non-astrophysical sources in the background noise as well as the loss of possible moving objects of interest. The resulting list contains the right ascension (RA) and declination (Dec) of all de- tected objects within each image. To search for mov- ing object candidates, we fit orbits through sets of four source positions, one from each of the four images in an observing field, using the methods described in Berstein & Khushalani (2000). A source was flagged as a moving object candidate if the resulting χ2 value from the fit was less than 10. We reduced the number of non-Trojan mov- ing object candidates flagged in this procedure by only considering sets of source positions consistent with ap- parent sky motion v less than 25(cid:48)(cid:48)/hr. For comparison, typical sky motions of known Trojans during the time of our observations lie in the range 14 ≤ v < 22(cid:48)(cid:48)/hr. Each moving object candidate was verified by aligning and blinking 100 × 100 pixel stamps clipped from each of the four images in the vicinity of the object. We re- jected all non-asteroidal moving object candidates (e.g., sets of four sources that were flagged by the previous orbit-fitting procedure, but that included cosmic ray hits and/or anomalous chip artifacts). We also rejected as- teroids that passed in front of background stars, coin- cided with a cosmic ray hit in one or more image, or otherwise traversed regions on the chip that would result in unreliable magnitude measurements. These objects numbered around 10, of which only 2 had on-sky posi- tions and apparent sky motions consistent with Trojans 3 Fig. 1. -- Locations of the 147 observed Subaru Suprime-Cam fields, projected in geocentric ecliptic longitude-latitude space (blue diamonds). The size of the diamonds corresponds to the total field of view of each image. The positions of numbered Trojans and non- Trojans during the time of our observations with apparent sky motions in the range 14 ≤ v < 22(cid:48)(cid:48)/hr are indicated by black squares and red dots, respectively. The solid, dashed, and dotted curves denote respectively the approximate 50%, 10%, and 5% relative density contours in the sky-projected L4 Trojan distribution (Szab´o et al. 2007). (see Section 2.2). Therefore, the removal of these objects from our Trojan data set is not expected to have any sig- nificant effect on the results of our analysis. Through the procedures described above, we arrived at an initial set of 1149 moving objects. 2.2. Selecting Trojans Observations were taken when the L4 point was near opposition, where the apparent sky motion of an object, v, is roughly inversely related to the heliocentric dis- tance. Since the short observational arc of ∼70 minutes prevented an accurate determination of the heliocentric distance from orbit-fitting, we resorted to using primar- ily sky motion to distinguish Trojans from Main Belt Asteroids and Hildas. Figure 2 shows the distribution of apparent velocities in RA and Dec for all numbered minor bodies as calculated from ephemerides generated by the JPL HORIZONS system for UT 2014 February 27 12:00 (roughly the middle of our first night of observa- tion). We have shown only objects with on-sky positions in the range of our observing fields (130◦ ≤ λ < 190◦ and −8◦ ≤ β < 8◦). A large majority of the objects with v < 20(cid:48)(cid:48)/hr are Trojans, while the relative por- tion of non-Trojan contaminants increases rapidly in the range 20(cid:48)(cid:48)/hr] ≤ v < 22(cid:48)(cid:48)/hr. We made an initial cut in angular velocity to consider only objects in the range 14 ≤ v < 22(cid:48)(cid:48)/hr. The contamination rate varies across the surveyed area as a function of apparent sky motion, as well as position on the sky. We carried out a more detailed contamina- tion analysis by applying a rough grid in the space of v, λ, and β spanning the range 14 ≤ v < 22(cid:48)(cid:48)/hr, 130◦ ≤ λ < 190◦, and −8◦ ≤ β < 8◦. We then binned all numbered objects contained in the JPL HORIZONS system into this three-parameter space and computed the Trojan fraction γ ≡ [numbered Trojans]/[all num- bered objects] in each bin. For bins containing no num- bered objects, we assigned a value γ = 0. Figure 3 shows the results of this analysis. We note that while the non-Trojan contamination rate over the whole sur- veyed area among objects with apparent sky motions be- tween 20 and 22(cid:48)(cid:48)/hr is high, there are regions in the sky where the Trojan fraction is 1. Similarly, for objects with v < 20(cid:48)(cid:48)/hr, there are regions near the edges of the ecliptic longitude space covered by our survey where the non-Trojan contamination rate is very high. In this paper, our operational Trojan data set includes 130140150160170180190λ [deg]864202468β [deg] 4 (cid:16)(cid:80)2 (cid:16)(cid:80)2 to be the error-weighted mean of the apparent magni- tudes calculated from the two g(cid:48) Suprime-Cam images, i.e., g = , with the corresponding uncertainty in the g magnitude defined by σ2 ; the i magnitude and uncer- tainty of each Trojan were defined analogously. k=1 mg,k/σ2 g,k k=1 1/σ2 g,k / k=1 1/σ2 g,k g = 1/ (cid:17) (cid:17) (cid:16)(cid:80)2 (cid:17) As mentioned previously in Section 2.2, orbit-fitting over the short observational arcs prevented us from pre- cisely determining the orbital parameters of the detected objects. Therefore, we did not directly convert the ap- parent magnitudes to absolute magnitudes using the best-fit heliocentric distances. Instead, we considered the difference between apparent V-band magnitude and ab- solute magnitude H of known Trojans, as computed by the JPL HORIZONS system for the time of our obser- vations. By fitting a linear trend through the computed apparent sky motions v as a function of the magnitude difference values, we obtained an empirical conversion between the sky motion and V − H. Since V − H de- pends also on the viewing geometry, we divided the eclip- tic longitude range 130◦ ≤ λ < 190◦ into 5◦ bins and de- rived linear fits separately for known Trojans within each bin. We calculated the apparent V-band magnitudes of the detected Trojans from our survey using the empirical mean colors g − r = 0.55 and V − r = 0.25 reported in Szab´o et al. (2007). Then, we translated these V values to H using the measured sky motions and our empirical V − H conversions. From the absolute magnitude, the diameter of a Tro- jan can be estimated using the relation D = 1329 × 10−H/5/ pv, where D is in units of kilometers, and we assumed a uniform geometric albedo of pv = 0.04 (Fern´andez et al. 2009). The brightest Trojan we de- tected in our survey has H = 10.0, corresponding to a diameter of 66.5 km, while the faintest Trojan has H = 20.3, corresponding to a diameter of 0.6 km. √ 2.4. Data completeness An analysis of the magnitude distribution of a popula- tion can be severely affected by detection incompleteness. Varying weather conditions during ground-based surveys lead to significantly different detection completeness lim- its for observations taken at different times, resulting in a non-uniform data set and biases in the overall mag- nitude distribution. Since we required a moving object candidate to be detected in all four images taken in an observed field, the epoch with the highest seeing in each set of four exposures determined the threshold magni- tude to which the survey was sensitive in that particular field. Here we set the seeing value of each chip image to be the median of the full-width half-max of the point- spread function for all non-saturated stars, as computed by SExtractor. If the highest seeing value across a four- image set was high, then the object detection pipeline would have missed many of the fainter objects positioned within the field. This is because the signal-to-noise of ob- jects of a given magnitude located on the worst image of the set would be significantly lower than that of the same objects located on an image taken at good seeing. As a result, the magnitude distribution of objects detected in an observing field with high-seeing images would be strongly biased toward brighter objects, and when com- Fig. 2. -- Distribution of apparent RA and Dec velocities for numbered Trojans (black squares) and non-Trojans (red dots) with positions in the range 130◦ ≤ λ < 190◦ and −8◦ ≤ β < 8◦. The dotted curves denote the v = 14(cid:48)(cid:48)/hr and v = 22(cid:48)(cid:48)/hr contours, which separate the Trojans from the majority of non-Trojan minor bodies. only those objects detected in regions of the (λ, β,v) space with Trojan fraction γ = 1 (i.e., zero expected contamination). The resulting data set contains 557 Tro- jans. Choosing a slightly more lenient Trojan selection criterion (e.g., including objects detected in bins that do not contain numbered objects, where the Trojan fraction is technically unknown) was not found to significantly affect the main results of our magnitude and color dis- tribution analysis. 2.3. Photometric calibration The apparent magnitude m of an object detected in our images is given by m = m0 − 2.5 log10(f ) = m0 + ms, where f is the measured flux, m0 is the zero-point mag- nitude of the corresponding chip image, and the survey magnitude has been defined as ms ≡ −2.5 log10(f ). The flux of each object was calculated by SExtractor through a fixed circular aperture with a diameter of 5 pixels. We computed the zero-point magnitude for each chip image by matching all bright, non-saturated sources detected by SExtractor to reference stars in the SDSS DR9 cata- log and fitting a line with slope one through the points (ms, m). The maximum allowed position difference for a match between image and reference stars was set at 0.5(cid:48)(cid:48), which resulted in an average of ∼150 matched stars per chip image. We used Sloan g-band and i-band reference star magnitudes to calibrate images taken in the g(cid:48) and i(cid:48) filters of Suprime-Cam, respectively. Consequently, the calculated apparent magnitudes of our detected Trojans are effectively Sloan g and i magnitudes, which greatly facilitates the comparison of our computed colors with those derived for Trojans in the SDSS Moving Object Catalog (see Section 3.2). Both the error in the zero-point magnitude σ0 and the error in the measured flux σf (reported by SEx- tractor) contribute to the error in the apparent magni- tude σ, which we calculate using standard error propaga- 0 + (2.5(σf /f )/ ln(10))2. We set the g magnitude of each Trojan detected in our survey tion methods: σ =(cid:112)σ2 252015105vRA ["/hr]505101520vDEC ["/hr] 5 Fig. 3. -- Map of Trojan fraction γ among numbered minor bodies at various locations in the space of ecliptic longitude, ecliptic latitude, and apparent sky motion during the time of our observations. Regions in dark red have 0% predicted contamination by non-Trojans, and our operational Trojan data set includes only objects detected in observed fields located within these regions. magnitude distribution. Figure 4 shows the highest measured seeing for each group of four chip images taken in an observing field, plotted with respect to the time of first epoch. A large portion of our first night of observation was plagued by very high seeing - as high as 3.0(cid:48)(cid:48) at times. We chose a cutoff seeing value of 1.2(cid:48)(cid:48) and defined a filtered Tro- jan set containing only Trojans detected in images with seeing below this value. We used signal-to-noise to establish a limiting mag- nitude for our magnitude distribution analysis. Instead of attempting to generate an empirical model of the de- tection efficiency for the faintest objects, we elected to stipulate a conservative S/N threshold of 8. We defined the upper magnitude limit of our filtered Trojan data set to be the magnitude of the brightest object (detected at seeing less than 1.2(cid:48)(cid:48)) with S/N < 8. The limiting mag- nitude was determined to be H = 16.4. The final filtered Trojan set contains 150 objects, and when fitting for the total magnitude distribution, we assume that this set is complete (i.e., is not characterized by any size-dependent bias). 2.5. Trojan colors The color c of each Trojan is defined as the difference between the g and i magnitudes: c ≡ g − i. When cal- Fig. 4. -- Seeing of the worst image (highest seeing) in every chip for each set of four exposures in the 147 observed fields, plotted as a function of the first exposure time. The dotted line indicates the cutoff at 1.2(cid:48)(cid:48) seeing - objects in fields with higher seeing (red triangles) are not included in the filtered Trojan set. bined with the full body of data, would affect the overall JD - 2456715 6 Fig. 5. -- Comparison of the measured standard deviation error in the difference of g magnitudes, σ∆g, binned in 0.5 mag intervals (blue squares) with the corresponding values from our empirical model combining the measured photometric magnitude errors with a constant contribution from asteroid rotation (green squares). The binned medians of the photometric errors only are denoted by black crosses. The agreement between the measured and modeled σ∆g values shows that the assumption of a magnitude-invariant contri- bution to the dispersion in ∆g from asteroid rotation is good. culating the uncertainty of each color measurement, we must consider the effect of asteroid rotation in addition to the contribution from photometric error. The oscilla- tions in apparent amplitude seen in a typical asteroidal rotational light curve arise from the non-spherical shape of the object and peak twice during one full rotation of the asteroid (see, for example, the review by Pravec et al. 2002). The average observational arc of ∼70 minutes for each object may correspond to a significant fraction of a characteristic light curve oscillation period, which can consequently lead to a large variation in the apparent brightness of an object across the four epochs. In the absence of published light curves for faint Tro- jans, we must develop a model of the rotational contri- bution to the color uncertainty in order to accurately de- termine the total uncertainty of our color measurements. We took advantage of the fact that we obtained two de- tections of an object in each filter to estimate the effect of asteroid rotation empirically. We considered the dif- ference between the two consecutive g(cid:48)-band magnitude measurements ∆g ≡ g1 − g2. The standard deviation in ∆g values, σ∆g, contains a contribution from the pho- tometric error given by the quadrature sum of the indi- vidual magnitude uncertainties as defined in Section 2.3: g,2. Figure 5 compares the stan- σ∆g,phot = dard deviation of ∆g values measured in 0.5 mag bins (blue squares) to the corresponding error contribution from the photometric uncertainty only, σ∆g,phot (black crosses). g,1 + σ2 σ2 (cid:113) From the plot, it is evident that an additional con- tribution attributable to rotation is needed to account for the uncertainty in ∆g seen in our data. We observe that the discrepancy between σ∆g and the photometric error contribution is large at low magnitudes, where the photometric error is small, and decreases with increas- (cid:113) σ2 ∆g,phot + σ2 ing magnitude and increasing photometric error. This overall trend suggests that the total error in ∆g can be empirically modeled as a quadrature sum of the photo- metric contribution and the rotational contribution, i.e., σ∆g = ∆g,rot, where the rotational contri- bution is magnitude-invariant. We found that a rota- tional contribution of σ∆g,rot ∼ 0.08 gives a good match with the measured standard deviation in ∆g, and like- wise for ∆i. In Figure 5, the green squares show the binned ∆g standard deviation values calculated from our empirical model combining both photometric and rota- tional contributions. The rotational contribution to the color uncertainty was estimated as a quadrature sum of the rotational con- tributions to ∆g and ∆i: σc,rot ∼ 0.11 mag. The total color uncertainty is therefore given by: (cid:113) σc = g + σ2 σ2 i + σ2 c,rot, (1) where σg and σi are the errors in the g and i magnitudes derived from the flux and calibrated zero-point magni- tude errors only (see Section 2.3). We note that the effects of bad seeing and detection incompleteness discussed in Section 2.4 are not expected to affect the bulk properties of the Trojan color distribu- tion (e.g., mean color), except at the faintest magnitudes, where there is a slight bias toward redder objects. This is because we used the measured g magnitudes to convert to absolute magnitudes H. For objects with a given g mag- nitude, redder objects are slightly brighter in the i(cid:48)-band (lower i magnitude) than less-red objects. The predicted effect is small, since the characteristic difference in g − i color between objects in the red and less-red populations is only ∼0.15 mag (see Section 3.2). In our color anal- ysis, we strove to avoid this small color bias by limiting the analysis to objects with H magnitudes less than 18, thereby removing the faintest several dozen objects from consideration. 3. ANALYSIS In this section, we first present our analysis of the total Trojan magnitude distribution, using both data for faint Trojans from our Subaru survey and data for brighter L4 Trojans listed in the Minor Planet Center catalog. Next, the distribution of measured g − i colors and its magnitude dependence studied in the context of the less- red and red color populations described in Wong et al. (2014). 3.1. Total magnitude distribution The cumulative magnitude distribution N (H) of all ob- jects in the filtered Trojan set (Section 2.4) as a function of H magnitude is shown in Figure 6. The main feature of the magnitude distribution is a slight rollover in slope to a shallower value at around H ∼ 15. Following pre- vious analyses of the magnitude distribution of Trojans (e.g., Jewitt et al. 2000), we fit the differential magnitude distribution, Σ(H) = dN (M )/dH, with a broken power , measured , model 7 Fig. 7. -- Cumulative magnitude distribution of all L4 Trojans contained in the Minor Planet Center (MPC) catalog brighter than H = 12.3 (white squares; corrected for incompleteness in the range H = 11.3 − 12.3 following the methodology of Wong et al. 2014) and the cumulative magnitude distribution of the filtered Trojan set from our Subaru observations, approximately scaled to reflect the true overall number and binned by 0.1 mag (green dots). The error bars on the scaled Subaru Trojan magnitude distribution (in green) denote the 95% confidence bounds derived from scaling the Poisson errors on the Subaru survey data. The uncertainties on the corrected MPC catalog Trojan magnitude distribution in the range H = 11.3 − 12.3 are much smaller than the points. The best-fit broken power law curves describing the MPC and Subaru data are overplotted (solid black and dashed red lines, respectively), with 0, α(cid:48) the power law slopes (in the text: α(cid:48) 1, α1, and α2) indicated for their corresponding magnitude regions. broken power law to the corresponding magnitude dis- tribution, we set η = 1. We used an affine-invariant Markov Chain Monte Carlo (MCMC) Ensemble sampler (Foreman-Mackey et al. 2013) with 50,000 steps to estimate the best-fit parame- ters and corresponding 1σ uncertainties. The magnitude distribution of the filtered Trojan set is best-fit by a bro- ken power law with parameter values α1 = 0.45 ± 0.05, α2 = 0.36+0.05−0.09, H0 = 11.39+0.31−0.37, and Hb = 14.93+0.73−0.88. The cumulative magnitude distribution for this best-fit model is plotted in Figure 6 as a solid black line. We note that while the difference between the two power-law slopes is small, it is statistically significant: Marginaliz- ing over the slope difference ∆α = α1 − α2, we obtained ∆α = 0.10+0.09−0.08, which demonstrates that the difference between the two power-law slopes is distinct from zero at the 1.25σ level. We also fit the magnitude distribution of known bright Trojans contained in the Minor Planet Center (MPC) catalog by repeating the analysis described in Wong et al. (2014), this time including only L4 Trojans. Fitting the distribution of bright L4 Trojans likewise to a broken power law of the form shown in Eq. (2), we obtained a best-fit distribution function with α(cid:48) 0 = 0.91+0.19−0.16, 1 = 0.43 ± 0.02, H(cid:48) α(cid:48) 0 = 7.22+0.24−0.25, and H(cid:48) b = 8.46+0.49−0.54. We have denoted the best-fit parameters for the MPC Trojan magnitude distribution with primes to distinguish them from the best-fit parameters for the Subaru Tro- jans; the subscripts on the slope parameters α are ad- justed to reflect the magnitude ranges they correspond Fig. 6. -- Cumulative absolute magnitude distribution of the fil- tered Trojan set from our Subaru observations, binned by 0.1 mag (green dots). The best-fit broken power law curve describing the distribution is overplotted (solid black line), with the power law slopes indicated. The dashed line is an extension of the α1 = 0.45 slope and is included to make the slope rollover more discernible. law Σ(α1, α2,H0, HbH) (cid:26)10α1(H−H0), = 10α2H+(α1−α2)Hb−α1H0 , for H < Hb for H ≥ Hb , (2) where the power law slope for brighter objects α1 changes to a shallower faint-end slope α2 at a break magnitude Hb. The parameter H0 defines the threshold magnitude for which Σ(H0) = 1. We fit the model curve in Eq. (2) to our filtered Tro- jan magnitude distribution using a maximum likelihood method similar to the one used in Fraser et al. (2008) for their study of Kuiper belt objects. We defined a like- lihood function L that quantifies the probability that a random sampling of the model distribution will yield the data: L(α1, α2, H0, HbHi) ∝ e−N(cid:89) Pi. (3) i Here, Hi is the H magnitude of each detected Trojan and Pi = Σ(α1, α2, H0, HbHi) is the probability of detecting an object i with magnitude Hi given the underlying dis- tribution function Σ. N is the total number of objects detected in the magnitude range under consideration and is given by (cid:90) Hmax −∞ N = η(H)Σ(α1, α2, H0, HbH) dH, (4) where in the case of our filtered Trojan set we have Hmax = 16.4. We have included the so-called "efficiency" function η(H) that represents an empirical model of the incompleteness in a given data set and ensures that the best-fit distribution curve corrects for any incomplete- ness in the data. Our filtered Trojan data set was de- fined to remove incompleteness contributions from both bad seeing and low signal-to-noise, so when fitting the 0.45±0.05 0.36 +0.05 -0.09 0.91 +0.19 -0.16 0.43±0.02 0.36 +0.05 -0.09 0.45±0.05 Subaru (scaled) Catalogued L4 Trojans 8 to in the overall distribution. The cumulative magnitude distribution of MPC L4 Trojans through H = 12.3 (cor- rected for catalog incompleteness following the method- ology described in Wong et al. 2014) is shown in Figure 7 along with the best-fit curve. Earlier studies of L4 Trojans in this size range have reported power law slopes that are consistent with our values: Yoshida & Nakamura (2008) fit MPC L4 Trojans with magnitudes in the range 9.2 < H < 12.3 and de- rived a slope of 0.40±0.02. We note that their fit did not take into account the incompleteness in the MPC cata- log, which we estimated in Wong et al. (2014) to begin at H = 11.3. Therefore, the slope value in Yoshida & Nakamura (2008) is somewhat underestimated. Jewitt et al. (2000) carried out a survey of L4 Trojans and com- puted slopes of 0.9 ± 0.2 and 0.40 ± 0.06 over size ranges corresponding to our reported α(cid:48) 1 slopes, respec- tively. Our best-fit power law slopes calculated for bright MPC L4 Trojans are consistent with these previously- published values at better than the 1σ level. 0 and α(cid:48) We combined the magnitude distributions of faint Sub- aru Trojans and brighter catalogued Trojans to arrive at the overall magnitude distribution of L4 Trojans, shown in Figure 7, where the cumulative absolute mag- nitude distribution of faint Trojans was approximately scaled to match the overall number of MPC Trojans at H ∼ 12. The error bars indicate 95% confidence bounds derived from Poisson errors on the Subaru survey data and reflect the uncertainty associated with scaling up the survey magnitude distribution to approximate the magnitude distribution of the total L4 population. The combined Subaru and MPC data sets cover the entire magnitude range from H = 7.2 to a limiting magni- tude of Hmax = 16.4. The best-fit power law distribu- tion slopes in the region containing the overlap between 1 = 0.43 ± 0.02 and the MPC and Subaru data sets (α(cid:48) α1 = 0.45 ± 0.05) are statistically equivalent, indicating that the magnitude distribution between H ∼ 10 and H ∼ 15 is well-described by a single power law slope. The overall magnitude distribution is characterized by three distinct regions: The brightest L4 Trojans have a power law magnitude distribution with slope α0 = 0.91. At intermediate sizes, the magnitude distribution rolls over to a slope of α1 ∼ 0.44. Finally, the faintest objects detected in our Subaru survey are characterized by an even shallower magnitude distribution slope of α2 = 0.36. These three regions are separated at the break magni- tudes H(cid:48) b = 8.46+0.49−0.54 and Hb = 14.93+0.73−0.88, which corre- spond to Trojans of size 135+38−27 km and 7+3−2 km, respec- tively. In a previous study, Yoshida & Nakamura (2005) de- tected 51 faint L4 Trojans near opposition using the Suprime-Cam instrument and found the magnitude dis- tribution to be well-described by a broken power law with a break at around H ∼ 16 separating a brighter- end slope of 0.48 ± 0.02 from a shallower faint-end slope of 0.26±0.02. The methods used in the Yoshida & Naka- mura (2005) fits differed from ours in several ways: The data was binned prior to fitting, and the two slopes were determined from independent fits of the bright and faint halves of their data. To better compare the distribu- tion of Trojans studied by Yoshida & Nakamura (2005) with the best-fit distribution we derived from our Subaru Fig. 8. -- Cumulative magnitude distribution of the less-red and red L4 Trojan populations, as constructed using the methods of Wong et al. (2014) for objects in the Minor Planet Center catalog through H = 12.3 (cyan triangles and red squares, respectively). These distributions have been scaled to correct for catalog and categorization incompleteness; the error bars denote the 95% confi- dence bounds and are derived from the binomial distribution errors associated with correcting for uncategorized less-red and red Tro- jans, as well as the uncertainties from the catalog incompleteness correction in the range H = 11.3 − 12.3. The best-fit curves de- scribing the distributions are overplotted (solid blue and red lines, respectively), with the power law slopes indicated for their corre- sponding magnitude regions. survey, we reanalyzed their data using the techniques de- scribed in this paper. Fitting all of the Yoshida & Naka- mura (2005) magnitudes through H = 17.9 (90% com- pleteness limit) to a broken power-law, we obtained the two slopes α1,Y&N = 0.44+0.07−0.06 and α2,Y&N = 0.26+0.06−0.04 and a roll-over at Hb,Y&N = 15.11+0.89−1.02. These values are consistent with the corresponding best-fit values for α1, α2, and Hb from the analysis of our Subaru survey data at better than the 1σ level. 3.2. Color distribution Previous spectroscopic and photometric studies of Tro- jans have noted bimodality in the distribution of various properties, including the visible (Szab´o et al. 2007; Roig et al. 2008; Melita et al. 2008) and near-infrared (Emery et al. 2011) spectral slope, as well as the infrared albedo (Grav et al. 2012). Wong et al. (2014) demonstrated that the bimodal trends were indicative of two separate pop- ulations within the Trojans, which are referred to as the less-red and red populations, in accordance with their relative colors. It was further shown that the magnitude distributions of these two color populations are distinct, with notably different power-law slopes in the magnitude range H ∼ 9.5 − 12.3. We repeated the color population analysis presented in Wong et al. (2014), using only L4 Trojans. Objects were categorized as less-red or red primarily based on their visible spectral slopes, which were derived from the g, r, i, and z magnitudes listed in 4th release of the SDSS Moving Object Catalog (SDSS-MOC4). We estimated the incompleteness in color categorization via the frac- tion of objects in each 0.1 mag bin that we were able to categorize as either less-red or red; when fitting the 0.33 +0.04 -0.03 0.84 +0.22 -0.16 0.61 +0.07 -0.06 magnitude distributions of the color populations up to a limiting magnitude of H = 12.3, the categorization in- completeness was factored into the efficiency function η along with the estimated incompleteness of the MPC cat- alog. See Wong et al. (2014) for a complete description of the methodology used. 2 = 0.33+0.04−0.03, H R The magnitude distributions of the color populations were fit using the same techniques that we applied in the total magnitude distribution fits in Section 3.1. The red population magnitude distribution is best-fit by a broken 1 = 0.84+0.22−0.16, αR power law with αR 0 = 7.30+0.30−0.26, and H R b = 8.82+0.35−0.44. The less-red population has a magnitude distribution more consistent with a sin- gle power law: Σ(α1, H0H) = 10α1(H−H0); here we com- 1 = 0.61+0.07−0.06 puted the following best-fit parameters: αLR 0 = 8.18+0.31−0.29. Figure 8 shows the cumulative and H LR magnitude distributions for the less-red and red popu- lations, scaled to correct for incompleteness, along with the best-fit curves. The error bars indicate the 95% con- fidence bounds derived from the incompleteness correc- tion. The best-fit slopes calculated above for the L4 color populations are consistent within the errors to the corre- sponding values in Wong et al. (2014) derived from the color analysis of both L4 and L5 Trojans. To determine whether the previously-studied bimodal- ity in color among the brighter Trojans carries through to the fainter Trojans from our Subaru observations, we constructed a histogram of the g−i color distribution for all Trojans contained in the SDSS-MOC4 catalog and compared it to the histogram of g − i colors for Tro- jans detected in our Subaru survey. As discussed in Sec- tion 2.4, the bulk distribution of colors is not expected to be affected by detection incompleteness due to bad seeing, and we include all Trojans brighter than H = 18. The two histograms are plotted in Figure 9. The bi- modality in the color histogram of brighter SDSS-MOC4 9 objects is evident. We fit a two-peaked Gaussian to the color distribution of brighter SDSS-MOC4 Trojans and found that the less-red and red populations have mean g − i colors of µ1 = 0.73 and µ2 = 0.86, respectively. On the other hand, while there is some asymmetry in the color distribution of faint Trojans detected by our Subaru survey, there is no robust bimodality. This can be mostly attributed to the large contribution of asteroid rotation to the variance in the color measurements, the magnitude of which is comparable to the difference be- tween the mean red and less-red colors (see Section 2.5). As such, it is not possible to categorize individual Tro- jans from our Subaru observations into the less-red and red populations based on their colors and construct mag- nitude distributions of the color populations, as was done in Wong et al. (2014) for the brighter SDSS-MOC4 Tro- jans. Instead, we considered bulk properties of the distribu- tion. In particular, we calculated the mean g − i color as a function of H magnitude for the combined set of SDSS-MOC4 Trojans and faint Trojans from our Sub- aru observations in order to assess whether the result- ing trend is consistent with extrapolation of the best- fit color magnitude distributions obtained previously for the brighter objects (i.e., curves in Figure 8). Here, we assumed that the mean colors of the two color popu- lations are invariant across all magnitudes. The mean g− i color and uncertainty in the mean for the data were computed in 1 mag bins and are plotted in Figure 10 in blue. We see that the mean color is consistent with a monotonically-decreasing trend with increasing magni- tude, or equivalently, decreasing size. To derive the ex- trapolated mean color values from the best-fit color mag- Fig. 9. -- Histogram of the g−i color distribution for Trojans con- tained in the SDSS-MOC4 catalog brighter than H = 12.3 (green) and fainter Trojans detected in our Subaru survey (blue). There is clear bimodality in the distribution of brighter objects in the SDSS- MOC4 catalog, while the distribution of faint Trojan colors does not display a clear bimodality. This is likely due to the large rela- tive uncertainties associated with the measurement of faint Trojan colors due to asteroid rotation. Fig. 10. -- Mean g − i colors and corresponding uncertainties for the combined set of Trojans detected in our Subaru survey and L4 Trojans listed in SDSS-MOC4 (blue squares with error bars). Red dots denote the predicted mean g − i color values computed from extrapolation of the best-fit less-red and red population mag- nitude distributions, assuming mean less-red and red colors of 0.73 and 0.86, respectively. The mean less-red and red g − i colors are indicated by dashed green lines. The monotonic decrease in mean g − i color indicates an increasing fraction of less-red Trojans with decreasing size. The agreement between the extrapolated values and the measured ones suggests that the best-fit color-magnitude distributions derived from bright catalogued Trojans likely extend throughout the magnitude range studied by our Subaru survey. 0.00.20.40.60.81.01.2g−i010203040506070CountSUBARU dataSDSS-MOC4Mean red color Mean less-red color 10 nitude distributions, we calculated the expected mean color at each bin magnitude as a weighted mean c(cid:48)(H) = (µ1×ΣLR(H)+µ2×ΣR(H))/(ΣLR(H)+ΣR(H)), where µ1,2 are the mean g − i colors of the less-red and red populations, respectively, as derived previously from fit- ting the color distribution of SDSS-MOC4 Trojans. The functions ΣLR(H) and ΣR(H) are the best-fit differential magnitude distributions for the less-red and red popula- tions presented earlier. The resulting model mean color values are denoted in Figure 10 by red dots. The extrapolated mean colors show very good agree- ment with the mean colors derived from the data. This suggests that the best-fit magnitude distributions for the less-red and red color populations shown in Figure 8 continue past the limiting magnitude of the MPC data analysis (H = 12.3) and likely extend throughout most of the magnitude range covered by our Subaru observa- tions. We conclude that the fraction of less-red objects in the overall L4 Trojan population increases steadily with increasing magnitude (decreasing size), and that L4 Tro- jans fainter than H ∼ 16, or equivalently, smaller than ∼4 km in diameter, are almost entirely comprised of less- red objects. 4. DISCUSSION The analysis of faint Trojans detected in our Subaru survey offers the most complete picture of the L4 Trojan population to date, refining the known absolute magni- tude distribution over the entire range from H = 7.2 to H = 16.4 and providing the distribution of Trojan col- ors down to kilometer-sized objects. The most notable features in the total magnitude distribution (Figure 7) are the two slope transitions at H ≈ 8.5 and H ≈ 15.0. Such breaks in the power-law shape are common features in the magnitude distributions of small body populations throughout the Solar System and are generally attributed to collisional evolution (see, for example, the review by Durda et al. 1998). Previous studies of the Trojans' col- lisional history sought to explain the observed bright-end break at H ≈ 8.5 using collisional modeling (e.g. Wong et al. 2014; Marzari et al. 1997; de El´ıa & Brunini 2007) and found that, given the very low intrinsic collisional probability of the Trojan clouds, the bright-end slope transition is best reproduced by assuming that a break was present at the time when the Trojans were emplaced in their current location. In other words, the collisional activity among the large Trojans over the past ∼4 Gyr is likely not sufficient to have produced the bright-end break at H ≈ 8.5 starting from a single power-law slope. We propose that it is the faint-end break at H ≈ 15.0 that represents the transition to the part of the Tro- jan population that has reached collisional equilibrium since emplacement; objects brighter than the faint-end break have not reached collisional equilibrium and there- fore largely reflect the primordial magnitude distribution of the Trojans at the time of capture by Jupiter. Turning to the color distributions, we recall that the magnitude distribution fits we obtained for the less-red (LR) and red (R) color populations through H = 12.3 (see Figure 8) show highly distinct slopes for objects fainter than the bright-end break (0.33+0.04−0.03 for the R population, and 0.61+0.07−0.06 for the LR population). This indicates that the fraction of LR objects in the over- all population increases with increasing magnitude. In Section 3.2, we found that this general trend continues throughout the magnitude range covered in our Subaru survey. As was done in Wong et al. (2014) based on the color-magnitude analysis of L4 and L5 Trojans listed in the Minor Planet Center catalog, we posit that the LR vs. R magnitude distribution slope discrepancy can be explained if R objects convert to LR objects upon collision. Such a process would naturally account for the relative flattening the R population's magnitude dis- tribution slope and the simultaneous steepening of LR population's magnitude distribution slope. The R-to-LR conversion model assessed here suggests that the LR and R Trojans have similar interior com- positions, with the difference in color confined to the exposed surface layer. Current models of Solar System formation and evolution indicate that the Trojans may have been sourced from the same body of material as the KBOs, located in the region of the disk beyond the primordial orbit of Neptune (Morbidelli et al. 2005). Re- cent observational studies of KBOs have revealed that the Kuiper belt is comprised of several sub-populations, among which are the so-called "red" and "very red" small KBOs (Fraser et al. 2008; Peixinho et al. 2012). Brown et al. (2011) hypothesized that this color bimodality may be attributable to the wide range of heliocentric distances at which the KBOs formed. In this scenario, all of these objects were accreted from a mix of rock and volatile ices of roughly cometary composition. Immediately following the dissipation of the primordial disk, the surface ices on these bodies began sublimating from solar irradiation, with the retention of a particular volatile ice species on an object's surface being determined primarily by the temperature of the region where the object resided: Ob- jects located at greater heliocentric distances would have retained that ice species on their surfaces, while those that formed at lesser heliocentric distances would have surfaces that were completely depleted in that ice species. Brown et al. (2011) proposed that the continued irradi- ation of volatile ices led to a significant darkening of the surface and the formation of a robust irradiation mantle, which served to protect ices in the interior from sublimat- ing away. The precise effect of irradiation on the surface is likely dependent on the types of volatile ices retained on the surface. Therefore, the presence or absence of one particular volatile ice species may be the key factor in producing the observed color bimodality in the small KBOs. Specifically, objects that retained that volatile ice species on their surfaces formed a "very red" irradia- tion mantle, while those that lost that volatile ice species from their surfaces formed a "red" irradiation mantle. Wong et al. (2014) suggested that the LR and R Jupiter Trojans may have been drawn from the same two sources as the "red" and "very red" KBOs, respec- tively. Upon a catastrophic impact, the irradiation man- tle on a Trojan's surface would disintegrate and any ex- posed volatile ices in the interior would sublimate away within a relatively short timescale, leaving behind col- lisional fragments comprised primarily of water ice and rock. Without the differing collection of volatile ices on the surface to distinguish them, the fragments of LR ob- jects and R objects would be spectroscopically identical to each other. Subsequent irradiation of these pristine fragments would raise the spectral slope slightly, but not 11 velocity of collisional fragments. At smaller sizes (below about 1 km in diameter), the intrinsic material strength of the target becomes the dominant factor; here, smaller objects tend to have fewer cracks and defects than larger objects, and therefore the collisional strength increases with decreasing size. Our model computed the collisional evolution of the two color populations separately. At each time step, the simulation considered the number of collisions between objects of the same color, as well as collisions involving objects of different colors. The conversion of red objects to less-red fragments through shattering was modeled by placing all collisional fragments into less-red bins, regard- less of the color of the target or the impactor. After run- ning simulations for various values of the parameters (α∗ 2, k, c), we found that a large number of test runs yielded final total and color magnitude distributions that were consistent with the best-fit distributions of catalogued Trojans presented in Section 3. To determine which run best reproduced the observed distributions, we compared the simulation results directly with the best-fit distri- bution curves and minimized the chi-squared statistic. Here, χ2 was computed as the sum of χ2 values for the total, LR, and R magnitude distributions. The test run that resulted in the best agreement with the data has an initial total distribution with faint-end slope α∗ 2 = 0.44, a strength scaling parameter c = 1.0, and an initial R- to-LR number ratio k = 5.5. Plots comparing the final simulated distributions from this test run to the best-fit distribution curves are shown in Figure 11. It is impor- tant to note that the simulated total magnitude distri- bution from the collisional model is not sensitive to the R-to-LR conversion, and the ability of our simulations to reproduce the observed total magnitude distribution to the same extent as would result if volatile ices were retained on the surface. All collisional fragments, regard- less of the surface color of their progenitor bodies, would eventually attain the same surface color, which would be relatively less-red when compared to the color of R Tro- jans. As a consequence, collisional evolution of the Tro- jan population since emplacement would have gradually depleted the number of R Trojans while simultaneously enriching the number of LR Trojans. To assess the hypotheses mentioned above, we ran a series of numerical simulations to model the collisional evolution of the L4 Trojan population since emplace- ment, following the methodology used in Wong et al. (2014). Given the low intrinsic collisional probability of Trojans, we defined the initial magnitude distribution as a broken power law of the form described in Eq. (2) with a bright-end distribution identical to that of the currently-observed L4 population (α1 = 0.91, H0 = 7.22, and Hb = 8.46). We varied the initial faint-end slope α∗ 2 across different trials. The initial population for each trial consisted of objects with absolute magnitudes in the range H = 7 → 30, divided into 75 logarithmic diameter bins. We constructed the initial color populations by taking constant fractions of the total initial population across all bins. The initial R-to-LR number ratio, k, ranged from 4 to 6, in increments of 0.5. The collisional evolution was carried out over 4 Gyr in 100000 time steps of length ∆t = 40000. At each time step, the expected number of collisions Ncoll between bodies belonging to any pair of bins is given by Ncoll = 1 4 (cid:104)P(cid:105)NtarNimp∆t(Dtar + Dimp)2, (5) where Ntar and Nimp are the number of objects in a tar- get bin with diameter Dtar and an impactor bin with di- ameter Dimp, respectively; (cid:104)P(cid:105) = 7.79×10−18 yr−1 km −2 is the intrinsic collision probability for Trojan-Trojan col- lisions calculated by Dell'Oro et al. (1998) for L4 Trojans. For objects with diameter Dtar, there exists a minimum impactor diameter Dmin necessary for a shattering colli- sion. Dmin is defined as (Bottke et al. 2005) Dmin = Dtar, (6) (cid:33)1/3 (cid:32) 2Q∗ V 2 imp D where Vimp = 4.66 km s−1 is the L4 impact velocity cal- culated by Dell'Oro et al. (1998), and Q∗ D is the colli- sional strength of target. In our algorithm, we utilized a size-dependent strength scaling law based off one used by Durda et al. (1998) in their treatment of collisions among small main-belt asteroids: D = c· 10· (155.9D−0.24 + 150.0D0.5 + 0.5D2.0) J kg −1. Q∗ (7) Here, we included a normalization parameter c to ad- just the overall scaling of the strength; c varied in in- crements of 0.5 from 1 to 10 in our test trials. The strength scaling model used here has a transition from a gravity-dominated regime for large objects to a strength- dominated regime for smaller objects. For large Trojans, the collisional strength increases rapidly with increasing size, since the impact energy required to completely shat- ter a large object is primarily determined by the escape Fig. 11. -- Comparison between the results from the best test run of our collisional simulation (dotted lines) and the observed L4 Trojan magnitude distributions (solid lines). The initial total magnitude distribution for the best test run is a broken power law with α1 = 0.91, α∗ 2 = 0.44, H0 = 7.22, Hb = 8.46, an initial R- to-LR number ratio k = 5.5 and a strength normalization factor c = 1.0. Black, red, and blue colors indicate the magnitude dis- tribution of the total, red, and less-red populations, respectively. The consistency of the model results and the best-fit distributions demonstrates that the proposed conversion of R objects to LR frag- ments upon collision is capable of explaining the different shapes of the LR and R magnitude distributions. 789101112131415Absolute magnitude, H100101102103Cumulative number 12 Fig. 12. -- Comparison of the best-fit cumulative magnitude dis- tribution curve from the survey data (solid line) with the final pre- dicted distribution from two collisional simulation runs: [1] test run assuming bodies with collisional strength given by Eq. (7) (dashed line) and [2] test run assuming strengthless bodies with collisional strength scaling given by Eq. (8) (dot-dash line). The better agree- ment of the latter suggests that Trojans have very low material strength, similar to comets. holds regardless of any assumptions made about the na- ture of the color populations. The similarity between the best initial test distribution and the current total magnitude distribution reaffirms the conclusion of previous studies that collisions have not played a major role in shaping the magnitude distribu- tion of large Trojans since emplacement. Meanwhile, the R-to-LR collisional conversion model yields simulated fi- nal color magnitude distributions that match the best-fit color magnitude distributions of catalogued Trojans well. Furthermore, we computed the expected trend in mean g − i color from the simulated color magnitude distribu- tions through the magnitude region spanning the Subaru Trojan data and found that the trend is consistent with the measured mean g − i colors from the data (as shown in Figure 10). To determine whether the collisional model can repro- duce the faint-end break at H ≈ 15.0, we compared the simulated final magnitude distribution from the best test run with the observed magnitude distribution of faint Trojans from our Subaru data (Figure 12). All of the sim- ulation test runs predict a break at around H = 14− 15; however, the slope that the simulated distributions roll over to is almost identical to the slope ahead of the break. In the case of the best test run, the faint-end rollover in the simulated final magnitude distribution is barely discernible. In terms of collisional equilibrium, this means that the predicted equilibrium slope is around αeq ∼ 0.43. Meanwhile, the actual faint-end slope de- rived from fitting the Subaru data is somewhat shallower (α2 = 0.36+0.05−0.09). One possible explanation for this discrepancy is that the collisional strength of Trojans may not be well- described by the strength model defined in Eq. (7). To explore this possibility, we considered the case where Tro- jans have negligible material strength and are loosely- held conglomerates of rock and ice similar to comets. We modified the collisional strength scaling relation to incorporate only the effect of self-gravity by removing the transition to a strength-dominated regime that we included in the previous model. The new strength for- mula is given by: D = c · 5D2.0 J kg Q∗ −1. (8) Rerunning the collisional simulations with this new col- lisional strength scaling, we established a new best test run for strengthless bodies with α∗ 2 = 0.45, k = 5.5, and c = 0.5. The simulated color magnitude distribu- tions in the strengthless case were found to be gener- ally consistent with the ones produced in the original non-strengthless model and likewise predict the observed trend in mean g − i color through H ∼ 18. The simu- lated final magnitude distribution from this model in the vicinity of the faint-end break is shown in Figure 12. We can see that the predicted distribution for strengthless Trojans provides a better match to the data than the previous case of non-strengthless bodies. This indicates that Trojans may have very low material strength, which would make them more comparable to comets than to main belt asteroids and lend support to the hypothesis presented earlier that the Jupiter Trojans formed in the primordial trans-Neptunian region and were later scat- tered inward during a period of dynamical instability. 5. CONCLUSION We detected 557 Trojans in a wide-field survey of the leading L4 cloud using the Suprime-Cam instrument on the Subaru Telescope. All objects were imaged in two filters, and the g − i color was computed for each ob- ject. After removing objects imaged during bad seeing and establishing a limiting magnitude of H = 16.4, we computed the best-fit curves describing the overall mag- nitude distribution. In addition, we examined the distri- bution of g−i colors for the faint objects detected by our survey and compared it to an extrapolated model based on the magnitude distributions of bright, catalogued ob- jects in the less-red and red Trojan populations. The color-magnitude distribution analysis was supplemented by collisional simulations, from which we made predic- tions about the formation and evolution of the Trojan population. The main results are summarized below: • The overall magnitude distribution of L4 Trojans is described by three power-law slopes: The dis- tribution of the brightest objects follows a power law slope of α0 = 0.91+0.19−0.16. At intermediate sizes, the magnitude distribution rolls over to a slope of α1 ∼ 0.44. Finally, the faintest objects are characterized by an even shallower magnitude dis- tribution slope of α2 = 0.36+0.05−0.09. These three regions are separated by rollovers in the magni- tude distribution located at H(cid:48) b = 8.46+0.49−0.54 and Hb = 14.93+0.73−0.88, which correspond to objects with diameters of 135+38−27 km and 7+3−2 km, respectively. • The faint-end break in the overall Trojan magni- tude distribution at H ∼ 15 is reproduced by our collisional simulations and indicates the transition between objects that have not experienced signif- icant collisional evolution and objects that have achieved collisional equilibrium. 12131415161718Absolute magnitude, H102103104105Cumulative numberBest-fit distributionCollisional model, with strengthCollisional model, strengthless • The shallow faint-end slope (α2 = 0.36+0.05−0.09) is consistent with Trojans having very low material strength, similar to comets. • The mean g−i color of Trojans follows a general de- creasing trend with increasing magnitude, or equiv- alently, decreasing size. At faint magnitudes, this trend is consistent with the extrapolation of magni- tude distribution fits computed for bright objects in the less-red and red populations. Less-red ob- jects dominate among objects smaller than ∼5 km in diameter. • The discrepant best-fit slopes of the color- 13 magnitude distributions for objects smaller than ∼50 km and the monotonically-decreasing trend in mean g − i color with decreasing size are consis- tent with the conversion of red objects to less-red fragments upon collision. ACKNOWLEDGMENTS This research was supported by Grant NNX09AB49G from the NASA Planetary Astronomy Program and by the Keck Institute for Space Studies. The authors also thank an anonymous reviewer for constructive comments that helped to improve the manuscript. REFERENCES Bernstein, G., & Khushalani, B. 2000, AJ, 120, 3323 Bertin, E. 2006, in Astronomical Data Analysis Software and Systems XV (ASP Conf. Ser. 351), ed. C. Gabriel, C. Arviset, D. Ponz, & S. Enrique (San Francisco, CA: ASP), 112 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Bottke, W. F., Durda, D. D., Nesvorn´y, et al. 2005, Icar, 175, 111 Brown, M. E., Schaller, E. L., & Fraser, W. C. 2011, ApJ, 739, L60 de El´ıa, G. C., & Brunini, A. 2007, A&A, 475, 375 Dell'Oro, A., Marzari, F., Paolicchi, P., Dotto, E., & Vanzani, V. 1998, A&A, 339, 272 Dotto, E., Fornasier, S., Barucci, M. A., et al. 2006, Icar, 183, 420 Durda, D. D., Greenberg, R., & Jedicke, R. 1998, Icar, 135, 431 Emery, J. P., Burr, D. M., & Cruikshank, D. P. 2011, AJ, 141, 25 Fern´andez, Y. R., Jewitt, D., & Ziffer, J. E. 2009, AJ, 138, 240 Foreman-Mackey, D., Hogg, D. W., Lang D., & Goodman, J. 2013, arXiv:1202.3665 Fornasier, S., Dotto, E., Hainaut, O., et al. 2007, Icar, 190, 622 Fraser, W. C., Kavelaars, J. J., Holman, M. J., et al. 2008, Icar, 195, 827 Marzari, F., Farinella, P., Davis, D. R., Scholl, H., & Campo-Bagatin, A. 1997, Icar, 125, 39 Marzari, F., & Scholl, H. 1998, Icar, 131, 41 Melita, M. D., Licandro, J., Jones, D. C., & Williams, I. P. 2008, Icar, 195, 686 Miyazaki, S., Komiyama, Y., Sekiguchi, M., et al. 2002, PASJ, 54, 833 Morbidelli, A., Levison, H. F., Tsiganis, K., & Gomes, R. 2005, Natur, 435, 462 Nakamura, T., & Yoshida, F. 2008, PASJ, 60, 293 Nesvorn´y, D., & Morbidelli, A. 2012, AJ, 144, 117 Peixinho, N., Delsanti, A., Guilbert-Lepoutre, A., Gafeira, R., & Lacerda, P. 2012, A&A, 546, 86 Pravec, P. Harris, A. W., & Micha(cid:32)lowski, T. 2002, in Asteroids III, ed. W. F. Bottke, A. Cellino, P. Paolicchi, & R. P. Binzel (Tucson, AZ: University of Arizona Press), 113 Roig, F., Ribeiro, A. O., & Gil-Hutton, R. 2008, A&A, 483, 911 Szab´o, Gy. M., Ivezi´c, Z, Juri´c, M., & Lupton, R. 2007, MNRAS, 377, 1393 Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Grav, T., Mainzer, A. K., Bauer, J. M., Masiero, J. R., & Nugent, Natur, 435, 459 C. R. 2012, ApJ, 759, 49 Jewitt, D. C., Trujillo, C. A., & Luu, J. X. 2000, AJ, 120, 1140 Wong, I., Brown, M. E., & Emery, J. P. 2014, AJ, 148, 112 Yang, B., & Jewitt, D. 2007, AJ, 134, 223 Yoshida, F., & Nakamura, T. 2005, AJ, 130, 2900 Yoshida, F., & Nakamura, T. 2008, PASJ, 60, 297
1803.01971
2
1803
2018-06-29T21:08:33
Exoplanets Torqued by the Combined Tides of a Moon and Parent Star
[ "astro-ph.EP" ]
In recent years, there has been interest in Earth-like exoplanets in the habitable zones of low mass stars ($\sim0.1-0.6\,M_\odot$). Furthermore, it has been argued that a large moon may be important for stabilizing conditions on a planet for life. If these two features are combined, then an exoplanet can feel a similar tidal influence from both its moon and parent star, leading to potentially interesting dynamics. The moon's orbital evolution depends on the exoplanet's initial spin period $P_0$. When $P_0$ is small, transfer of the exoplanet's angular momentum to the moon's orbit can cause the moon to migrate outward sufficiently to be stripped by the star. When $P_0$ is large, the moon migrates less and the star's tidal torques spin down the exoplanet. Tidal interactions then cause the moon to migrate inward until it is likely tidally disrupted by the exoplanet and potentially produces rings. While one may think that these findings preclude the presence of moons for the exoplanets of low mass stars, in fact a wide range of timescales are found for the loss or destruction of the moon; it can take $\sim10^6-10^{10}\,{\rm yrs}$ depending on the system parameters. When the moon is still present, the combined tidal torques force the exoplanet to spin asynchronously with respect to both its moon and parent star, which tidally heats the exoplanet. This can produce heat fluxes comparable to those currently coming through the Earth, arguing that combined tides may be a method for driving tectonic activity in exoplanets.
astro-ph.EP
astro-ph
Accepted for publication in The Astronomical Journal Preprint typeset using LATEX style emulateapj v. 12/16/11 8 1 0 2 n u J 9 2 . ] P E h p - o r t s a [ 2 v 1 7 9 1 0 . 3 0 8 1 : v i X r a EXOPLANETS TORQUED BY THE COMBINED TIDES OF A MOON AND PARENT STAR The Observatories of the Carnegie Institution for Science, 813 Santa Barbara St., Pasadena, CA 91101, USA; [email protected] Accepted for publication in The Astronomical Journal Anthony L. Piro ABSTRACT In recent years, there has been interest in Earth-like exoplanets in the habitable zones of low mass stars (∼ 0.1 − 0.6 M(cid:12)). Furthermore, it has been argued that a large moon may be important for stabilizing conditions on a planet for life. If these two features are combined, then an exoplanet can feel a similar tidal influence from both its moon and parent star, leading to potentially interesting dynamics. The moon's orbital evolution depends on the exoplanet's initial spin period P0. When P0 is small, transfer of the exoplanet's angular momentum to the moon's orbit can cause the moon to migrate outward sufficiently to be stripped by the star. When P0 is large, the moon migrates less and the star's tidal torques spin down the exoplanet. Tidal interactions then cause the moon to migrate inward until it is likely tidally disrupted by the exoplanet and potentially produces rings. While one may think that these findings preclude the presence of moons for the exoplanets of low mass stars, in fact a wide range of timescales are found for the loss or destruction of the moon; it can take ∼ 106 − 1010 yrs depending on the system parameters. When the moon is still present, the combined tidal torques force the exoplanet to spin asynchronously with respect to both its moon and parent star, which tidally heats the exoplanet. This can produce heat fluxes comparable to those currently coming through the Earth, arguing that combined tides may be a method for driving tectonic activity in exoplanets. Subject headings: celestial mechanics - planet–star interactions - planets and satellites: dynamical evolution and stability - planets and satellites: general 1. INTRODUCTION It has been revealed in recent years that terrestrial ex- trasolar planets are basically ubiquitous in our Galaxy (Burke et al. 2015; Mulders et al. 2015), and, although extrasolar moons have yet to be discovered, certainly many of these exoplanets have moons just like the plan- ets in our own solar system. Whether or not a planet has a moon is not just a minor curiosity, but potentially fundamental to the question of whether these planets host life. In the case of the Earth, the Moon's relatively large size allows it to stabilize the Earth's obliquity. This may be crucial for stabilizing conditions on Earth for a sufficiently long time to allow life to develop (Ward & Brownlee 2000, although also see Lissauer et al. 2012; Li & Batygin 2014). Another factor that may affect a planet's ability to host life is the presence of internal heating and volcan- ism. The associated tectonic activity due to the move- ment of plates sitting atop a fluid mantle can trap atmo- spheric gases such as carbon dioxide into rocks and help stabilize the climate (Walker et al. 1981; Sleep & Zahnle 2001; Foley & Driscoll 2016). Tectonic activity can also cycle fresh rock and minerals out of deeper regions of the planet, providing the building blocks and nutrients for life. Heating can provide the energy needed to drive biochemical reactions as is seen from hydrothermal vents on Earth. In the case of the Earth, this heating is driven by a combination of radioactivity, latent heat, and heat from formation, but in principle a moon could also pro- vide a source of heating through tidal interactions, as is seen for Io. The easiest way to drive such tidal heating would be for a moon via an eccentric orbit. In the ab- sence of a large eccentricity though there could also be cases where a planet is being tidally torqued by both its moon and parent star. The competing torques ensure that the planet is never perfectly synchronized to either the moon or star, so that tidal heating can continue to persist. Motivated by these possibilities, here I consider the evolution of a planet tidally torqued to a similar level by both its moon and parent star. Particular focus is on stars in the mass range of M∗ ≈ 0.1− 0.6 M(cid:12), since they have two attractive properties: (1) there has been strong interest in studying exoplanets in the habitable zones of these stars, and (2) their habitable zones are closer in where tidal interactions with the star are more im- portant. Of course various aspects of star-planet-moon tidal interactions have been investigated previously. As early as in Counselman (1973), it was noted such systems would evolve toward one of three states, (1) the moon mi- grates inward until it reaches the planet, (2) the moon migrates outward until it escapes from the planet, and (3) the moon finds a stable state where its orbital fre- quency and the planetary spin frequency are at a mutual resonance. Since this work, the problem has continued to be studied in various ways over a wide variety of possi- ble systems (e.g., Ward & Reid 1973; Touma & Wisdom 1994; Neron de Surgy & Laskar 1997; Barnes & O'Brien 2002; Sasaki et al. 2012; Sasaki & Barnes 2014; Adams & Bloch 2016). Here I focus on particular aspects of this previous work by highlighting the possibility of tidal heating and the importance of the initial spin period of the planet in determining the resulting dynamics. In Section 2, I motivate and summarize the parameter range for the star-planet-moon systems I will be consid- ering. In Section 3, I present the set of equations I use to solve for the orbits and tides. In Section 4, I summa- 2 rize the characteristic timescale of the problem and solve for the evolution of the exoplanet's spin in the limit of no orbital evolution. This provides some intuition that is useful for solving the more detailed evolution of the system. In Section 5, I explore the full evolution of the star-planet-moon system over a range of parameters, and I conclude in Section 6 with a summary. 2. SETTING THE STAGE Before diving into the details of tidal interactions, it is helpful to motivate the range of parameters that will be considered. In Figure 1, I plot the habitable zone for stars in the mass range of M∗ = 0.1 − 0.6 M(cid:12) (green shaded region). This is estimated as the range of dis- tances at which an exoplanet would receive the same flux as at distances of 0.8 − 1.7 AU from the Sun (this is the more optimistic range from the work of Kasting et al. 1993–a less optimistic range would be 0.95− 1.4 AU). In comparison to the Sun, the habitable zone must be fairly close to the parent star because the luminosity varies strongly with mass. If an exoplanet with an exomoon is within this range of distances from the star, it is possible that the star's gravity is sufficiently strong to unbind that moon. The critical distance within which a moon can remain is pro- portional to the so-called Hill sphere, RH, so that acrit,m = f RH = f a∗ , (1) for which I use a constant factor f = 0.49 (Domingos et al. 2006, as appropriate for prograde orbits) for this work (although note that a value of f = 0.36 has also been argued for in the past by Holman & Wiegert 1999). This relation can be inverted to find a critical distance of the planet from the star, within which the moon would be stripped away. This is given by (cid:19)1/3 (cid:18) Mp 3M∗ acrit,∗ = am f (cid:18) 3M∗ (cid:18) am (cid:19)1/3 (cid:19)(cid:18) Mp Mp (cid:19)−1/3(cid:18) M∗ (cid:19)1/3 aM M(cid:12) M⊕ = 0.52 AU, (2) where aM = 3.84 × 1010 cm is the distance between the In Figure 1, I plot acrit,∗ for differ- Earth and Moon. ent values of am/aM (blue dashed lines). In all cases, I assume the mass of the exoplanet is Mp = M⊕, where M⊕ = 5.97×1027 g is the mass of the Earth, and the mass of the exomoon is Mm = MM, where MM = 7.35× 1025 g is the mass of the Moon. From comparing these crit- ical distances to the habitable zone, one can see that the moon must be sufficiently close to the exoplanet to prevent being stripped. For example, for M∗ = 0.5 M(cid:12) and a∗ = 0.3 AU, then the moon must be at a distance am (cid:46) 0.7 aM from the exoplanet to stay bound. These considerations roughly set the parameters I will be us- ing for the initial conditions of the dynamical evolution calculations. 3. BASIC EQUATIONS I next present the set of equations I will be using to solve for the dynamics of the planet's spin and the or- bital separations. The basic strategy is to focus on the Fig. 1.- The habitable zone for stars in the mass range of M∗ = 0.1 − 0.6 M(cid:12) (green shaded region). This is defined as the range of distances which match the same flux received at distances of 0.8 − 1.7 AU from our Sun. Blue dashed lines indicate acrit,∗, the critical distance for the planet from the star, within which the moon would be stripped away. These are shown for different values of am/aM as labeled. In all cases I use Mp = M⊕ and Mm = MM. secular evolution of the star-planet-moon system rather than follow each orbit individually. In this sense, the con- servations equations used represent averages over many orbits. This allows the evolution of these systems to be considered over much longer timescales of ∼ 105−1010 yrs rather than being focused on the much shorter timescale of the orbits themselves. Further simplifications include assuming that the spin angular momentum of the planet is parallel to the orbital angular momentum of both the moon and planet. Consider the configuration of a star, planet, and moon, with masses M∗, Mp, and Mm, respectively, as shown schematically in Figure 2. The semi-major axis of the orbit of the star and planet is a∗ and the semi-major axis of the planet and moon is am. The orbits are all assumed to be circular. Tidal forces from the star and moon each generate two bulges on the planet on opposite sides. The frequency of these bulges in the frame of the planet are ω∗ = 2(n∗ − Ω) and ωm = 2(nm − Ω) due to the star and moon, respectively, where the factor of 2 represents the two bulges, Ω is the spin of the planet, and the orbital frequencies are n2∗ = GM∗/a3∗ and n2 m = GMp/a3 m. These bulges are not perfectly aligned with the position of the mass generating the bulge because the planet is not a perfect fluid. Instead, the bulge can proceed or lag behind. This is emphasized in Figure 2, where the bulge caused by the star proceeds the star by an angle α and the bulge caused by the moon lags behind by an angle β (note that these angle are exaggerated here and are 3 Fig. 2.- Schematic showing the combined tidal effects on an planet from both a star and moon (definitely not drawn to scale). Two tidal bulges are raised by each the star and moon, which proceed or lag behind the actual position of the forcing body depending on the time for the planet to react to the tides (parameterized by the time lag factor τ ) and the relative frequencies of the planet's spin and the orbits. In this specific example nm > Ω > n∗, so that the bulge from the star proceeds ahead of the star's position by an angle α, while the bulge from the moon lags behind the moon's position by an angle β. much smaller in real systems). These angles depend on the "time lag" of the tidal forcing of the planet τ and tidal forcing frequencies such that α ∝ τ ω∗ and β ∝ τ ωm. In Figure 2, it is assumed that the planet is spinning faster than its orbit around the star, thus ω∗ < 0 and the bulge from the star proceeds its position. Due to the misalignment between the bulge and the object causing the tide, the planet's spin can be torqued up or down. These torques over secular timescales can be approximated as (Ogilvie 2014) and N∗ = Nm = 3 2 3 2 k2τ ω∗ GM 2∗ R5 p a6∗ , k2τ ωm mR5 GM 2 p a6 m , (3) (4) from the star and moon, respectively, where Rp is the radius of the planet and k2 is the Love number which represents the planet's rigidity. In the specific case of Figure 2, N∗ < 0 and Nm > 0 because of whether the bulge proceeds of lags. These torques can be rewritten in terms of the orbital frequencies as (cid:18) Rp (cid:19)3 (cid:19)(cid:18) Rp (cid:18) Mm a∗ n2∗, (cid:19)3 N∗ = 3 2 and k2τ ω∗M∗R2 p (5) 3 2 Nm = k2τ ωmMmR2 p (6) Note that because I am assuming that Mp (cid:29) Mm, there is an additional factor of Mm/Mp for Nm. Mp n2 m. am Note that for this work I express the tidal torque using a constant time lag model (Mignard 1979, 1980, 1981; Hut 1981; Heller et al. 2011) rather than a constant geometric lag model (MacDonald 1964; Goldreich 1966; Murray & Dermott 1999), which would be parameterized with a quality factor Q. The relation between the two is k2τ ω = σk2/Q with σ = sgn ω (see Efroimsky & Makarov 2013 for a discussion of the limitations and implications of different tidal implementations). I choose this formalism simply because it provides a more smooth transition from a positive torque (when the planet is spinning slowly) to zero torque (when the planet is tidal locked) to a negative torque (when the planet is spinning quickly). In contrast, if I leave Q fixed, the torque changes more abruptly as the planet's spin evolves, which causes numerical issues when evolving the orbits and spins. Another issue is that I consider the torque by the star from the star's bulge and the torque by the moon from the moon's bulge but not vice versa (for example, the torque by the star from the moon's bulge). This is a reasonable approximation for the secular limit, because over long timescales the moon's bulge, for example, will appear at all different positions with respect to the star and therefore will average to zero torque. If the orbits were followed individually, there would be more compli- cated behavior on shorter timescales, but this is beyond the secular focus of the present study. The torques change the spin angular momentum of the planet as d dt (IpΩ) = N∗ + Nm, (7) where Ip and Ω are the moment of inertia and spin fre- quency of the planet, respectively. To conserve angu- lar momentum, the orbital separations must also change. The differential equations that describe these are (cid:104) Mp(GM∗a∗)1/2(cid:105) d dt = −N∗, (cid:105) (8) (9) and (cid:104) d dt Mm(GMpam)1/2 + Imnm = −Nm, where Im is the moment of inertia for the moon. For simplicity, I assume that the moon is tidally locked to the planet, which gives rise to the term Imnm in StarPlanetMoonTidal bulgefrom moonTidal bulgefrom starαβΩa*am Taking the derivative of this expression results in Em = IpΩ dΩ dt + GMpMm 2a2 m dam dt . (19) Substituting the part of the torque due to the moon's tide on dΩ/dt and doing some algebra one finds Em =− ωm 2 Nm = = 3 4 k2τ ω2 mMmR2 p Ip ω2 m 4τsyn,m (cid:18) Mm (cid:19)(cid:18) Rp (cid:19)3 Mp am n2 m (20) This change of energy represents the maximum amount of heating possible on the planet due to asynchronous rotation with respect to the moon. Performing a similar set of arguments on the tidal forcing from the star results in (cid:19)2(cid:35)−1 (13) ,(14) E∗ = − ω∗ 2 N∗ = Ip ω2∗ 4τsyn,∗ = 3 4 k2τ ω2∗M∗R2 p (cid:18) Rp (cid:19)3 a∗ n2∗. (21) for the heating of the planet due to the star. I take the total heating rate on the exoplanet to be the sum of the two tidal heating contributions, E = E∗ + Em. (22) In detail, the tidal heating can occur at different loca- tions depending on how the tides are damped, but this requires a more sophisticated treatment of the tides that is outside the scope of this work. 4 Equation (9). Since relatively little angular momentum is in the moon's spin, whether or not this tidal locking occurs does not impact my main conclusions. For the star, I assume that its spin is not changing appreciably from the small tidal torque of the planet and that it can be ignored. From the above differential equations a number of key timescales can be identified. Rewriting Equation (7), one finds (cid:18) dΩ dt = n∗ τsyn,∗ 1 − Ω n∗ where the associated synchronization timescales are (cid:19) + (cid:19) (cid:18) Mp (cid:18) Mp M∗ nm 1 − Ω nm τsyn,m (cid:18) (cid:19)3 (cid:19)(cid:18) a∗ n−1∗ , (cid:19)3 (cid:19)2(cid:18) am Rp n−1 m , , (10) (11) (12) and τsyn,m ≡ τsyn,∗ ≡ λ 3k2τ n∗ 3k2τ nm Mm Rp and λ = Ip/MpR2 p is the radius of gyration for the planet. The orbital separations also change as described by the differential equations da∗ dt = − a∗ τmig,∗ 1 − Ω n∗ and dam dt = − am τmig,m 1 − Ω nm 1 − 6 5 (cid:18) (cid:19)(cid:34) , (cid:19) (cid:18) Rm am where the last term is due to synchronous spin of the moon and I assume Im = (2/5)MmR2 m. For the change in the orbital orbital separations, λ (cid:18) τmig,∗ ≡ (6k2τ n∗)−1 and τmig,m ≡ (6k2τ nm)−1 (cid:18) Mp (cid:18) Mp M∗ (cid:19)5 (cid:19)(cid:18) a∗ n−1∗ , (cid:19)5 (cid:19)(cid:18) am Rp Mm Rp n−1 m . (15) (16) for the migration timescales. Besides the planet's spin and orbital separations, tides will also change the eccentricity of the orbits, which is governed by the equation (Ogilvie 2014) (cid:18) Rp (cid:19)5 am 1 e de dt = − 3 2 k2τ (18nm − 11Ω) Mm Mp nm. (17) 4. CHARACTERISTIC TIMESCALES AND ANALYTIC EVOLUTION SOLUTIONS This happens on a roughly similar timescale to τmig,m. Nevertheless, I assume circular orbits in this work be- cause one of my main goals is to highlight the fact that tidal heating is possible even with no eccentricity. Tidal heating can be present for circular orbits because when both the moon and star exert their tides on the planet, the planet is always forced to spin asynchronously with respect to each of them. To estimate the strength of this heating, first consider the total energy of the com- bined spinning planet plus gravitational interaction with a moon is Em = 1 2 IpΩ2 − GMpMm 2am . (18) The detailed evolution of the star-planet-moon system can be complicated because of the many parameters that can be varied. It is therefore helpful to consider some of the key timescales that will govern the evolution as well as solving a simpler set of evolution equations to provide some intuition on how the evolution will proceed. 4.1. Orbital and Tidal Timescales First we consider the four key timescales for the syn- chronization and migration of the orbits that were iden- tified in Section 3. Substituting physical values, these are τsyn,∗ = 1.8 × 1010 R⊕ τsyn,m = 3.7 × 109 (cid:19)−2 M(cid:12) (cid:19)(cid:18) M∗ (cid:19)−2 (cid:19)(cid:18) Mm (cid:19)−2 (cid:19)(cid:18) M∗ (cid:19)−1 M(cid:12) MM M⊕ yr, M⊕ 580 s 580 s 1 AU (cid:18) k2τ /λ (cid:19)−3(cid:16) a∗ (cid:18) k2τ /λ (cid:19)−3(cid:18) am (cid:18) k2τ (cid:19)−5(cid:16) a∗ (cid:18) k2τ (cid:19)−5(cid:18) am (cid:19)−1(cid:18) Mp (cid:17)6 (cid:19)−1(cid:18) Mp (cid:19)6 (cid:19)−1(cid:18) Mp (cid:17)8 (cid:19)−1(cid:18) Mm (cid:19)8 1 AU 191 s 191 s M⊕ MM aM yr, yr, R⊕ (cid:18) Rp (cid:18) Rp (cid:18) Rp (cid:18) Rp R⊕ × × × × (23) (24) τmig,∗ = 2.5 × 1017 τmig,m = 2.5 × 1011 5 when the orbital separations do not evolve. In other words, I make the approximation that τmig,∗, τmig,m (cid:29) τsyn,∗, τsyn,m. It is found for this case that a particularly simple analytic solution is possible for the planet's spin evolution. Assuming that am and a∗ are constant, Equation (10) becomes a differential equation simply in Ω(t). Integrat- ing this then results in Ω(t) = Ωeq + (Ω0 − Ωeq) exp (−t/τsyn) , (28) where Ω0 ≡ Ω(t = 0) is the initial spin frequency, Ωeq is the equilibrium spin frequency as t → ∞, and the total synchronization time of the planet is defined to be τsyn ≡ τsyn,mτsyn,∗ τsyn,m + τsyn,∗ . (29) (25) The easiest way to estimate Ωeq is to just set dΩ/dt = 0 with Equation (10), and then solve for Ω, resulting in. yr, aM R⊕ (26) where R⊕ = 6.37 × 108 cm is the radius of the Earth, and I have used λ = 0.33 and k2 = 0.3, which are mo- tivated by empirical measurements and fits to the shear modulus and stiffness of the present day Earth (Williams 1994; Henning et al. 2009; Ray & Egbert 2012; Heller & Barnes 2013; Driscoll & Barnes 2015). The time lag is set to τ = 638 s (Lambeck 1977; Neron de Surgy & Laskar 1997), which for the current values of the Earth– Moon system gives a migration rate of the moon of dam/dt = 3.8 cm yr−1 and spindown rate of the planet of dP/dt = −2.2 ms century−1, where P = 2π/Ω is the planet's spin period. As a check on the general frame- work used here, these rates are both in agreement with the currently measured values for the Earth and Moon. At in Equations (23)–(26), the ordering of the timescales is τsyn,m (cid:46) τsyn,∗ (cid:46) τmig,m (cid:28) τmig,∗. Thus for the Sun-Earth-Moon system, the synchronization of the Earth with the Moon's orbit is occurring the fastest of any of these processes. As emphasized in the discussion in Section 1, where things potentially get interesting is when τsyn,∗ ≈ τsyn,m, so that both the tides acting from the star and moon are impacting the planet on similar timescales. Setting τsyn,∗ ≈ τsyn,m using Equations (23) and (24), one can estimate that the typical separation between the planet and moon where this occurs is specific values least used the for (cid:17)(cid:18) Mm (cid:19)1/3(cid:18) M∗ (cid:19)−1/3 (cid:16) a∗ am ≈ 0.5 0.3 AU MM 0.5 M(cid:12) aM. (27) Comparing this estimate for am with Figure 1, one can see that τsyn,∗ ≈ τsyn,m will naturally occur for low mass stars over a wide range of the parameter space where moons can remain bound to planets. 4.2. Analytic Solutions without Orbital Migration To get more intuition for what will happen to the sys- tem when τsyn,∗ ≈ τsyn,m, I consider the simplified case Ωeq ≡ τsyn,mn∗ + τsyn,∗nm τsyn,m + τsyn,∗ . (30) Equation (28) shows that the planet's spin just exponen- tially decays on a timescale τsyn to the final equilibrium spin of Ωeq. In general, this solution is not exactly correct because am can also potentially evolve on similar timescales as shown by Equation (26). This in turn changes τsyn,∗ and τsyn,m, so that they are not constant as assumed to derive Equation (28). An example where this fails is for the Sun-Earth-Moon system, where the migration of the Moon cannot be ignored. Nevertheless, the full evolutions of the star-planet- moon system will demonstrate that the concept of Ωeq plays an important role. In particular, note that from Equation (30) it is apparent that Ωeq cannot exactly equal either n∗ or nm. This means that there will be tidal forcing on the planet by both the star and the moon, and tidal heating will always play some role in a star-planet- moon system. 5. FULL TIME EVOLVED SOLUTIONS Now that the main background has been covered, I consider the full evolution of the star-planet-moon sys- tem. This is solved by numerically integrating forward in time Equations (10), (13), and (14), which are three coupled differential equations in the dependent variables Ω, a∗, and am, respectively. 5.1. Example Evolution It is helpful to first just focus on one fiducial example that exemplifies the main features on the solutions, which is presented in Figure 3. For this specific case, I use M∗ = 0.5 M(cid:12), Mp = M⊕, and Mm = MM (in all further examples, Mp and Mm use these values for simplicity). The initial orbital separations are a∗ = 0.3 AU and am = 0.5 aM. Finally, the initial spin of the planet must be chosen, in this case I use P0 = 2π/Ω0 = 7 hrs (note that the Earth is generally thought to have had an initial spin period of roughly 6 hrs). From this example, a number of important features are seen that will inform our more detailed parameter survey below. First, in the upper panel of Figure 3, I summarize the key periods of the system. These are the actual spin 6 Fig. 4.- Although the planet can spin down until Ω ≈ Ωeq, this is not a stable equilibrium. Because nm > Ωeq > n∗, this means that (1) the moon torques up the planet, which then (2) moves the moon closer to conserve angular momentum, so that (3) at the new spin equilibrium again nm > Ωeq because Ωeq > n∗. Thus this loops continue and the moon moves in toward the planet until it is disrupted. not stable. As shown by Equation (30), if spin equilib- rium is reached then the planet is not tidally locked with either the moon or the star, and thus the tidal forces still persist. Given the relative frequencies, in this situ- ation the ordering is nm > Ωeq > n∗, and therefore the planet will be spun down by the star but spun up by the moon. Thus we find that an equilibrium is reached, but the equilibrium is not stable as summarized in Figure 4. Since nm > Ωeq, the moon torques up the planet. This then moves the moon closer to the planet to conserve an- gular momentum. With the new orbital separation, once again nm > Ωeq, and thus the process repeats until the moon moves close enough to the planet to be disrupted. This is seen in the middle panel of Figure 3, which plots the evolution of the moon's orbital separation (red solid line). At early times, am increases as the planet spins down and donates its angular momentum to the moon's orbit until leveling off when P ≈ Peq. This does not last though because of the unstable situation and eventually am comes crashing back to the planet. Also plotted in Figure 3 is the critical radius for the moon to be lost to the parent star acrit,m (black dashed line) given by Equation (1). In this particular case, am < acrit,m for the entire evolution, but if this critical radius is ever exceeded, one should expect the moon to be tidally stripped from the planet. Finally, in the bottom panel of Figure 3, I summarize the tidal heating rates. At early times, this is dominated by the star, and at late times the moon. But in either case, it must remain non-zero because the combined tides always make sure the planet is asynchronous with the orbits of both the star and moon. Also plotted is the heat flux of ≈ 4 × 1020 erg s−1 that is currently emanat- ing through the Earth (black dashed line, Dye 2012) due to a combination of radioactivity, latent heat, and heat left over from the Earth's formation. This demonstrates that the tidal heating is actually similar or exceeds this value for a timescale of ≈ 108 yrs, and thus one might expect such a planet to have tectonic activity similar to the Earth (or perhaps even Io, which has an even greater heating rate) during this time. Although compared to the timescale necessary for life to develop, this appears too short (albeit, such a timescale is obviously very un- certain). 5.2. Is the Moon Disrupted or Lost? Fig. 3.- An example time evolution of the star-planet-moon sys- tem, using the parameters M∗ = 0.5 M(cid:12), Mp = M⊕, and Mm = MM, with initial orbital separations a∗ = 0.3 AU and am = 0.5 aM, and an initial spin period for the planet of P0 = 7 hrs. The top panel summarizes the main periods of the system Peq = 2π/Ωeq (dashed turquoise line), P = 2π/Ω (solid red line), 4π/ω∗ (dotted blue line), and 4π/ωm (long dashed purple line). The middle panel is the moon's orbital separation (red solid line) in comparison to the critical radius acrit,m (dashed black line) at which the moon would be tidally stripped by the star from Equation (1). The bot- tom panel compares tidal heating rates, which includes the heating from the star E∗ (dotted blue line), the moon Em (long dashed pur- ple line), and the total heating (solid red line). Also plotted is the current heat flux coming up through the Earth of ≈ 4×1020 erg s−1 (dashed black line). period of the planet P = 2π/Ω (solid red line), the equi- librium spin period related to the analytic solutions from Section 4.2 Peq = 2π/Ωeq (dashed turquoise line), 4π/ω∗ (dotted blue line), and 4π/ωm (long dashed purple line). These latter two periods are the periods of the tidal forc- ing from the star and moon, respectively, with an extra factor of 2 in each case to cancel the factor of 2 from the two tidal bulges. Also, for these periods the absolute value is plotted since they can be either negative or pos- itive. At early times we see that P ≈ 4π/ω∗ ≈ 4π/ωm. This is because the fast initial spin of the planet dom- inates setting the tidal frequencies. Furthermore, both of the toques are negative, and the planet is spinning down. At a time of ≈ 2 × 108 yrs, the planet has spun down sufficiently that Ω < nm, which causes the torque from the moon to instead want to spin the planet up. This switch in the sign of the moon's torque can be seen in the plot of 4π/ωm, which shows a cusp at this time. Now the combined torques of the moon (spinning the planet up) and the star (spinning the planet down) push the planet's spin toward P ≈ Peq. Even though the planet has reached this spin equilib- rium, the story is not over because this equilibrium is Moon torquesup planetMoon movescloser to conserveangular momentumNew spinequilibrium is slowerthan moon's orbit 7 Equating this to acrit,m given by Equation (1) and solving for P0 provides the critical initial period for the planet Pcrit,0 = (cid:18) 3M∗ (cid:19)1/3(cid:18) Mm (cid:17)−1/2(cid:18) M∗ (cid:19)1/6 (cid:19)−1(cid:18) Rp (cid:19)2 (cid:19)1/6 MM R⊕ Mp , 2πλM 1/2 p R2 p (Gf a∗)1/2Mm (cid:18) Mp ×(cid:16) a∗ M⊕ = 5.7 0.3 AU 0..5 M(cid:12) hrs. (32) This matches fairly closely the critical value for the initial period found from the evolutions in Figure 5. In detail, this critical period does not exactly hold. This is because it matters how much angular momentum the moon initially has as well. For example, in Figure 6 for each of the evolutions the initial spin period of the planet is set to P0 = 9 hrs (in Figure 5, this P0 was found to be sufficiently long that the moon is expected to mi- grate back into the planet). But varying the moon's ini- tial orbital separation from 0.2aM to 0.7aM demonstrates that if the moon is initially sufficiently far, then it will be easier to strip. This effect is not captured by Equation (32). Another physical mechanism that will play a role in determining the fate of the moon is the role of tidal resonances. Of particular interest may be the evection resonance, where the moon's perihelion precession rate becomes equal to the orbital period of the planet (e.g., Touma & Wisdom 1998). This can alter the moon's orbit dramatically, but it depends in detail on the rate of the orbital evolution and the width of the resonance. Due to these complications, I save a more detailed study of these processes for future work. 5.3. Fate of the Disrupted Moon In the cases where the moon is forced to tidally migrate back toward the planet, there are likely two potential outcomes: (1) the moon is tidally disrupted, or (2) the moon directly impacts the planet. Here it is argued that at least for a system similar to the Earth-Moon, tidal disruption is more likely. The moon will migrate inward until its radius hits the Roche lobe, i.e., the equipotential surface where material will no longer be gravitationally bound to the moon. As long as Mm/Mp (cid:28) 1, this can be approximated by the condition (Frank et al. 2002) (cid:18) (cid:19)−1/3 Rm = 0.462 1 + Mp Mm at ≈ 0.462 (cid:18) Mm (cid:19)1/3 Mp at, (33) where at denotes the semi-major axis when tidal disrup- tion occurs. At the moment of disruption, the L1 La- grange point is located a distance dt ≈ 0.7(Mm/Mp)1/3at from the center of mass of the moon. Thus, for tidal dis- ruption to occur, then at (cid:38) Rp + dt, otherwise the planet and moon are too close together and direct impact will occur instead. Rewriting this inequality by making use of Equation (33), Fig. 5.- Similar to Figure 3, but varying the initial spin period P0 of the planet with values of 3 hrs (blue curves), 5 hrs (turquoise curves), 7 hrs (green curves), 9 hrs (yellow curves), 11 hr (orange curves), and 13 hrs (red curves). All other parameters are the same as in Figure 3. This demonstrates when P0 = 7 hrs, the moon will likely be disrupted, but if the initial spin is shortened to P0 = 5 hrs, then the moon is likely stripped by the star. In the example from the previous section and in Figure 3, it appears the moon will be forced to migrate into the planet rather than expelled by the star. But this example also showed that the moon was fairly close to being removed by the star if only the moon migrated out a little further (compare the red solid line and black dashed lines in the middle panel of Figure 3). To better explore what controls the fate of the moon, in Figure 5 we plot the evolution of the star-planet- moon system for a variety of different initial spins for the planet. The initial spins vary from from periods of 3 hrs (blue curves) to 13 hrs (red curves). This demonstrates that the initial spin of the planet can have a dramatic affect on the fate of the moon. The reason is that the moon's migration is driven by the extraction of angular momentum from the planet's spin, and the more spin the planet has, then the further out the moon will migrate. This dependency on P0 is not emphasized in the work of Sasaki et al. (2012), which used a constant value of P0 = 6 hrs for Earth-like planets (although see some of the discussion in Sasaki & Barnes 2014). Whether the moon is disrupted or lost can be addressed with some simple arguments. The initial angular momen- tum in the planet's spin is 2πλMpR2 p/P0. If all of this angular momentum goes into the moon's orbit, with an- gular momentum Mm(GMpam)1/2, then the orbital sep- aration of the moon would be am = 4π2λ2MpR4 p GM 2 mP 2 0 . (31) 2.16 + 1.51. (34) (cid:18) Mp Mm (cid:19)1/3 (cid:38) Rp Rm 8 Fig. 6.- Similar to Figure 3, but with P0 = 9 hrs and varying the orbital separation of the moon over the values 0.7aM (blue curves), 0.6aM (turquoise curves), 0.5aM (green curves), 0.4aM (yellow curves), 0.3aM (orange curves), and 0.2aM (red curves). All other parameters are the same as in Figure 3. This demonstrates that for 0.7aM, the moon is sufficiently far away to begin with that only with a small amount of migration it will be stripped by the star. Simply using the properties of the Earth and Moon, the left side is 9.35 while the right side is 5.19, thus the result is tidal disruption. Another way to understand this is to multiple both sides of this expression by Rm/Rp, which results in (cid:18) (cid:104)ρp(cid:105) (cid:19)1/3 (cid:38) 0.46 + 0.70 Rm Rp , (cid:104)ρm(cid:105) (35) where (cid:104)ρp(cid:105) and (cid:104)ρm(cid:105) are the average density of the planet and moon, respectively. Since the righthand side of Equation (35) has a value between 0.46 to 1.16 (since Rm < Rp), this demonstrates that as long as the den- sity of the planet is similar or greater than the moon, then tidal disruption is expected. Conversely, direct im- pact only occurs when the density of the planet is much less than the moon (similar arguments were presented in Metzger et al. 2012, in a somewhat different context). Following disruption, the moon will likely form some sort of ring-like structure around the planet. This may indicate that some rocky planets around low mass stars should be expected to have circumplanetary rings. To get some idea of what such rings may look like, consider that at the moment of disruption at a separation at, the orbital angular momentum of the moon is Jt = [G(Mp + Mm)at]1/2Mm ≈ 1.47[G(Mp + Mm)Rm]1/2M 1/6 p M 5/6 m . (36) where for the second expression I have used Equation Fig. 7.- Same as Figure 5, but with a∗ = 0.6 AU. This greatly extends the evolution to longer timescales and the moon gets tidally disrupted for a larger range of initial spin periods for the planet P0. Fig. 8.- Same as Figure 5, but with M∗ = 0.2 M(cid:12) and a∗ = 0.1 AU. The small separation greatly speeds up the evolution. (33). For the values of the Earth-Moon system, this gives an angular momentum of Jt ≈ 6 × 1040 erg s. This same amount of angular momentum should be roughly stored in the resulting rings, which would be given by Jr ≈ (GMpRr)1/2Mr, (37) 9 where Rr and Mr are the radius and mass of the rings, respectively. The actual mass that goes into the rings would depend on the details of the ring dynamics, but even if ≈ 10% of the moon's mass went into the rings it would imply a radius of Rr ≈ 2 × 1011 cm (and an even larger radius if the ring mass is smaller). Such a large radius cannot be maintained for the rings. Material interior to the so-called fluid Roche limit will remain in rings, where the fluid Roche limit is given by (Murray & Dermott 1999) (cid:19)1/3 (cid:18)(cid:104)ρp(cid:105) (cid:104)ρ(cid:105) , RFRL = 2.46Rp (38) where (cid:104)ρ(cid:105) is the average density of the material that makes up the rings. This results in typical ring radii of ≈ 2×109 cm. Searching for such rings (using the methods described in, for example, Barnes & Fortney 2004; Ohta et al. 2009; Zuluaga et al. 2015) may verify that pro- cesses as discussed here took place in a specific system. Exterior to the fluid Roche limit, material can coalesce into a new moon that is some fraction of the mass of the moon before. This new moon could then migrate inward again and the process repeat. This suggests that the planet may go through phases where it alternatively has a moon or rings, which has actually been suggested to be the case for Mars and Phobos (Hesselbrock & Minton 2017). More work is needed to better understand the evolution and duty cycle of such rings. 5.4. Exploration of More Evolutions The above discussions spell out some of the general features of the evolution, but there are many parameters that can be varied for a star-planet-moon system. Thus here I highlight some other example evolutions to provide some sense to the diversity of potential results. In each example, I consider a range of values for P0, since this factor has proven to be key in determining the moon's fate. In the above examples, I set a∗ = 0.3 AU for the initial separation so that the planet would be within the hab- itable zone, but what if the planet is much further out? In Figure 7, I consider such a case with a∗ = 0.6 AU. The two main results of this change are that (1) the moon avoids being tidally stripped by the star for a larger range of initial spin periods for the planet and (2) it takes a much longer timescale for the moon to mi- grate back into the planet. For the first case, this is simply because the ability of the star to strip the moon depends most strongly on the separation, as shown by Equation (1). For the second case, this is because the tidal timescales depend on a high power of the separa- tion. In addition, tidal heating also stays strong for a longer period of time, exceeding the current heat flux on Earth for up to ∼ 109 yrs. In Figure 8, I consider the case of a smaller star with M∗ = 0.2 M(cid:12), which requires an initial separation of Fig. 9.- Same as Figure 5, but with M∗ = 0.2 M(cid:12) and In comparison to Figure 8, the larger separation a∗ = 0.3 AU. dramatically lengthens the evolution. a∗ = 0.1 AU for the planet to be near the habitable zone. Although the mass of the star is smaller, the planet's or- bital separation plays a much stronger role in setting the tidal interaction timescales. Thus, the evolution occurs much faster in this case with the fate of the moon be- ing decided within ∼ 106 yrs. In comparison, in Figure 9, just by tripling the initial separation to a∗ = 0.3 AU, now the evolution can occur for 1010 yrs or longer. 6. DISCUSSION AND CONCLUSIONS In this work, I have investigated the tidal interac- tions of star-planet-moon systems. In comparison to the broader work of Sasaki et al. (2012), this study has a more specific focus on Earth-Moon-like systems around low mass stars, which is motivated by recent surveys find- ing planets in the habitable zones of these stars. Fur- thermore, since these habitable zones are relatively close to the star, the tidal interactions between the planet and star are naturally comparable to the interactions between the planet and moon. I especially highlight the role of the initial spin period of the planet P0 in determining the fate of the moon, and I use a constant τ formalism rather than a quality factor for assessing the impact of the tides. Solving for the time evolution of these systems, my main conclusions are as follows. • The combined tidal interactions cause the moon to eventually be stripped by the star or migrate back toward the planet. • Which of these fates befall the moon depends sen- sitively on the initial spin period of the planet P0, 10 with a small spin making it more likely for the moon to be stripped. • In cases where the moon migrates into the planet, the moon will be tidally disrupted rather than di- rectly impact the planet because of the relatively similar densities of the rocky planet and moon. This may produce rings around rocky planets. • The combined tidal interactions force the planet to always spin asynchronously with respect to the both the moon and star, often generating an amount of tidal heating similar to the current heat flux coming up through the Earth for up to ∼ 109 yrs. • The overall evolution of this system until the time the moon is stripped or disrupted depends on the initial separation of the planet and star, and can be very greatly from less than 106 yrs to greater than 1010 yrs. In the future, as extrasolar moons are inevitable discov- ered, the formalism presented here can be used to bet- ter understand the lifetime and fate of these moons (al- though in some cases additional planetary bodies may also impact the moon, something outside the scope of this work). If instead the presence of rings around rocky planets orbiting low mass stars is found, it would pro- vide evidence that processes as described here have oc- curred in specific systems. Alternatively, if no rings or moons are ever present, it could indicate that the moons are stripped because of the short initial spin when the planet is formed, providing insight into the planet for- mation process. Further calculations are needed to un- derstand the details of these rings, how they evolve, and how long they should be present. It is interesting to ponder the implications of the long timescale found for the loss of the moon through tidal dis- ruption or stripping. One might assume that the fate of a moon would be determined relatively early during the formation process of the solar system and planet. But in fact, without fine tuning the parameters much, a moon could orbit a planet for well over 1010 yrs before being tidally disrupted or lost from the planet completely! If such a planet could harbor life, this would presumably be sufficiently long for an advanced civilization to develop, only to be subject to a catastrophic event. Neverthe- less, as we have seen over the last century, technology advances quickly. And although our species still seems to be sorting out issues more local to home, one might hope that in a relatively short amount of time in com- parison to astrophysical timescales, an advanced society may be able to overcome such a unique challenge. I thank Johanna Teske for suggestions on a previous draft of this manuscript, and Jason Barnes for answer- ing my questions about his work. I also thank Kon- stantin Batygin, Alexandre Correia, Michael Efroimsky, Jim Fuller, and Valeri Makarov for their feedback. REFERENCES Adams, F. C., & Bloch, A. M. 2016, MNRAS, 462, 2527 Barnes, J. W., & Fortney, J. J. 2004, ApJ, 616, 1193 Barnes, J. W., & O'Brien, D. P. 2002, ApJ, 575, 1087 Burke, C. J., Christiansen, J. L., Mullally, F., et al. 2015, ApJ, 809, 8 Counselman, III, C. C. 1973, ApJ, 180, 307 Domingos, R. C., Winter, O. C., & Yokoyama, T. 2006, MNRAS, 373, 1227 Driscoll, P. E., & Barnes, R. 2015, Astrobiology, 15, 739 Dye, S. T. 2012, Reviews of Geophysics, 50, RG3007 Efroimsky, M., & Makarov, V. V. 2013, ApJ, 764, 26 Foley, B. J., & Driscoll, P. E. 2016, Geochemistry, Geophysics, Geosystems, 17, 1885 Frank, J., King, A., & Raine, D. J. 2002, Accretion Power in Astrophysics: Third Edition, 398 Goldreich, P. 1966, AJ, 71, 1 Heller, R., & Barnes, R. 2013, Astrobiology, 13, 18 Heller, R., Leconte, J., & Barnes, R. 2011, A&A, 528, A27 Henning, W. G., O'Connell, R. J., & Sasselov, D. D. 2009, ApJ, 707, 1000 Hesselbrock, A. J., & Minton, D. A. 2017, Nature Geoscience, 10, 266 Holman, M. J., & Wiegert, P. A. 1999, AJ, 117, 621 Hut, P. 1981, A&A, 99, 126 Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108 Lambeck, K. 1977, Philosophical Transactions of the Royal Society of London Series A, 287, 545 Li, G., & Batygin, K. 2014, ApJ, 790, 69 Lissauer, J. J., Barnes, J. W., & Chambers, J. E. 2012, Icarus, 217, 77 MacDonald, G. J. F. 1964, Reviews of Geophysics and Space Physics, 2, 467 Metzger, B. D., Giannios, D., & Spiegel, D. S. 2012, MNRAS, 425, 2778 Mignard, F. 1979, Moon and Planets, 20, 301 -. 1980, Moon and Planets, 23, 185 -. 1981, MNRAS, 194, 365 Mulders, G. D., Pascucci, I., & Apai, D. 2015, ApJ, 798, 112 Murray, C. D., & Dermott, S. F. 1999, Solar system dynamics Neron de Surgy, O., & Laskar, J. 1997, A&A, 318, 975 Ogilvie, G. I. 2014, ARA&A, 52, 171 Ohta, Y., Taruya, A., & Suto, Y. 2009, ApJ, 690, 1 Ray, R. D., & Egbert, G. D. 2012, Geophysical Journal International, 189, 400 Sasaki, T., & Barnes, J. W. 2014, International Journal of Astrobiology, 13, 324 Sasaki, T., Barnes, J. W., & O'Brien, D. P. 2012, ApJ, 754, 51 Sleep, N. H., & Zahnle, K. 2001, J. Geophys. Res., 106, 1373 Touma, J., & Wisdom, J. 1994, AJ, 108, 1943 -. 1998, AJ, 115, 1653 Walker, J. C. G., Hays, P. B., & Kasting, J. F. 1981, J. Geophys. Res., 86, 9776 Ward, P., & Brownlee, D. 2000, Rare earth : why complex life is uncommon in the universe Ward, W. R., & Reid, M. J. 1973, MNRAS, 164, 21 Williams, J. G. 1994, AJ, 108, 711 Zuluaga, J. I., Kipping, D. M., Sucerquia, M., & Alvarado, J. A. 2015, ApJ, 803, L14
1001.1617
1
1001
2010-01-11T10:11:20
Numerical Simulations of Highly Porous Dust Aggregates in the Low-Velocity Collision Regime
[ "astro-ph.EP" ]
A highly favoured mechanism of planetesimal formation is collisional growth. Single dust grains, which follow gas flows in the protoplanetary disc, hit each other, stick due to van der Waals forces and form fluffy aggregates up to centimetre size. The mechanism of further growth is unclear since the outcome of aggregate collisions in the relevant velocity and size regime cannot be investigated in the laboratory under protoplanetary disc conditions. Realistic statistics of the result of dust aggregate collisions beyond decimetre size is missing for a deeper understanding of planetary growth. Joining experimental and numerical efforts we want to calibrate and validate a computer program that is capable of a correct simulation of the macroscopic behaviour of highly porous dust aggregates. After testing its numerical limitations thoroughly we will check the program especially for a realistic reproduction of various benchmark experiments. We adopt the smooth particle hydrodynamics (SPH) numerical scheme with extensions for the simulation of solid bodies and a modified version of the Sirono porosity model. Experimentally measured macroscopic material properties of silica dust are implemented. We calibrate and test for the compressive strength relation and the bulk modulus. SPH has already proven to be a suitable tool to simulate collisions at rather high velocities. In this work we demonstrate that its area of application can not only be extended to low-velocity experiments and collisions. It can also be used to simulate the behaviour of highly porous objects in this velocity regime to a very high accuracy.The result of the calibration process in this work is an SPH code that can be utilised to investigate the collisional outcome of porous dust in the low-velocity regime.
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. technicalpaperv10 June 15, 2018 c(cid:13) ESO 2018 Numerical Simulations of Highly Porous Dust Aggregates in the Low-Velocity Collision Regime Implementation and Calibration of a Smooth Particle Hydrodynamics Code R. J. Geretshauser1, R. Speith1, C. Guttler2, M. Krause2, and J. Blum2 1 Institut fur Astronomie und Astrophysik, Abteilung Computational Physics, Eberhard Karls Universitat Tubingen, Auf der Morgenstelle 10, D-72076 Tubingen, Germany e-mail: [email protected] Braunschweig, Germany 2 Institut fur Geophysik und extraterrestrische Physik, Technische Universitat zu Braunschweig, Mendelssohnstr. 3, D-38106 Preprint online version: January 11, 2010; accepted by Astronomy & Astrophysics December 22, 2009 ABSTRACT Context. A highly favoured mechanism of planetesimal formation is collisional growth. Single dust grains hit each other due to relative velocities caused by gas flows in the protoplanetary disc, which they follow. They stick due to van der Waals forces and form fluffy aggregates up to centimetre size. The mechanism of further growth is unclear since the outcome of aggregate collisions in the relevant velocity and size regime cannot be investigated in the laboratory under protoplanetary disc conditions. Realistic statistics of the result of dust aggregate collisions beyond decimetre size is missing for a deeper understanding of planetary growth. Aims. Joining experimental and numerical efforts we want to calibrate and validate a computer program that is capable of a correct simulation of the macroscopic behaviour of highly porous dust aggregates. After testing its numerical limitations thoroughly we will check the program especially for a realistic reproduction of the compaction, bouncing and fragmentation behaviour. This will demonstrate the validity of our code, which will finally be utilised to simulate dust aggregate collisions and to close the gap of fragmentation statistics in future work. Methods. We adopt the smooth particle hydrodynamics (SPH) numerical scheme with extensions for the simulation of solid bodies and a modified version of the Sirono porosity model. Experimentally measured macroscopic material properties of SiO2 dust are implemented. By simulating three different setups we calibrate and test for the compressive strength relation (compaction experiment) and the bulk modulus (bouncing and fragmentation experiments). Data from experiments and simulations will be compared directly. Results. SPH has already proven to be a suitable tool to simulate collisions at rather high velocities. In this work we demonstrate that its area of application can not only be extended to low-velocity experiments and collisions. It can also be used to simulate the behaviour of highly porous objects in this velocity regime to a very high accuracy. A correct reproduction of density structures in the compaction experiment, of the coefficient of restitution in the bouncing experiment and of the fragment mass distribution in the fragmentation experiment show the validity and consistency of our code for the simulation of the elastic and plastic properties of the simulated dust aggregates. The result of this calibration process is an SPH code that can be utilised to investigate the collisional outcome of porous dust in the low-velocity regime. Key words. hydrodynamics - methods: laboratory - methods: numerical - planets and satellites: formation - planetary systems: formation - planetary systems: protoplanetary disks 1. Introduction In gaseous circumstellar discs, the potential birthplace of plan- etary systems, dust grains smaller than a micrometre grow to kilometre-sized planetesimals, which themselves proceed to ter- restrial planets and cores of giant planets by gravity-driven run- away accretion. Depending on their size the dust grains and aggregates perform motions in the disc and relative to each other. Brownian motion, radial drift, vertical settling and tur- bulent mixing cause mutual collisions (Weidenschilling 1977; Weidenschilling & Cuzzi 1993). Since real protoplanetary dust particles are not available for experiments in the laboratory much of the following work has been carried out with dust analogues such as SiO2 (Blum & Wurm 2008). Theoretical models also refer to micro- scopic and macroscopic properties of these materials. Initially the dust grains hit and stick on contact by van der Waals forces (Heim et al. 1999). In this process, which has been investi- gated experimentally (Blum et al. 2000; Blum & Wurm 2000; Krause & Blum 2004) and numerically (Dominik & Tielens 1997; Paszun & Dominik 2006, 2008, 2009; Wada et al. 2007, 2008, 2009), they form fluffy aggregates with a high degree of porosity. Due to restructuring the aggregates gain a higher mass to surface ratio and reach higher velocities. Blum & Wurm (2000, 2008) and Wada et al. (2008) showed that collisions among them lead to fragmentation and mass loss. Depending on the model of the protoplanetary disc, which provides the kinetic colli- sion parameters, this means that direct growth ends at aggre- gate sizes of a few centimetres. However, Wurm et al. (2005) and Teiser & Wurm (2009) have demonstrated in laboratory ex- periments in the centimetre regime and with low-porosity dust that the projectile can stick partially to the target at velocities of 0 1 0 2 n a J 1 1 . ] P E h p - o r t s a [ 1 v 7 1 6 1 . 1 0 0 1 : v i X r a 2 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates more than 20 m s−1. Thus, collisional growth beyond centimetre size seems to be possible and the exact outcome of the fragment distribution is crucial for the understanding of the growth mech- anism. Numerical models that try to combine elaborate pro- toplanetary disc physics with the dust coagulation prob- lem (Weidenschilling et al. 1997; Dullemond & Dominik 2005; Brauer et al. 2008; Zsom & Dullemond 2008; Ormel et al. 2007, 2009) have to make assumptions about the outcome of collisions between dusty objects for all sizes and relative velocities. Since data for these are hardly available, in the most basic versions of these models perfect sticking, in more elaborate ones power- law fragment distributions from experiments and observations (Mathis et al. 1977; Davis & Ryan 1990; Blum & Munch 1993) are assumed. The results of alike simulations highly depends on the assumed fragmentation kernel. However, the given experi- mental references have been measured only for small aggregate sizes. The influence of initial parameters such as the rotation of the objects or their porosity have not been taken into account for the size regimes beyond centimetre size, although they might play an important role (Sirono 2004; Ormel et al. 2007). A new approach to model the growth of protoplanetary dust aggregates was recently developed by Guttler et al. (2009a) and Zsom et al. (2009) who directly implemented the results of dust aggregation experiments into a Monte-Carlo growth model. They found that bouncing of protoplanetary dust aggregates plays a major role for their evolution as it is able to inhibit fur- ther growth and changes their aerodynamic properties. Although their model relies on the most comprehensive database of dust aggregate properties and their collisional behaviour, they were unable to make direct predictions for any arbitrary set of colli- sion parameters and were thus obliged to perform extrapolations over orders of magnitude. Some of these extrapolations are based on physical models which need to be supported by further exper- iments and where this is not possible by sophisticated numerical models such as the one presented here. Sticking, bouncing, and fragmentation are the important col- lision outcomes which need to be implemented into a coagu- lation code and affect the results of these models. Thus, it is not only important to correctly implement the exact thresholds between these regimes but also details concerning the outcome such as the fragment size distribution, the compaction in bounc- ing collisions, and maybe even the shape of the aggregates after they merged in a sticking collision. Due to the lack of impor- tant input information the necessity for a systematic study of all relevant collision parameters arises and will be addressed in this work. Because of restrictions in size and realistic environment pa- rameters this task cannot be achieved in the laboratory alone. There has been a lot of work lately on modelling the behaviour of dust aggregates on the basis of molecular interactions be- tween the monomers (Paszun & Dominik 2006, 2008, 2009; Wada et al. 2007, 2008, 2009). However, simulating dust aggre- gates with a model based on macroscopic material properties such as density, porosity, bulk and shear moduli and compres- sive, tensile and shear strengths remains an open field since these quantities are rarely available. The advantage of this approach over the molecular dynamics method, which is computationally limited to a few ten thousand monomers, is the accessibility of aggregate sizes beyond the centimetre regime. Recently, Jutzi et al. (2008) have implemented a porosity model into the smooth particle hydrodynamics (SPH) code by Benz & Asphaug (1994). It was calibrated for pumice material using high-velocity impact experiments (Jutzi et al. 2009b) and utilised to understand the formation of an aster- oid family (Jutzi et al. 2009a). However, pumice is a mate- rial whose strength parameters decrease when it is compacted (crushed). Additionally, the underlying thermodynamically en- hanced porosity model is designed to describe impacts of some km s−1. Thus, it is perfectly suitable to simulate high-velocity collisions of porous rock-like material. In contrast, collisions between pre-planetesimals occur at relative velocities of some tens of m s−1 and compressive, shear and tensile strengths are increasing with increasing den- sity (Blum & Schrapler 2004; Blum et al. 2006; Guttler et al. 2009b). Schafer et al. (2007) have used an SPH code based on the porosity model by Sirono (2004) to simulate collisions be- tween porous ice in the m s−1 regime. They found that a suit- able choice of relations for the material parameters can pro- duce sticking, bouncing or fragmentation of the colliding ob- jects. Therefore, they stressed the importance of calibrating the material parameters of porous matter with laboratory measure- ments. Numerical molecular-dynamics simulations are about to use a sufficient number of monomers to be close enough to the continuum limit and to provide the required material parameters (e.g. Paszun & Dominik 2008, reproducing experimental results of Blum & Schrapler 2004). They represent an important sup- port to the difficult experimental determination of these quanti- ties. In Guttler et al. (2009b) we have measured the compressive strength relation for spherical SiO2 dust aggregates for the static case in the laboratory and have given a prescription how to apply this to the dynamic case using a compaction calibration experi- ment and 2D simulations. We have pointed out relevant bench- mark features. It was shown that the code is in principal capable of simulating not only fragmentation but also bouncing, which cannot be seen in molecular-dynamics codes so far. In this work we present our SPH code with its technical de- tails (Sect. 2), experimental reference (Sect. 3), and numerical properties (Sect. 4). On the basis of the compaction calibration simulation we demonstrate that the results converge for increas- ing spatial resolution and choose a sufficient numerical resolu- tion (Sect. 4.2). We investigate the differences between 2D and 3D numerical setup (Sect. 4.3) and thereby improve some draw- backs in Guttler et al. (2009b). Artificial viscosity will be pre- sented as stabilising tool for various problems in the simulations and its influence on the physical results will be pointed out (Sect. 4.4). Most prominently we continue the calibration process started in Guttler et al. (2009b) utilising two further calibration experi- ments for bouncing (Sect. 5.2) and fragmentation (Sect. 5.3). In the end we possess a collection of material parameters, which is consistent for all benchmark experiments. Finally, the SPH code has gained enough reliability to be used to enhance our infor- mation about the underlying physics of dust aggregate collisions beyond the centimetre regime. In future work it will be applied to generate a catalogue of pre-planetesimal collisions and their outcome regarding all relevant parameters for planet formation. Jointly with experiments and coagulation models this can be im- plemented into protoplanetary growth simulations (Guttler et al. 2009a; Zsom et al. 2009) to enhance their reliability and predic- tive power. R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 3 2. Physical Model and Numerical Method 2.1. SmoothParticleHydrodynamics The numerical Lagrangian particle method smooth particle hy- drodynamics (SPH) was originally introduced by Lucy (1977) and Gingold & Monaghan (1977) to model compressible hydro- dynamic flows in astrophysical applications. Later, the method has been extended (Libersky & Petschek 1990), and improved extensively (e.g. Libersky et al. 1993; Benz & Asphaug 1994; Randles & Libersky 1996; Libersky et al. 1997) to simulate elastic and plastic deformations of solid materials. A compre- hensive description of SPH and its extensions can be found in Monaghan (2005). In the SPH scheme, continuous solid objects are discretized into interacting mass packages, so called "particles". These par- ticles form a natural frame of reference for any deformation and fragmentation that the solid body may undergo. All spatial field quantities of the object are approximated onto the particle posi- tions xi by a discretized convolution with a kernel function W. The kernel W depends on particle distance x j − xi and has com- pact support, determined by the smoothing length h . We are using the standard cubic spline kernel (Monaghan & Lattanzio 1985) but normalised such that its maximum extension is equal to one smoothing length h. We apply a constant smoothing length. This also allows to model fragmentation of solid objects in a simple way. Fragmentation occurs when some SPH particles within the body lose contact with their adjacent particles. Two fragments are completely separated as soon as their respective subsets of parti- cles have reached a distance of more than 2h so that their kernels do not overlap any more. Time evolution of the SPH particles is computed according to the Lagrangian form of the equations of continuum mechan- ics, while transferring spatial derivatives by partial integration onto the analytically given kernel W. 2.2. Continuummechanics A system of three partial differential equations forms the frame- work of continuum mechanics. As commonly known they follow from the constraints of conservation of mass, momentum and en- ergy. The first, accounting for the conservation of mass, is called continuity equation + ρ = 0. ∂vα ∂xα dρ dt Following the usual notation ρ and vα denote density and veloc- ity, respectively. Greek indices run from 1 to d, the dimension of the problem. Einstein summing notation holds throughout the entire paper. In contrast to the usual SPH scheme, where the SPH density ρi is calculated directly from the particle distribution, we solve the continuity equation according to (1) dρi dt = −ρiXj m j ρ j (vα i ) j − vα ∂Wi j(h) ∂xα i (2) The conservation of momentum is ensured by the second equation dvα dt = 1 ρ ∂σαβ ∂xβ . In SPH formulation, the momentum equation reads (3) (4) dvα i dt = Xj m j σαβ i ρ2 i + σαβ j ρ2 j ∂Wi j(h) ∂xβ i .  Due to the symmetry in the interaction terms, conservation of momentum is ensured by construction. Additionally we ap- ply the standard SPH artificial viscosity (Monaghan & Gingold 1983). This is essential in particular for stability at interfaces with highly varying densities. The influence of artificial vis- cosity on our simulation results is investigated thoroughly in Sec. 4.4. The third equation, the energy equation, is not used in our model. Hence, we assume that kinetic energy is mainly con- verted into deformation energy and energy dissipated by viscous effects is converted into heat and radiated away. The stress tensor σ can be split into a part representing the pure hydrostatic pres- sure p and a traceless part for the shear stresses, the so called deviatoric stress tensor S αβ. Hence, σαβ = −pδαβ + S αβ. (5) Any deformation of a solid body leads to a development of in- ternal stresses in a specifically material dependent manner. The relation between deformation and stresses is not taken into ac- count within the regular equations of fluid dynamics. Therefore, they are insufficient to describe a perfectly elastic body and have to be extended. The missing relations are the constitutive equa- tions, which depend on the strain tensor ǫαβ = 1 2 ∂x′α ∂xβ + ∂x′β ∂xα ! . (6) It represents the local deformation of the body. The primed co- ordinates denote the positions of the deformed body. Following Hooke's law a proportional relation between de- formation is assumed involving the material dependent shear modulus µ = µ(ρ), which depends itself on the density: S αβ ∝ 2µ ǫαβ − 1 d δαβǫγγ! . (7) However, this is only the constitutive equation for the traceless shear part. For the hydrostatic part of the stress tensor we adopt a modification of the Murnaghan equation of state which is part of the Sirono (2004) porosity model: p(ρ) = K(ρ′ 0)(ρ/ρ′ 0 − 1), (8) where ρ′ material at zero external stress, and K(ρ) is the bulk modulus. 0 is the so called reference density, the density of the The density dependence of the bulk K(ρ) and shear µ(ρ) moduli is modelled by a power law (e.g. Randles & Libersky 1996). Here the sum runs over all in- teraction partners j of particle i, m j is the particle mass of par- ticle j, and Wi j denotes the kernel for the particular interaction. Although this approach is more expensive, as it requires to solve an additional ordinary differential equation for each particle, it is more stable for high density contrasts and it avoids artifacts due to smoothing at boundaries and interfaces. K(ρ) = 2µ(ρ) = K0(ρ/ρi)γ. (9) Although according to Sirono (2004) ρi is the initial density of the material at the beginning of the simulation, we, in contrast, want to ensure that our dust material possesses the same bulk modulus K(ρ) even for simulations with different initial densi- ties. In this work the dust material has two different densities 4 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates at the beginning of the bouncing (Sect. 5.2) and fragmentation (Sect. 5.3) calibration setup. According to Sirono (2004) the ma- terials should feature two different ρi. As a consequence K(ρ), and in particular K0, depends on the initial setup. Since we want to compare and validate K0 by using two different setups we have to fix a unique ρi for all simulations. We choose ρi such that K(ρi) = K0 is the bulk modulus of the generic uncompressed dust material that is produced by the random ballistic deposition (RBD) method (Blum & Schrapler 2004). The time evolution of the pressure is directly given by the time evolution of the density (Eq. 1) and since the pressure is a scalar quantity it is intrinsically invariant under rotation. However, in order to gain a frame invariant formulation of the time evolution of the deviatoric stress tensor, i.e. the stress rate, correction terms have to be added. A very common formulation for SPH (see e.g. Benz & Asphaug 1994; Schafer et al. 2007) is the Jaumann rate form dS αβ = 2µ ǫαβ − 1 d δαβ ǫγγ! + S αγRγβ + S βγRγα, (10) dt where Rαβ is the rotation rate tensor, defined by Rαβ = 1 2 ∂vα ∂xβ − ∂vβ ∂xα! and ǫαβ denotes the strain rate tensor ǫαβ = 1 2 ∂vα ∂xβ + ∂vβ ∂xα! . (11) (12) i /∂xβ To determine rotation rate and strain rate tensor and thus the evo- lution of the stress tensor Eq. 10 in SPH representation, the SPH velocity derivatives ∂vα i have to be calculated. The standard SPH formulation, however, does not conserve angular momen- tum due to the discretization error by particle disorder, which leads to a rotational instability and in particular inhibits mod- elling rigid rotation of solid bodies. To avoid this error, we apply a correction tensor Cγβ (Schafer et al. 2007) according to ∂vα i ∂xβ i i )Xγ ∂Wi j ∂xγ i = Xj j − vα m j ρ j Cγβ, (13) (vα where the correction tensor Cγβ is the inverse of 2.3. Porosityandplasticity Following the Sirono (2004) model, the porosity Φ is modelled by the density of the porous material ρ and the constant matrix density ρs Φ = 1 − ρ ρs = 1 − φ, (17) In the following we will use the filling factor φ = ρ/ρs. We model plasticity by reducing inner stresses given by σαβ. For this reason we need constitutive relations describing the be- haviour of the material during plastic deformation. These rela- tions are specific for each material and have to be determined empirically. Particularly for highly porous materials it is ex- tremely difficult to acquire them. Therefore, it is an advanta- geous feature of our model that it is based on measurements by Blum & Schrapler (2004), Blum et al. (2006) and Guttler et al. (2009b). The main idea of the adopted plasticity model is to reduce inner stress once the material exceeds a certain plasticity cri- terion. In the elastic case, described by Eqns. 7 and 8, inner stresses grow linearly with deformation. Hence, the material re- turns to its original shape at vanishing external forces. Reducing inner stresses, i.e. deviating from the elastic deformation path, reduces the internal ability of the material to restore its original shape. Therefore, by stress reduction deformation becomes per- manent, i.e. plastic. Following and expanding the approach by Sirono (2004), we treat plasticity for the pure hydrostatic pres- sure p and the deviatoric stress tensor S αβ separately. For the deviatoric stress tensor we follow the approach by Benz & Asphaug (1995) and Schafer et al. (2007) assuming that our material is isotropic, which makes the von Mises yield cri- terion applicable. This criterion is characterised by the shear strength Y, which in our model is a composite of the compres- sive and tensile strengths: Y(φ) = pΣ(φ) T (φ). The suitability of this choice was already demonstrated in Guttler et al. (2009b). Since Y(φ) is a scalar we have to derive a scalar quantity from S αβ for reasons of comparability. We do this by calculating its second irreducible invariant J2 = S αβS αβ. Finally, the reduction of the deviatoric stress is implemented in the following way (xα j ) i − xα ∂Wi j ∂xγ i , (14) S αβ → f S αβ, where f = minhY2(φ)/3J2, 1i . (18) The hydrostatic pressure is limited by the tensile strength T (φ) for p < 0 and by the compressive strength Σ(φ) for p > 0: p(φ) = ( Σ(φ) φ > φ+ c T (φ) φ < φ− c . (19) c ≤ φ ≤ φ+ c the material is in the elastic regime and Eq. For φ− 8 is applied. φ− c denote the filling factors where the elas- tic path intersects the tensile strength and compressive strength, respectively (see Fig. 1). c and φ+ The pressure reduction process is implemented such that at any time step p is computed with Eq. 8. If for a given φ p(φ) > Σ(φ) and φ > φ+ c the pressure p(φ) is reduced to Σ(φ). The deformation becomes irreversible once the new reference density ρ′ 0 is computed through Eq. 8 and the elastic path is shifted towards higher densities. Hereby also the limiting filling factors φ− c are set anew. In principal there are two pos- sible implementations for this. (1) Plasticity becomes effective immediately and ρ′ 0 is computed whenever p > Σ. (2) Plasticity becomes effective after pressure decrease, which is equivalent to c . We tested both implementations. For our understanding φ < φ+ c and φ+ m j ρ j Xj that is m j ρ j Xj (xα j − xα i )Xγ ∂Wi j ∂xγ i Cγβ = δαβ. (15) This approach leads by construction to first order consistency where the errors due to particle disorder cancel out and the con- servation of angular momentum is ensured. Only this allows that rigid rotation can be simulated correctly. Finally, the sound speed of the material is given by (16) cs(ρ′ 0) = qK(ρ′ 0)/ρ′ 0. Together with Eq. 9 this relation shows that the soundspeed is a strong function of density. This behaviour has been seen in moleculardynamics simulations by Paszun & Dominik (2008), but there is no data available from laboratory measurements. Up to this point the set of equations describes a perfectly elastic solid body. Additionally, the material simulated in this work are SiO2 dust aggregates, which features a high degree of porosity and, thus, plasticity. The modifications accounting for these features are described in the next section. R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 5 p φ− c φ′ 0 φ+ c plastic tension i o n s i c t s p l a c o m p r e s atio elastic eform φ′ 1 n d Σ(φ) φ T (φ) Fig. 1. In the modified Sirono porosity model the regime of elastic deformation is limited by the compressive strength Σ(φ), which represents the transition threshold to plastic compression for p > Σ(φ), and the tensile strength T (φ), which represents the transition threshold to plastic tension or rupture for p < T (φ). Returning from one of the plastic regimes to vanishing external pressure via an elastic path leads to a φ′ 1 that differs from the initial φ′ 0. Hence, the material has been deformed irreversibly. possibility (1) is closer to the underlying physical process. In ad- dition it proved to be more stable. According to the benchmark parameters the results were equivalent. For the tensile regime, i.e. for φ < φ− c , we do not adopt the damage and damage restoration model presented in Sirono (2004). This damage model for brittle material such as rocks or pumice has been developed for SPH by Benz & Asphaug (1994, 1995) and was recently used by Jutzi et al. (2008, 2009a,b). It is assumed that a material contains flaws, which are acti- vated and develop under tensile loading (Grady & Kipp 1980). Schafer et al. (2007) did not find the model applicable in their simulations of porous ice because it includes compressive dam- age effects. Brittle material like pumice and rocks tend to dis- integrate when they are compressed: they crush. Whereas in our material, highly porous SiO2 dust, tensile and compressive strength increase with compression. This is due to the fact that the monomers are able to form new bonds when they get in con- tact. Therefore, we adopt the same approach as in the compres- sive regime and reduce the pressure p(φ) to T (φ) once the ten- sile strength is exceeded. Finally, the material can rupture due to plastic flow. However, material that is plastically stretched can be compressed again up to its full strength. By choosing this ap- proach a "damage restoration model" is implemented in a very natural way. Finally, a remark has to be made about energy. Apart from energy dissipation by numerical and artificial viscosity we as- sume intrinsic energy conservation. We suppose that heat pro- duction in the investigated physical processes is negligible. Therefore, our model is limited to a velocity regime below the sound speed cs of the dust material (≈ 30 m/s, see Blum & Wurm 2008; Paszun & Dominik 2008). By choosing the approach of modelling plasticity via stress reduction, we assume that most of the energy is dissipated by plastic deformation, since the re- duction of internal stresses accounts for the reduction of internal energy. Since we do not solve the energy equation thermodynami- cally enhanced features like any phase transition such as melting and freezing cannot be simulated. This scheme also does not fea- ture a damage model. Especially in the section about fragmenta- tion (Sect. 5.3) the presence of any flaws in the material cannot be taken into account yet, although they might have influence on the resulting fragment distribution. 3. Experimental reference 3.1. Materialparameters The material used for the calibration experiments are highly- porous dust aggregates as described by Blum & Schrapler (2004), consisting of spherical SiO2 spheres with a diameter of 1.5 µm. For these well defined dust aggregates it is possible to reproducibly measure macroscopic material parameters like tensile strength, compressive strength, and, potentially, also the shear strength as needed for the SPH porosity model (see sec- tion 2.3). The tensile strength for this material was measured for highly porous and compacted aggregates (φ = 0.15 .. 0.66) by Blum & Schrapler (2004). These measurements support a linear dependence between the tensile strength and the number of con- tacts per monomer (increasing with increasing φ), which yields the tensile strength as T (φ) = −(cid:16)102.8+1.48φ(cid:17) Pa . (20) The compressive strength was measured in the experimental counterpart of this paper (Guttler et al. 2009b) with an experi- mental setup to determine the static omni-directional compres- sion (ODC), whereby the sample is enclosed from all sides and the pressure is constant within the sample. We found that the compressive strength curve can be well described by the analytic function Σ(φ) = pm · φ2 − φ1 φ2 − φ − 1!∆·ln 10 , (21) where the free parameters were measured to be φ1 = 0.12, φ2 = 0.58, ∆ = 0.58, and pm = 13 kPa. However, these parame- ters were measured with a static setup and we expect a different strength for a dynamic collision. The parameters φ1 and φ2 deter- mine the range of the volume filling factor, where φ1 is defined by the uncompressed material and φ2 corresponds to the densest packing. These parameters are expected to be the same for the dynamic compression, while pm and ∆ are treated as free param- eters, which we will calibrate in the forthcoming sections. The last important material parameter to describe the dust aggregates is the elasticity. We can determine the Young's modulus from measurement and simulation of the sound speed (Blum & Wurm 2008; Paszun & Dominik 2008), which is c = 30 m s−1. From this approach the bulk modulus for the uncompressed material is K0 = ρic2 = 300 kPa with ρi = 300 kg m−3. However, other plausible calculations (Weidling et al. 2009) indicate an elastic- ity of rather 1 kPa. We will therefore also vary this parameter. 3.2. Calibrationexperiment As a setup for an easy and well-defined calibration experiment (see Guttler et al. 2009b), we chose a glass bead with a diame- ter of 1 to 3 mm, which impacts into the dust aggregate material with a velocity between 0.1 and 1 m s−1 under vacuum condi- tions (pressure 0.1 mbar). We were able to measure the deceler- ation curve, stopping time, and intrusion depth of the glass bead 6 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates (for various velocities and projectile diameters) and the com- paction of the dust under the glass bead (for a 1.1 mm projectile with a velocity of 0.65 m s−1). These results will serve for cali- brating and testing the SPH code. For the measurement of the deceleration curve, we used an elongated epoxy projectile instead of the glass bead. The bottom shape and the mass resembled the glass bead, while the lower density and the therefore longer extension made it possible to observe the projectile during the intrusion. The projectile was observed by a high-speed camera (12,000 frames per second) and the position of the upper edge was followed with an accu- racy of ≈ 3 µm. We found that – independently of velocity and projectile diameter – the intrusion curve can be described by a sine curve h(t) = −D · sin π 2 t T s! ; t ≤ T s. (22) Here, D and T s denote the intrusion depth and the stopping time of the projectile. The stopping time is defined as the time be- tween the first contact and the deepest intrusion at zero velocity. In Sect. 5.1.2 we will use the normalised form of the intrusion curve with D = T s = 1. For the intrusion depth we found good agreement with a lin- ear behaviour of D = 8.3 · 10−4 m2 s kg ! · mv A , (23) where v is the impact velocity of the projectile and m and A = πr2 are the mass and cross-sectional area of the projec- tile, respectively. We found the stopping time T s to be indepen- dent of the velocity and only depending on the projectile size (3.0 ± 0.1 ms for 1 mm projectiles and 6.2 ± 0.1 ms for 3 mm projectiles). The compaction of dust underneath the impacted glass bead was measured by x-ray micro-tomography. The glass bead diam- eter and velocity for these experiments correspond to the com- paction calibration setup described in Table 1. The dust sample with embedded glass bead was positioned onto a rotatable sam- ple carrier between an x-ray source and the detector. During the rotation around 360◦, 400 transmission images were taken, from which we computed a 3D density reconstruction with a spatial resolution of 21 µm. The according results of the density recon- struction can be found in Guttler et al. (2009b), where we found that roughly one sphere volume under the glass bead is com- pressed to a volume filling factor of φ ≈ 0.23, while the sur- rounding volume is nearly unaffected with an original volume filling factor of φ ≈ 0.15. In this work, we will focus on the vertical density profile through the centre of the sphere and the compressed material (see section 4). 3.3. Furtherbenchmarkexperiments In this section, we will present two further experiments which will be used for the validation of the SPH code in sections 5.2 and 5.3. Heisselmann et al. (2007) performed low-velocity col- lisions (v = 0.4 m s−1) between cubic-shaped, approx. 5 mm- sized aggregates of the material as described in Sect. 3.1 and found bouncing whereby approx. 95% of the energy was dis- sipated in a central collision. Detailed investigation of the com- paction in these collisions (Weidling et al. 2009) revealed signif- icant compaction of the aggregates (from φ = 0.15 to φ = 0.37) after approx. 1,000 collisions. The energy needed for this com- paction is consistent with the energy dissipation as measured by Heisselmann et al. (2007). Fig. 2. Cumulative mass distribution of the fragments after a dis- ruptive collision, which can be described by a power law. The divergence for low masses is due to depletion of small aggre- gates because of the camera resolution. A further experiment deals with the disruptive fragmenta- tion of dust aggregates (for details see Guttler et al. 2009a,b). In this case, a dust aggregate with a diameter of 0.57 mm consist- ing of 1.5 µm spherical SiO2 dust with a volume filling factor of φ = 0.35 collides with a solid glass target at a velocity of 8.4 m s−1 (also see Fig. 20 in Guttler et al. 2009b). The projectile fragments and the projected sizes of these fragments are mea- sured with a high-speed camera with a resolution of 16 µm per pixel. As the mass measurement is restricted to the 2D images, the projected area of each fragment is averaged over a sequence of images where it is clearly separated from others. From this projected area, the fragment masses are calculated with the as- sumptions of a spheric shape and an unchanged volume filling factor. Fig. 2 shows the mass distribution in a cumulative plot. For the larger masses, which are not depleted due to the finite camera resolution, we find good agreement with a power-law distribution µ!α n(m) dm = m dm , (24) where m is the normalised mass (fragment mass divided by pro- jectile mass) and µ = 0.22 is a measure for the strength of frag- mentation, being defined as the mass of the largest fragment divided by the mass of the projectile. The exponent α has a value of 0.67. A similar distribution was already described by Blum & Munch (1993) for aggregate-aggregate fragmentation of ZrSiO4 aggregates with a comparable porosity. 4. Numerical Issues Before we perform the calibration process, some numerical is- sues have to be resolved. For instance, it is unfeasible to sim- ulate the dust sample, into which the glass bead is dropping in the compaction calibration experiment presented in section 3, as a whole. It is also unfeasible to carry out all necessary compu- tations in 3D. Therefore, we will only simulate part of the dust sample and explore at which size spurious boundary effects will emerge. Most of the calibration process has been conducted in 2D, but the differences between 2D and 3D results will be dis- cussed and quantified. R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 7 sphere particles dust particles boundary particles v0 rsample vertical density profile Fig. 3. Compaction calibration setup in 2D or, respectively, cross-section of 3D compaction calibration setup. A glass sphere impacts into a dust sample (radius rsample) with initial velocity v0. The dust sample is surrounded by boundary particles at its bot- tom. Their acceleration is set to zero at every time step. The ver- tical density profile at maximum intrusion serves as calibration feature. In this context we make use of 2D simulations in cartesian coordinates although the symmetry of the problem suggests the usage of cylindrical coordinates. However, the SPH scheme in cylindrical or polar coordinates battles with the problem of a sin- gularity at the origin of the kernel function. There exist only few attempts to solve this issue (e.g. Omang et al. 2006), but they are still under development and require high implementation efforts. Since in our case 2D simulations only serve as indicator for cal- ibration and 3D simulations are aimed at, we stick to cartesian coordinates. The glass bead is simulated with the Murnaghan equation of state p(ρ) = (cid:18) K0 ρ0!n n (cid:19)" ρ − 1# , (25) following the usual laws of continuum mechanics as presented in section 2.2. The compaction calibration setup is initialised with the numerical parameters shown in Table 1, unless stated otherwise in the text. Our tests showed that the maximum in- trusion depth and the density profile are the calibration parame- ters which are most sensitive to changes of the numerical setup. Density profiles (e.g. Figs. 4 and 5) display the filling factor φ along a line through the centre of the sphere and perpendicular to the bottom of the dust sample (see Fig. 3). The origin in the diagrams represents the surface of the unprocessed dust sample. The glass sphere has been removed in these figures. 4.1. Computationaldomainandboundaryconditions In 2D simulations we tested the effect of changing size and shape of the dust sample. Initially the particles were set on a triangular lattice with a lattice constant of 25 µm. To be geometrically con- sistent with the cylindric experimental setup, firstly, we utilised a box (width 8 mm) and varied its depth: 1.375 mm, 2.2 mm, 3.3 mm, and 5.5 mm. This is equivalent to 2.5 ×, 4 ×, 6 ×, and 10 × rsphere. Comparing the density profiles (Fig. 4, top), two features are remarkable: (1) The maximum filling factor at the top of the dust sample (φ ≈ 0.27 at D ≈ −0.6 mm) and the in- trusion depth D is nearly the same for all dust sample sizes. (2) Physical Quantity Symbol Value Unit Glass bead Bulk density(∗) Bulk modulus(∗) Murnaghan exponent(∗) Radius Impact velocity Dust sample Initial density Bulk density Reference density Initial filling factor Bulk modulus ODC mean pressure ODC max. filling factor ODC min. filling factor ODC slope ρ0 K0 n r v0 ρi ρs ρ′ 0 φi K0 pm φ2 φ1 ∆ 2540 5 × 109 kg m−3 Pa - 0.55 × 10−3 m 4 0.65 300 2000 300 0.15 3 × 105 1300 0.58 0.12 0.58 m s−1 kg m−3 kg m−3 kg m−3 - Pa Pa - - - Table 1. Numerical parameters for the compaction calibration setup. ODC stands for omni-directional compression relation (Eq. 21). Quantities marked by (*) represent the parameters for sandstone in Melosh (1989) which we adopt for glass here. For dsample < 3.3 mm we find spurious density peaks at the lower boundaries (D ≈ −1.4 mm and D ≈ −2.2 mm). In order to reduce the computation time we simulated the dust sample as a semicircle with the same radius variation as above. The resulting density profiles are shown in Fig. 4 (bot- tom). In contrast to the corresponding simulations with the box- shaped samples we find for rsample ≤ 1.375 mm an increased maximum filling factor and a slightly reduced intrusion depth. Due to the greater amount of volume lateral to the intrusion channel, material can be pushed aside more easily than inside the narrow boundaries of the semicircle. Therefore, a higher frac- tion of the material is compressed to higher filling factors. For rsample > 3.3 mm the spurious boundary effects become negli- gible within the compaction calibration setup and the density structure shows no significant difference for box-shaped and semicircle shaped dust samples. Hence, all computations of section 4 are conducted on the basis of a semicircle in 2D or a hemisphere in 3D with a radius of rsphere = 3.3 mm. In all cases the dust sample is bordered by a few layers of boundary particles. The acceleration of these particles is set to zero at each integration step, simulating reflecting boundary con- ditions. Apart from that, i.e. in terms of equation of state, they are treated like dust particles. We also tested damping boundary con- ditions by simulating two layers of boundaries. The outer layer was treated as described above, the inner (sufficiently large) layer was simulated with a high artificial α-viscosity. Since there was no significant difference in the outcome we fix all bound- aries in the afore mentioned way and treat them as reflecting. 4.2. ResolutionandConvergence Jutzi (2008) found in his studies of a basalt sphere impacting into a porous target that the outcome of his simulations strongly depends on the resolution. With a calibration setup similar to the one used in this paper Geretshauser (2006) confirms that also in simulations of the type presented in this work a strong res- olution dependence is present. He has found that the intrusion depth of the glass bead can be doubled by doubling the resolu- 8 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates φ r o t c a f g n i l l i f φ r o t c a f g n i l l i f 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 -5 -5 dsample = 1.375 mm dsample = 2.2 mm dsample = 3.3 mm dsample = 5.5 mm -2 -3 -4 sample height [mm] rsample = 1.1 mm rsample = 1.375 mm rsample = 2.2 mm rsample = 3.3 mm rsample = 5.5 mm -3 -4 sample height [mm] -2 φ r o t c a f g n i l l i f φ r o t c a f g n i l l i f 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 -2.5 -2.5 -1 0 -1 0 lc = 100 µm lc = 50 µm lc = 25 µm lc = 12.5 µm -1 -1.5 -2 intrusion depth [mm] lc = 100 µm lc = 50 µm lc = 25 µm -1.5 -2 sample height [mm] -1 -0.5 0 -0.5 0 Fig. 4. Vertical density profile at maximum intrusion for the compaction calibration setup and different shapes of the 2D dust sample (box and semicircle). Depth dsample of an 8 mm wide box (top) and radius rsample of the semicircle (bottom) were varied. In both cases spurious boundary effects appear for dsample < 3.3 mm and rsample < 3.3 mm, respectively. tion. Since the calibration experiments presented in Guttler et al. (2009b) are extremely sensitive even to minor changes of the setup, the convergence properties of porosity model and underly- ing SPH method will be investigated carefully in this paragraph. Additionally, we will study the differences in the outcome of 2D and 3D setup. For the 2D convergence study particles were initially put on a triangular lattice again. The lattice constants lc were 100, 50, 25, and 12.5 µm for the compaction calibration setup. The smooth- ing length h was kept constant relative to lc at a ratio of 5.6 × lc. The maximum number of interaction partners was Imax ≈ 180, the average Iav ≈ 100 and the minimum Imin ≈ 30. In the 3D convergence study we are using a cubic lattice with edge lengths lc = 100, 50, and 25 µm. The latter was simulated with 3.7 million SPH particles, which represent the limit of our computational resources. We fixed h = 3.75 × lc which yielded Imax ≈ 370, Iav ≈ 240, and Imin ≈ 70. The results are presented in Fig. 5. In contrast to the plots in Fig. 4 the glass sphere is not removed here. Coming from the right side of the plot, the filling factor rapidly decreases from a high value outside the scope of the plot indicating the sphere. The filling factor reaches its min- imum at an artificial gap between sphere and surface of the dust sample. The width of this gap is about one smoothing length h. The existence of the gap has two reasons: (1) Sphere mate- Fig. 5. Convergence study of the vertical density profile for 2D (top) and 3D (bottom) compaction calibration setups. The filling factor increase towards the surface of the dust sample accounts for the glass bead which is not removed in this plot. Simulations were performed for different spatial resolutions. All curves show a characteristic filling factor minimum between sphere and dust sample and a characteristic filling factor maximum indicating the dust sample surface. rial and dust material have to be separated by artificial viscosity due to stability reasons. This will be discussed below. (2) The volume of the sphere represents an area of extremely high den- sity and pressure with respect to the dust sample. This area is smoothed out due to the SPH method. The width of the smooth- ing is given by the smoothing length. Although a clear conver- gence behaviour is visible in Fig. 5 for both the 2D and the 3D case, a more unique convergence criterion has to be found. For this purpose we choose the maximum intrusion depth which proved to be very sensitive to resolution changes. The shape of the filling factor profile offers two choices to determine the intru- sion depth: (1) The filling factor minimum which represents the middle between sphere and dust sample and (2) the filling fac- tor maximum (peak) of the dust material left to the gap between sphere and dust sample. Fig. 6 shows the results for both cases in 2D (top) and 3D (bottom). The error bars around the minimum values represent the smoothing length and give an indication of the maximum er- ror. The position of the density peak remains almost constant, converging to ≈ −0.9 mm (2D) and ≈ −0.65 mm (3D), respec- tively, for higher resolutions. The position of the density min- imum at low resolutions significantly differs from the position R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 9 ] m m [ n o i s u r t n i . x a m ] m m [ n o i s u r t n i . x a m 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 0 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 0 2D density min 2D density peak 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 -1 3D density min 3D density peak 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 resolution [µm-1] Fig. 6. Convergence study of the maximum intrusion depth for the 2D (top) and 3D (bottom) compaction calibration setups. Filled symbols represent the position of the filling factor peak of the dust material, whereas empty symbols denote the position of the minimum filling factor at the gap between glass bead and dust material. The values are derived from the density profiles in Fig. 5. The smoothing length is indicated by the error bars. While the peak position remains almost constant at ≈ −0.9 mm (2D) and ≈ −0.65 mm (3D) with increasing spatial resolution, the position of the filling factor minimum quickly converges to the same value. This is due to the artificial separation of dust and glass materials, which is in the order of a smoothing length. of the density peak, but converges quickly to the same intrusion depth with higher resolutions. However, the differences between the extrema remain well within one smoothing length. This is due to the fact that sphere and dust sample are separated by about one smoothing length. Comparing 2D and 3D convergence the 3D case seems to converge more quickly. Due to the findings of this study we choose a spatial resolu- tion of lc = 25 µm for further simulations in 2D. In the 3D case lc = 50 µm is sufficient, but lc ≤ 50 µm is desirable if it is feasi- ble. After defining suitable values for the spatial resolution we now turn to the numerical resolution, which for the SPH scheme is given by the number of interaction partners of each single par- ticle. For the investigation of this feature we performed a study utilising the 2D compaction calibration setup with a spatial res- olution of lc = 25 µm and varied the ratio between smoothing length and lattice constant h/lc from 2 to 7 in steps of one. h/lc determines the initial number of interaction partners that is smoothed over. The resulting maximum, average, and minimum interactions Imax, Iav, and Imin and the corresponding smoothing φ r o t c a f g n i l l i f 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 h = 0.05 mm h = 0.075 mm h = 0.1 mm h = 0.125 mm h = 0.15 mm h = 0.175 mm -2.5 -2 -1.5 -1 -0.5 sample height [mm] Fig. 7. Convergence study for the density profile using the 2D setup and varying the smoothing length h. Through this variation the number of interaction partners is varied according to Table 2. The glass bead has been removed in this plot. For h ≤ 0.075 mm clear signs of instabilities are visible. For h ≥ 0.1 mm the filling factor has the same value and its position remains constant. The smoothing of the boundary of the dust sample is increased for increasing h. Table 2. Parameters for the convergence study regarding inter- action numbers h h/lc Imin Iav 0.050 mm 0.075 mm 0.100 mm 0.125 mm 0.150 mm 0.175 mm 2 3 4 5 6 7 3 10 16 23 32 43 13 30 53 82 116 158 Imax 25 55 92 142 205 274 Tcomp Nsteps 16.2 h 14.8 h 19.8 h 19.0 h 21.6 h 24.3 h 132401 96912 71175 56347 46782 39980 lengths h can be found in Table 2. Additionally, we measured the computation time Tcomp that the simulations took on 4 cores of a cluster with Intel Xenon Quad-Core processors (2.66 GHz) for a simulated time of 5 ms and the number of integration steps Nstep of our adaptive Runge-Kutta Cash-Karp integrator. Comparing the density profiles in Fig. 7 (where the glass bead has been removed) instabilities in the form of filling fac- tor fluctuations due to insufficient interaction numbers appear for smoothing lengths h ≤ 0.075 mm, i.e. for Iav ≤ 30. For h ≥ 0.1 mm the density profile maintains essentially the same shape: The position and height of the filling factor peak remains nearly the same and φ smoothly drops to ≈ 0.18 towards the bot- tom of the dust sample. Only the sharp edge at the top of the dust sample is smoothed out over a wider range due to the increased smoothing length. Table 2 shows that the number of integration steps Nstep is de- creasing with increasing interaction numbers. This is due to the fact that elastic waves inside the dust sample are smoothed out over a wider range causing the adaptive integrator to increase the duration of a time step, since density fluctuations do not have to be resolved as sharply as at smaller smoothing. As expected, the computation time Tcomp is generally increasing with increasing number of interactions. There are two exceptions: h = 0.075 mm and h = 0.125 mm. Here, the decrease of Nsteps overcompen- sates the increase of interactions leading to a decrease of Tcomp. Hence, a ratio h/lc ≈ 5 yields the necessary accuracy and an ac- 10 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates ceptable amount of computation time. This study also justifies the choice of h/lc = 5.6 in Guttler et al. (2009b) and we will stick to this ratio throughout this paper. According to these findings, for 3D simulations theoretically an average interaction number of I3/2 av ≈ 750 would be needed to achieve the same numerical resolution. However, alike sim- ulations are unfeasible and our choice of Iav ≈ 240 in 3D is equivalent to Iav ≈ 40 in 2D which should provide sufficient and reliable accuracy. 4.3. Geometricaldifference–2Dand3Dsetups As one can easily see in Fig. 6 2D and 3D simulations have sig- nificantly different convergence values for the intrusion depth. This deviation is caused by the geometrical difference of the 2D and 3D setup. The 2D setup (glass circle impacts into dust semi- circle) represents a slice through a glass cylinder and a semi- cylindrical dust sample, which implies an infinite expansion into the third spatial direction. Whereas the 3D setup represents a real sphere dropping into a "bowl" of dust. The relation for the intru- sion depth found by Guttler et al. (2009b) (see Sect. 3) contains a geometrical dependence: D ∝ mv/A, where D is the intrusion depth, m the mass of the impacting glass bead, v its impact ve- locity, and A its cross-section. m2D is a mass per unit length. Guttler et al. (2009b) already exploited this relation in order to determine a rough correction factor between 2D and 3D simual- tional setups: m3Dv A3D = 4 3 πr3ρ · v πr2 = 8 3π πr2ρ · v 2r = 8 3π m2Dv A2D . (26) Hence, the 2D intrusion depth has to be corrected by a factor of ≈ 8 3π in order to determine the 3D intrusion depth. The compar- ison is shown in Fig. 8. Again we choose the 2D and 3D data gained in the convergence study for the peak filling factor values shown in Fig. 6. Fig. 8 shows the original 2D data, the corrected 2D data, and the according 3D data (with error bars displaying the smoothing length). The 3D values nicely follow the rough correction and remain well within the maximum error. This comparison justifies the correction of the results in Guttler et al. (2009b). With the aid of this – now verified – correction factor the 2D data of the intrusion depth can be converted into 3D data and all calibration tests involving the intrusion depth can be car- ried out in 2D, which saves a significant amount of computation time. Comparing the vertical density profiles of 2D and 3D se- tups in Fig. 5 resolves also another issue that remained unsolved in Guttler et al. (2009b). According to the experimental data the filling factor drops to a value of φ ≈ 0.16 within ≈ 0.6 mm away from the bottom point of the glass bead. For high-resolution 2D simulations (Fig. 5, top) the filling factor does not drop to this value within the whole dust sample. However, the 3D simula- tions (Fig. 5, bottom) show that this effect again is due to the difference of 2D and 3D geometry. With the 3D setup the fill- ing factor drops to φ ≈ 0.16 within ≈ 0.9 mm. All deviations from experimental findings due to this effect, in particular the presence of a large amount of volume with φ ≤ 0.2 in the cumu- lated volume over filling factor diagram (Fig. 15 in Guttler et al. 2009b), can in principal be removed by switching to 3D simula- tions. They do not represent a fundamental error in the porosity model. ] m m [ n o i s u r t n i . x a m 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 0 2D peak corr 3D peak 2D density peak 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 resolution [µm-1] Fig. 8. Verification of the 2D-3D correction factor. Filled sym- bols denote the position of the filling factor peak of the dust material in Fig. 5. Triangles represent 3D and squares 2D val- ues. The conversion from 2D to 3D intrusion depth utilising the correction factor from Eq. 26 and Eq. 23 due to the geometrical difference is indicated by the line without symbols. The 3D val- ues are in very good agreement with the very rough theoretical prediction. They lie well within the errors. 4.4. ArtificialViscosity Since artificial viscosity plays an eminent role for the stability of SPH simulations we investigate its influence on the outcome of our compaction calibration setup. Only artificial α-viscosity is applied in our test case, but in a threefold way: (1) It damps high oscillation modes of the glass bead caused by the stiff Murnaghan equation of state (EOS). Thereby it enlarges the time step of our adaptive integrator and saves computation time. (2) It is used to provide the dust material with a basic stability. (3) It separates the areas of Murnaghan EOS and dust EOS and pre- vents a so called "cannonball instability". For all three cases also the influence of β-artificial viscosity has been tested, but its in- fluence on all benchmark parameters was negligible. (1) The choice of the α-viscosity of the glass bead proved to be very uncritical. We choose the canonical value α = 1.0. There was no influence on the physical benchmark parameters for all α values, except α = 0 which produces an instability. Values for α > 1.0 show no significant effect on the damping and the influence of 0.1 < α < 1.0 on it is not too high, but still observable. Hence, we stick to the canonical value. (2) Sirono (2004) applies no artificial viscosity to his porous ice material because of its spurious dissipative properties. Our findings, shown in Fig. 9, confirm this assessment regarding the dust material. Within our 2D compaction calibration setup (lc = 25 µm, h/lc = 5.6) we vary α from 0 to 2 and observe its influence on the density profile (Fig. 9, top) and the maximum intrusion represented by the filling factor peak of the dust mate- rial (Fig. 9, bottom). The position of the filling factor peak ranges from ≈ −0.92 mm with α = 0.0 to ≈ −0.62 mm at α = 2.0. This clearly demonstrates the dissipative feature of the α-viscosity, since a smaller amount of kinetic energy of the glass bead is transformed into plastic deformation with higher α. The residual energy must have been dissipated. However, the α-viscosity-intrusion curve seems to saturate at a value of ≈ −0.6 mm. The decrease of the maximum intrusion can also be seen in the density profile (Fig. 9, top). While the profile maintains nearly the same shape, R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 11 the height of the filling factor peak is decreasing with increas- ing α. Hence, an increasing artificial viscosity diminishes the peak pressure during compaction, which is via the compressive strength relation Σ(φ) directly responsible for the height of the filling factor peak. In contrast to Sirono (2004) we find it necessary to apply a small amount of α-viscosity to the dust material. For α < 0.1 the results show traces of an instability, which is also responsible for a rapid increase of the maximum intrusion. Therefore, we find it convenient to apply an artificial viscosity with α = 0.1 to the dust material, which holds for the previous simulations of this section as well as the following. The choice of a non-zero α, however, is also justified by experimental findings: After impacting into the dust sample the glass bead shortly oscillates due to the elas- tic properties of the dust. This oscillation is damped by internal friction, which we model with artificial viscosity. Therefore, by choosing a non-zero α we take into account the dissipative prop- erties that our dust material naturally has. A quantitative calibra- tion of this parameter, however, has to be left to future work. (3) During our first simulations with the 2D compaction cali- bration setup we observed what is sometimes described in the lit- erature as "cannonball instability": During the compaction pro- cess, when the glass bead intrudes into the dust material, single particles at the sphere's surface start to oscillate between the do- mains of the Murnaghan EOS and dust EOS. Due to the severe discrepancy of the "stiffness" of these two equations of state the particles collect a huge amount of kinetic energy until they are fast enough to generate a pressure on the dust material that ex- ceeds the compressive strength Σ(φ). Eventually they disengage from the sphere's surface like a cannonball and dig themselves into the dust sample causing a huge amount of unphysical com- paction. We tackle this problem by applying to all SPH particles with dust EOS, which interact with glass bead SPH particles, the same amount of α-viscosity as to the sphere, i.e. α = 1.0. In our simulations this is sufficient to prevent the "cannonball instability". The spurious dissipation caused by this measure is negligible. 5. Calibration 5.1. CompressiveStrength-CompactionProperties In this section we will refine and extend a study of the com- paction properties of the dust sample, which was already carried out in a similar, but less detailed way in Guttler et al. (2009b). There it turned out that the quantity having the most influence on the compaction was the compressive strength relation Σ(φ) (Eq. 21), which was measured in an omni-directional and static manner. In order to adopt this relation for dynamic compression the mean pressure pm and the "slope" of the Fermi-shaped curve ∆ can be treated as free parameters. The upper and lower bound- aries of the filling factor φ2 and φ1, respectively, remain constant even in the dynamic case. Guttler et al. (2009b) found that by lowering pm most of the features of the compaction calibration setup can be reproduced in a very satisfactory manner. Hereby the filling factor over compressive strength curve (see also Fig. 2 in Guttler et al. 2009b) is shifted towards lower pressures and the yield pressure for compression is lowered. Using only 2D simulations and a rough parameter grid Guttler et al. (2009b) fix pm = 1.3 kPa. The "slope" ∆ has not been considered. In this work we will consider ∆ and we will perform more accurate pa- rameter studies for pm. From the latter we will predict a rea- sonable choice for pm, which will represent the basis for a 3D simulation of the compaction calibration setup. The results of φ r o t c a f g n i l l i f ] m m [ n o i s u r t n i . x a m 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 α = 0.0 α = 0.05 α = 0.1 α = 0.5 α = 1.0 α = 2.0 -4 -3.5 -2.5 -3 -1 sample height [mm] -2 -1.5 -0.5 0 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 0 Max. intrusion (peak) 0.5 1 α viscosity 1.5 2 Fig. 9. Density profile (top) and maximum intrusion (bottom) for different values of artificial α-viscosity. The shape of the density profile hardly changes, but increasing α-viscosity decreases the maximum filling factor and the maximum intrusion depth. this simulation will be compared to results from the laboratory. For the comparison we use the same features as Guttler et al. (2009b). 5.1.1. Fixing free parameters Since there are no empirical data available for pm in the dynami- cal compressive strength curve we perform a parameter study in order to determine a suitable choice for this important quantity. For this study we make use of the 2D compaction calibration setup and vary pm from 0.13 to 13.0 kPa (Fig. 10 and 11), where 13.0 kPa represents the value for the static compressive strength curve. The effect of lowering pm can most clearly be seen in the vertical filling factor profile (Fig. 10, top). First of all, more material can be compressed and it is compressed to higher filling factors. As a consequence the glass bead intrudes deeper into the dust sample. From experimental results we expect an intrusion depth of about one sphere diameter (≈ 1 mm). With the aid of the empirical relation between momentum over impactor cross- section mvA−1 and intrusion depth D (Eq. 23) as well as the cor- rection factor between 2D and 3D intrusion depth (Eq. 26) we estimate that D3D ≈ 1 mm corresponds to D2D ≈ 1.42 mm. Fig. 11 shows the maximum intrusion over the stopping time for var- ious values of pm (labels). The estimated D2D is indicated by a 12 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates pm = 0.13 kPa pm = 0.26 kPa pm = 0.65 kPa pm = 1.30 kPa pm = 2.60 kPa pm = 6.50 kPa pm = 13.0 kPa -1.5 -0.5 sample height [mm] -1 0 0.5 pm variation φ r o t c a f g n i l l i f 0.6 0.5 0.4 0.3 0.2 0.1 0 0.4 0.35 0.3 0.25 φ r o t c a f g n i l l i f k a e p -2.5 -2 0.2 0.1 1 10 mean pressure pm [kPa] Fig. 10. Density profile (top) and maximum peak filling factor from the density profile over pm (bottom) for different values of pm, which represents the mean pressure in the compressive strength relation Σ(φ) (Eq. 21). Lowering pm from 13.0 kPa (static compressive strength) increases the compressed volume, its filling factor, and the maximum intrusion depth of the glass bead. The parameter study has been performed using the 2D compaction calibration setup. dashed line. We deduce that regarding intrusion depth a dynamic mean pressure pm = 0.26 kPa is a suitable choice. This is supported by the peak filling factor appearing in the filling factor profile (Fig. 10, bottom). Empirical data indicate that in case of the compaction calibration setup a peak filling factor φmax ≈ 0.3 can be expected. The comparison between 2D and 3D results (Sect. 4.3) has shown that the peak filling factor in the vertical density profile in the 2D case is generally higher than for the same situation in 3D. The equivalent of φ3D max ≈ 0.3 is a maximum filling factor of φ2D max ≈ 0.34 in 2D. This points to a choice of pm ≈ 0.3 kPa, which is consistent with the findings from the intrusion depth. The compressive strength relation Σ(φ) (Eq. 21) contains a second free parameter ∆, which accounts for the "slope" of the Fermi-shaped curve. In Guttler et al. (2009b) this parameter was chosen to be the same as the one of the static omni-directional compressive strength curve and a closer investigation was not carried out. In order to understand its effect on the compaction properties of the dust sample we are utilising the 2D compaction calibration setup again and vary ∆ from 0.55 to 0.80. The results are presented in Fig. 12. From the vertical density profile (Fig. ] m m [ n o i s u r t n i . x a m 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 Expected 2D depth pm variation 0.13 0.39 0.26 0.52 0.65 0.91 1.17 0.78 1.04 1.30 2.60 13.0 6.50 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 stopping time [ms] Fig. 11. Maximum intrusion over stopping time for different val- ues of the mean pressure pm (labels, in kPa) in the compressive strength relation Σ(φ) (Eq. 21) using the 2D compaction calibra- tion setup. The dashed line indicates the 2D intrusion depth that is equivalent to a 3D intrusion depth of ≈ 1 mm according to Eq. 23 and the 2D-3D correction factor from Eq. 26. This sup- ports the choice of the mean pressure pm = 0.26 kPa for further simulations. 12, top) it can be seen that increasing ∆ increases the intrusion depth, but not as effectively as by lowering the mean pressure pm. This is due to the fact that increasing ∆ is hardly increasing the peak filling factor in the vertical density profile. More vol- ume is compacted to lower filling factors. Whereas by lowering pm the total amount of volume that is compacted is smaller, but it is compacted to higher filling factors. This behaviour can be seen comparing the density profiles for the pm variation (Fig. 10, top) and ∆ variation (Fig. 12, top) and particularly from the cu- mulated volume over filling factor curve (Fig. 12, bottom). This figure shows the normalised fraction of compacted volume cor- responding to a volume filling factor > φ. The intersection of the curves with the y-axis represents the total compressed volume, which is increased from ≈ 7 to ≈ 9.5 sphere volumes. According experimental measurements here expect a value of roughly one sphere volume (see section 3.2). Especially for 0.18 < φ < 0.23 the compacted volume fraction is increased. The equivalent fig- ure for the pm variation can be found in Guttler et al. (2009b, Fig. 15). From the comparison of 2D and 3D calibration setups (section 4.3) we know that a huge amount of this compaction (especially in the lower filling factor regime) is due to the ge- ometrical difference. However, the experimental data do not in- dicate a particularly high amount of compaction to lower filling factors (rather the contrary) and, therefore, we maintain our ini- tial choice of ∆ = 0.58. 5.1.2. Comparison with experiments For the comparison with experiments we performed a simulation using the 3D compaction calibration setup (see Table 1) with two exceptions: (1) The bulk modulus of the dust material was set to K0 = 2 kPa (instead of K0 = 300 kPa) since findings pre- sented below indicated a much lower bulk modulus. However, the choice of K0 has little influence on the compaction proper- ties calibrated for in the compaction calibration setup. It is more important for bouncing and fragmentation, which will be shown below. (2) The mean pressure of the compressive strength re- R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 13 φ r o t c a f g n i l l i f 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 -2.5 / e r e h p s V V e m u l o v d e t a l u m u C 10 9 8 7 6 5 4 3 2 1 0 0.16 0.18 ∆ = 0.55 ∆ = 0.60 ∆ = 0.65 ∆ = 0.70 ∆ = 0.75 ∆ = 0.80 -1.5 -2 sample height [mm] -1 -0.5 0 ∆ = 0.55 ∆ = 0.60 ∆ = 0.65 ∆ = 0.70 ∆ = 0.75 ∆ = 0.80 0.2 0.22 0.24 filling factor φ 0.26 0.28 0.3 Fig. 12. Density profile (top) and normalised volume fraction of compacted volume over its corresponding filling factor (bottom) for different values of ∆ in the compressive strength relation (Eq. 21). Increasing ∆ increases the amount of compacted volume at filling factors 0.18 < φ < 0.23 and thereby the maximum intrusion depth. Whereas the peak filling factor in the density profile is not changed. lation (Eq. 21) was set to pm = 0.26 kPa as suggested in the previous section. The following features of the compaction calibration setup were measured in the laboratory and will be used here for com- parison: (1) the stopping time T s, (2) the intrusion curve of the projectile, (3) the vertical density profile, (4) the cumulated vol- ume over filling factor curve, and (5) the cross-section through sphere and dust sample displaying the filling factor. (1) The experiments show that the stopping time T s of the glass bead, i.e. the time from first contact with the dust sample until deepest intrusion, is nearly constant at T exp s = 3.0 ± 0.1 ms for 1 mm projectiles over different impact velocities (see section 3.1). In our simulation we find T sim s = 2.42 ± 0.05 ms, which is not in excellent agreement but also not too far off the experimen- tal results. (2) The intrusion curve h(t) was cleared from gravity effects and normalised through h′(t′) = h(t)/D and t′ = t/TS , where h(t) is the position of the bottom of the glass bead as a function of time, D the maximum intrusion depth, and T s the stopping time. At first contact t is h′(t′ = 0) = 0 and at deepest intrusion h′(t′ = 1) = −1 (see also Guttler et al. 2009b, section 3.2.2). The comparison is shown in Fig. 13: The intrusion curve generated by our simulation lies well within the data from the experiments with 1 mm and 3 mm spheres and only slightly below the sine curve fitted to the empirical data. (3) The comparison between the experimentally measured vertical density profile and the result from our simulation is shown in Fig. 14. The crosses represent the data from two exper- iments where the sphere has not been removed for the measure- ment. For this reason the filling factor reaches extremely high values at ≈ −1 mm indicating the bottom of the glass bead. The vertical density profile of our simulation is given by the solid line and the vertical dashed line is placed at its filling factor peak at −1.02 mm representing the maximum intrusion depth. Comparing with the experimentally measured maximum intru- sion depth of −1.07 mm this is an excellent result. Since a depth of ≈ −1 mm was aimed for using Eqns. 23 and 26 and the 2D intrusion depth study of the previous section this result also sup- ports the validity of these relations. In addition to the exact value of the intrusion depth our simulation also reproduces the shape of the given experimental vertical density profile very well. Only the step-like structure at ≈ −1.5 mm does not find an equivalent in our simulation but it is nicely interpolated. (4) The vertical density profile shows only a cut through the compressed volume. It contains information about the exact structure of the compression. The cumulated volume over fill- ing factor curve (Fig. 15) has the advantage that it represents the total compressed volume together with its filling factors, hence, these features are not fully independent but focus on different as- pects of the compression. The cumulated volume is normalised through the sphere volume. It accounts for the volume fraction of compacted area corresponding to a volume filling factor > φ. In general, the experimental reference and our simulation show a very good agreement. Slightly too much volume is compressed to high filling factors in the simulation which leads to an almost constant deviation for φ < 0.26. However, the slope is repro- duced very well. (5) The comparison between the cross-sections through sphere and dust sample along the z-axis (Fig. 16) reveals reasons for the excess of compressed volume. First, the cross-section of the sphere is artificially enhanced by the smoothing of its bound- aries, which is inherent of the SPH method. One effect of the smoothing is the existence of a gap between sphere and dust sample, which was already discussed in Sect. 4.2 and is also clearly visible in Fig. 16 (top). Hence, we assume that the dust sample actually begins, where it has its maximum compression. The fact that the sphere pokes out of the dust sample a bit more than in the experiment has its reason also in the artificial en- largement of the cross-section. Second, it can be seen that in the experimental reference (Fig. 16, bottom) the compacted region is much narrower and more concentrated underneath the sphere. In the simulated result the compacted region is a bit broader. This indicates that the shear strength seems to be a bit smaller than we assume. Third, the compaction reaches too high filling factors, which was already visible in the cumulated volume di- agram. Nevertheless, both cross-sections match very well, es- pecially with respect to the mediocre resolution. Remarkably, even the slight intrusion channel on the left and right side of the sphere, which features a slight compression, can be reproduced. 5.2. BulkModulus As it was mentioned in section 3.1 there are two estimates for the bulk modulus K0 for the uncompressed material with φ ≈ 0.15. K0 in our model is also the pre-factor for comput- ing the bulk modulus for higher filling factors with the aid of 14 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates dust particles boundary particles (glass) v0 gravity vf (a) (b) (c) (d) Fig. 17. Bouncing sequence for t = 0 ms (a), t = 10 ms (b), t = 18 ms (c), and t = 25 ms (d). The colour code indicates the filling factor. An aggregate consisting of dust particles (Sirono EOS, diameter 1.0 mm) hits a solid surface simulated by boundary glass particles (Murnaghan EOS, diameter 1.6 mm, thickness 0.1 mm) with a velocity of v0 = 0.2 m s−1. For this simulation a bulk modulus of K0 = 5.0 kPa and pm = 0.26 kPa were used. The aggregate hits the surface and starts to be compacted at its bottom (b). While the plastic deformation at the bottom increases, the aggregate is also deformed elastically: it gets broader (c). Eventually it leaves the surface with a final velocity v f (d). It features a permanent compaction, while the elastic deformation vanishes. (An animation of this figure is available in the online journal.) φ r o t c a f g n i l l i f 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 -3 Experiment 1 Experiment 2 pm = 0.26 kPa -2.5 -2 -1.5 -1 -0.5 0 sample height [mm] Fig. 13. Comparison between intrusion curves from drop exper- iments using 1 mm and 3 mm spheres and our 3D simulation (mean pressure pm = 0.26 kPa). The curves are normalised such that the origin represents first touch of sphere and dust sam- ple and (1,-1) denotes stopping time at maximum intrusion. The simulated curve lies slightly underneath the fit to the experimen- tal data, but well within the errors. The deviation is due to a smaller stopping time than in the experiments. Fig. 14. Comparison between experimentally measured (crosses, sphere not removed) and simulated (solid line, sphere removed) density profiles at maximum intrusion for the compaction cali- bration setup. The dashed line indicates the position of the sim- ulated maximum intrusion depth given by the density peak at −1.02 mm. The simulation was carried out in 3D using a mean pressure pm = 0.26 kPa for the compressive strength relation (Eq. 21). Both profiles are in excellent agreement. The fact that the step-like structure of the experimental data cannot be seen in the simulation is a minor drawback since it is interpolated nicely. a power law (Eq. 9). Since these indirect findings, 1 kPa (see Weidling et al. 2009) and 300 kPa (from sound speed measured by Blum & Wurm 2008; Paszun & Dominik 2008), differ by two orders of magnitude, we try to find a suitable value for K0 utilis- ing a calibration experiment. Simulating the low velocity collision setup by Weidling et al. (2009), a 3D dust sphere (φi = 0.15,1 mm diameter, 267737 par- ticles, lc = 12.5 µm) drops onto a solid surface (cylindrical, 1.6 mm diameter, 0.1 mm thick, 115677 particles, lc = 12.5 µm) with initial velocity v0 = 0.2 m s−1 (see Fig. 17). The material parameters are shown in Table 3. The bulk modulus K0 is var- ied with respect to two values of the mean pressure pm in the compressive strength relation (0.26 kPa and 1.3 kPa). During the impact a small region of the bottom of the dust sphere is com- pacted. Then the deformed sphere bounces off the target with reduced velocity v f (see Fig. 17). The latter effect was already observed in Guttler et al. (2009b) and demonstrates the ability of our code and the implemented porosity model to simulate the elastic properties of the dust correctly. In this study we have dou- bled the resolution and determine the coefficient of restitution εrest = v f v−1 0 depending on K0 (see Fig. 18). The bouncing calibration setup is equivalent with two cen- trally colliding dust aggregates at 0.4 m s−1, thus, our results are comparable to Heisselmann et al. (2007): εrest ≈ 0.2 and ≈ 95 % energy dissipation. Based on the results of the previous section, where pm = 0.26 kPa turned out to be a good choice for the mean pressure, our results of this experiment favour a bulk modulus K0 ≈ 5 kPa. This value is close to K0 = 1 kPa computed by Weidling et al. / e r e h p s V V e m u l o v d e t a l u m u C R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 15 Experiment 1 Experiment 2 pm = 0.26 kPa 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0.2 0.21 0.22 0.23 0.24 0.25 0.26 0.27 0.28 0.29 filling factor φ Fig. 15. Total cumulated volume over filling factor for drop ex- periments (crosses and triangles) and our 3D simulation with mean pressure pm = 0.26 kPa (solid line). The plot shows the cu- mulated mass fraction (normalised through the sphere volume) with a filling factor > φ over φ. Our simulation is in good agree- ment with the experimental findings. However, a higher amount of volume compressed to high filling factors leads to an almost constant deviation for φ < 0.26. The slope is reproduced very well. Physical Quantity Symbol Value Unit Glass plate Bulk density(∗) Bulk modulus(∗) Murnaghan exponent(∗) Radius Thickness Dust sample Initial density Bulk density Reference density Initial filling factor Bulk modulus ODC mean pressure ODC max. filling factor ODC min. filling factor ODC slope Impact velocity Radius ρ0 K0 n rplate dplate ρi ρs ρ′ 0 φi K0 pm φ2 φ1 ∆ v0 r 2540 5 × 109 4 0.8 × 10−3 0.1 × 10−3 various 260 and 1300 300 2000 300 0.15 0.58 0.12 0.58 0.2 0.5 × 10−3 kg m−3 Pa - m m kg m−3 kg m−3 kg m−3 - Pa Pa - - - m s−1 m Table 3. Numerical parameters for the bouncing calibration setup. ODC stands for omni-directional compression relation (Eq. 21). Quantities marked by (*) represent the parameters for sandstone in Melosh (1989), which we adopt for glass here. (2009) with a rough model. Our simulations yield a coefficient of restitution εrest = 0.19 (≈ 96 % energy dissipation) for K0 = 5.0 kPa, which is in excellent agreement with the experimental results. On the other hand for K0 = 500 kPa we find εrest = 0.09 (≈ 99 % energy dissipation), which is too far away from the reference. A high value for the bulk modulus K0 as it is given by the sound speed measurements is therefore excluded. Given the higher value pm = 1.3 kPa for the compres- sive strength curve εrest is raised for all choices of K0. For K0 = 1.0 kPa it yields εrest ≈ 0.7 and only ≈ 50 % of the Fig. 16. Cross section through glass bead (red) and dust sample (light blue) at maximum intrusion for our 3D simulation (with pm = 0.26 kPa, top) and the drop experiment (bottom). The colour indicates the spatially averaged filling factor. The den- sity structure underneath the glass bead match very well. Even the slight compression along the tight intrusion channel can be reproduced. In the simulated plot a gap between glass bead and the most dense area is clearly visible. This is due to the smooth- ing of the sphere and was discussed in Sect. 4.2. The gap has roughly the size of one smoothing length h. energy is dissipated. On the other hand for K0 = 300 kPa we find εrest ≈ 0.13 equivalent to ≈ 98% energy dissipation. For pm = 1.3 kPa it is harder to dissipate energy by plastic deforma- tion. Hence, less energy is dissipated in the compaction process and the coefficient of restitution is generally higher. This bouncing experiment fixes our choice for the bulk mod- ulus to K0 ≈ 5 kPa while maintaining pm = 0.26 kPa from the compaction experiment in the previous section. The next section will show that this choice is also consistent with the fragmenta- tion behaviour of the dust aggregates. 5.3. Fragmentation Since the intended field of application of our calibrated SPH code will be the simulation of pre-planetesimal collisions, it is of major importance to calibrate and test the fragmentation be- haviour of the simulated material. For this reason we simulate the fragmentation experiment described already in the second part of Sect. 3.3. 16 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates R C n o i t u t i t s e r f o t n e i c i f f e o c 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 pm = 0.26 kPa pm = 1.3 kPa Exp. value 1 10 100 1000 bulk modulus K0 [kPa] 100 10-1 10-2 10-3 n o i t c a r f s s a m e v i t a l u m u c 10-4 10-4 K0 = 3.00 kPa K0 = 3.50 kPa K0 = 4.00 kPa K0 = 4.50 kPa K0 = 5.50 kPa K0 = 6.50 kPa K0 = 4.50 kPa (Fit) 10-2 10-3 10-1 fragment mass / total mass 100 Fig. 18. Coefficient of restitution εrest over bulk modulus K0 for values of the mean pressure pm of the compressive strength re- lation. For the lower pm = 0.26 kPa (squares) more energy is dissipated by plastic deformation. For this reason the coefficient of restitution is significantly lowered compared to the higher pm = 1.3 kPa (crosses). Best agreement with the experimental reference of εrest = 0.2 (dashed line) is given for K0 ≈ 5 kPa and K0 ≈ 20 kPa, respectively. Fig. 19. Cumulative mass distribution of the fragments of a dust aggregate impacting on a glass plate for different values of K0. For low fragment masses the shape of all simulated curves dif- fers from the experimental curve in Fig. 2 due to the limited res- olution of the experimental setup. An increase of K0 leads to an increase in the slope α of the power-law fit. The best agreement with the experimentally measured slope 0.67 was found for the simulation with K0 = 4.5 kPa. In our simulation a dust aggregate (φi = 0.35, 189296 par- ticles) hits a glass plate (188478 particles) from below with an impact velocity of v0 = 8.4 m s−1. The spatial resolution (lc = 8.0 µm, h = 30.0 µm) was chosen such that a single parti- cle has less than 5 × 10−6 times the mass of the whole aggregate (6.8 × 10−8 kg) to resolve the same fragment masses as the ex- perimental reference. Gravitation was taken into account. A list of the other relevant parameters can be found in Table 4. We investigated the influence of the bulk modulus on the fragmentation behaviour. For this reason we varied K0 from 3.0 kPa to 6.5 kPa. For comparison with the reference experi- ment (Fig. 2) we also plot our fragmentation data in a cumu- lative way (Fig. 19) and find also a good a agreement with the power-law distribution of Eq. 24. The simulation was evaluated after 0.8 ms simulated time. The mass of a fragment is given by the sum of the mass of the SPH particles belonging to it. We consider two fragments as separated when they are not linked by SPH particles that interact with each other, i.e. when the clos- est SPH particles of two fragments have a distance more than a smoothing length. Guttler et al. (2009b) were facing the problem that using K0 = 300 kPa almost no fragmentation occurred (see Fig. 21 in this reference). Their speculations about a modification of the shear strength in order to resolve this problem proved to be wrong. The quantity with the most impact on the fragmen- tation behaviour of the dust aggregate is its bulk modulus. In general it can be said that an increase of the bulk modulus leads to an increasing slope α of the fragment distribution, whereas the size of the largest fragment (normalised through the total mass of the projectile) µ roughly remains constant at ≈ 20 % up to K0 = 6.0 kPa. For higher K0 mainly small chunks and single SPH particles will burst off the aggregate, which is only compacted but does not fragment. This reproduces the situation described in Guttler et al. (2009b) and Sect. 3. We calibrate for α, which is more sensitive to changes of K0 (see Table 5). Given the measured value of α = 0.67 we find an excellent agreement with our simulation using K0 = 4.5 kPa, which yields α = 0.673 ± 0.017. Moreover this simulation re- produces also the experimentally measured normalised mass of the largest fragment µ = 0.22 to a very high accuracy (µ = 0.234 ± 0.007). The slight overestimation may be caused by the fact that the increase in filling factor is not taken into account in the analysis of the experimental data (see Sect. 3), whereas in the simulation it is. The fragment distributions for different K0 and the best fit for the power-law are shown in Fig. 19. Setup and out- come of the simulation are displayed in Fig. 20. As already indi- cated in Sect. 5.2 the choice of pm = 0.26 kPa and K0 = 4.5 kPa is consistent with the results from the compaction and bouncing experiments. The fragmentation experiment serves as a proof for the validity of these choices and also for the consistency of the underlying porosity model. In contrast to the experiments we find no material sticking to the glass plate. This is not possible due to the setup of the simulation. As in Sect. 4.4 we use artificial viscosity to separate glass and dust materials. This leads to an additional pressure on the dust material which prevents sticking. 6. Discussion In this work we presented a smooth particle hydrodynamics (SPH) code equipped with extensions for continuum mechan- ics of solid bodies and an extended version of the Sirono (2004) porosity model. The code uses experimentally measured macro- scopic parameters such as tensile strength and a static compres- sive strength relation. In Guttler et al. (2009b) this code was used to determine a relation for the shear strength and an estimate for the mean pressure pm for the dynamic compressive strength re- lation (Eq. 21). The estimate was quite crude, though, due to the usage of 2D simulations only. This work profoundly investigated the numerical properties of the SPH code. Utilising the compaction calibration setup of Guttler et al. (2009b) as an example we determined an adequate size of the computational domain and adequate numerical and spatial resolutions. We have shown that the results for this setup R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 17 v0 boundary particles (glass) gravity dust particles Fig. 20. Fragmentation sequence for t = 0 ms (top left), t = 0.4 ms (top right), and t = 0.8 ms (bottom). The colour code indicates the filling factor. A projectile consisting of dust particles (Sirono EOS, diameter 0.57 mm) hits a solid surface simulated by boundary glass particles (Murnaghan EOS, diameter 1.6 mm, thickness 0.04 mm) with an impact velocity of v0 = 8.4 m s−1. A bulk modulus of K0 = 4.5 kPa was used for this simulation. While many small pieces, often single SPH particles, chip off, most of the aggregate is compacted to its maximum filling factor φmax = 0.58 and fractures. The bottom layer of the impacting sphere is hardly compacted. This uncompacted layer is still visible on the fragments. (An animation of this figure is available in the online journal.) Physical Quantity Symbol Value Unit K0 [kPa] Slope α Norm. largest fragment µ Glass plate Bulk density(∗) Bulk modulus(∗) Murnaghan exponent(∗) Radius Thickness Dust sample Initial density Bulk density Reference density Initial filling factor Bulk modulus ODC mean pressure ODC max. filling factor ODC min. filling factor ODC slope Impact velocity Radius ρ0 K0 n rplate dplate ρi ρs ρ′ 0 φi K0 pm φ2 φ1 ∆ v0 r various 300 2000 700 0.35 260 0.58 0.12 0.58 8.4 0.285 × 10−3 m Table 4. Numerical parameters for the fragmentation calibra- tion setup. ODC stands for omni-directional compression rela- tion (Eq. 21). Quantities marked by (*) represent the parameters for sandstone in Melosh (1989), which we adopt for glass here. 2540 5 × 109 4 0.8 × 10−3 0.04 × 10−3 kg m−3 Pa - m m kg m−3 kg m−3 kg m−3 - Pa Pa - - - m s−1 3.00 3.50 4.00 4.25 4.50 4.75 5.00 5.50 6.00 6.50 0.361 ± 0.004 0.429 ± 0.002 0.518 ± 0.011 0.523 ± 0.006 0.673 ± 0.017 0.834 ± 0.025 0.832 ± 0.063 0.836 ± 0.052 2.027 ± 0.121 0.910 ± 0.053 0.200 ± 0.008 0.172 ± 0.002 0.230 ± 0.009 0.194 ± 0.004 0.234 ± 0.007 0.196 ± 0.005 0.198 ± 0.011 0.220 ± 0.010 0.171 ± 0.002 0.390 ± 0.013 Table 5. Results from the fragmentation calibration setup. The slope α of the power-law is increasing with increasing bulk mod- ulus K0. Remarkably, the size of the normalised biggest fragment remains nearly constant around µ ≈ 0.2 for K0 ≤ 6.0 kPa. are converging for higher spatial resolutions. Boundary condi- tions are necessary for all calibration setups presented in this work. Their treatment, which is a difficult issue in SPH, was re- solved by using boundary particles with vanishing acceleration at every time step. The dissipative properties of the artificial vis- cosity and its role for the stability of the simulation were investi- gated. Artificial viscosity was used to separate dust and glass ma- 18 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates terial which are highly different in the "stiffness" of their equa- tions of state. We investigated the crucial differences between 2D and 3D compaction calibration setup and proved the validity of the cor- rection factor for the intrusion depth (Eq. 26) that was already used without confirmation in Guttler et al. (2009b). A major dif- ference between experiments and simulations, a compaction that went too far down in the dust sample and along with this a com- paction of too much volume, was resolved by using the 3D setup. Using a series of 2D simulations of the compaction calibra- tion setup we predicted a new, more accurate, value for the mean pressure pm of the dynamic compressive strength relation. A 3D simulation with this value pm = 0.26 kPa was carried out. The results were compared with the experimental reference using the same benchmark features as in Guttler et al. (2009b). We found a good agreement. In Guttler et al. (2009b) it was already demonstrated that our code is in principle capable of simulating the bouncing of dust aggregates. With the bouncing calibration setup we now investigated the ability of the SPH code of quantitatively sim- ulating the elastic properties of the dust and the energy dissi- pation by compaction. We found that the bulk modulus as well as the compressive strength relation have significant influence. For smaller values of pm (low compressive strength) and higher values of the bulk modulus more energy is dissipated. Using pm = 0.26 kPa from the compaction calibration setup we de- duced that the uncompressed dust samples (φ ≈ 0.15) have a bulk modulus K0 ≈ 5 kPa. With this result we were able to de- cide between two differing experimental values in favour of the indirect determination by Weidling et al. (2009). An important application of this code will be pre- planetesimal collisions. Therefore we also tested the code for its ability to simulate fragmentation of dust aggregates quantita- tively correct. For this we used a fragmentation calibration setup and identified the bulk modulus as the quantity with the most in- fluence on the fragment distribution. Again using pm = 0.26 kPa from the compaction calibration setup we found the best agree- ment with the empirical reference for the bulk modulus K0 = 4.5 kPa. Remarkably this is consistent with the findings from the bouncing calibration setup, which represents a test for a totally different behaviour. The problem of almost no fragmentation for K0 = 300 kPa described in Guttler et al. (2009b) was hereby traced back to a wrong assumption for the bulk modulus. 7. Conclusion Schafer et al. (2007) have shown that the outcome of pre- planetesimal collisions crucially depend on their material prop- erties. If the issue of collisional growth is to be investigated by computer simulations, they suggest an implementation of experi- mentally measured material parameters and thorough calibration and comparison with laboratory experiments. The SPH code presented in this work complies with their requirements. Based on the experimental preparatory work by Guttler et al. (2009b) this code has been successfully tested with three kinds of calibration setups for the correct simulation of compaction, bouncing, and fragmentation. We conclude that we have developed a tool that features a sufficient accuracy for the investigation of the outcome of pre- planetesimal collisions in parameter ranges that are not accessi- ble to laboratory experiments. This is based on the assumption that the macroscopic properties calibrated for in this work do not change with increasing size. However, due to the thermodynam- ically simple porosity model used in the SPH code the area of application is restricted to a certain range of collision velocities. Collisions where shock propagation cannot be neglected are out- side the physically meaningful limit of this model. Changes of the mechanical properties with temperature and other thermody- namical effects like sintering and melting cannot be simulated. An extension of the porosity model and its implementation and calibration are necessary to broaden the parameter range (e.g. supersonic impacts) and to consider a realistic environment for pre-planetesimal collisions in the protoplanetary disc. Nevertheless, within its area of application our code will be used to produce a "catalogue of collisions". The influence of ob- ject sizes, porosity, relative velocity, rotation, and impact param- eter on the fragment distribution of pre-planetesimal collisions will be investigated in future work. Acknowledgements. The authors want to thank Serena E. Arena for many fruit- ful discussions and helpful comments about the manuscript. The SPH simula- tions were performed on the university and bwGriD clusters of the computing centre (ZDV) of the University of Tubingen. We thank M.-B. Kallenrode and the University of Osnabruck for providing access to the XRT setup. We also thank the anonymous referee for clarifying comments and suggestions. This project was funded by the Deutsche Forschungsgemeinschaft within the Forschergruppe 759 "The Formation of Planets: The Critical First Growth Phase" under grants Bl 298/7-1, Bl 298/8-1, and Kl 650/8-1. References Benz, W. & Asphaug, E. 1994, Icarus, 107, 98 Benz, W. & Asphaug, E. 1995, Computer Physics Communications, 87, 253 Blum, J. & Munch, M. 1993, Icarus, 106, 151 Blum, J. & Schrapler, R. 2004, Phys. Rev. Lett., 93, 115503 Blum, J., Schrapler, R., Davidsson, B. J. R., & Trigo-Rodr´ıguez, J. M. 2006, ApJ, 652, 1768 Blum, J. & Wurm, G. 2000, Icarus, 143, 138 Blum, J. & Wurm, G. 2008, ARA&A, 46, 21 Blum, J., Wurm, G., Kempf, S., et al. 2000, Physical Review Letters, 85, 2426 Brauer, F., Dullemond, C. P., & Henning, T. 2008, A&A, 480, 859 Davis, D. R. & Ryan, E. V. 1990, Icarus, 83, 156 Dominik, C. & Tielens, A. G. G. M. 1997, ApJ, 480, 647 Dullemond, C. P. & Dominik, C. 2005, A&A, 434, 971 Geretshauser, R. 2006, Master's thesis, Eberhard-Karls-Universitat Tubingen Gingold, R. A. & Monaghan, J. J. 1977, MNRAS, 181, 375 Grady, D. E. & Kipp, M. E. 1980, Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 17, 147 Guttler, C., Blum, J., Zsom, A., Ormel, C. W., & Dullemond, C. P. 2009a, A&A, submitted Guttler, C., Krause, M., Geretshauser, R. J., Speith, R., & Blum, J. 2009b, ApJ, 701, 130 Heim, L.-O., Blum, J., Preuss, M., & Butt, H.-J. 1999, Physical Review Letters, 83, 3328 Heisselmann, D., Fraser, H., & Blum, J. 2007, in Proceedings of the 58th International Astronautical Congress 2007, IAC-07-A2.1.02 Jutzi, M. 2008, PhD thesis, Universitat Bern Jutzi, M., Benz, W., & Michel, P. 2008, Icarus, 198, 242 Jutzi, M., Michel, P., Benz, W., & Richardson, D. C. 2009a, MNRAS, submitted Jutzi, M., Michel, P., Hiraoka, K., Nakamura, A. M., & Benz, W. 2009b, Icarus, in press Krause, M. & Blum, J. 2004, Physical Review Letters, 93, 021103 Libersky, L. D. & Petschek, A. G. 1990, in Advances in the Free-Lagrange Method, ed. H. E. Trease, M. J. Fritts, & W. P. Crowley (Springer Verlag), 248–257 Libersky, L. D., Petschek, A. G., Carney, T. C., Hipp, J. R., & Allahdadi, F. A. 1993, J. Chem. Phys., 109, 67 Libersky, L. D., Randles, P. W., Carney, T. C., & Dickinson, D. L. 1997, Int. J. Impact Eng., 20, 525 Lucy, L. B. 1977, ApJ, 82, 1013 Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425 Melosh, H. J. 1989, Oxford monographs on geology and geophysics, Vol. 11, Impact cratering: A geologic process, ed. H. J. Melosh (Oxford University Press) Monaghan, J. J. 2005, Reports on Progress in Physics, 68, 1703 Monaghan, J. J. & Gingold, R. A. 1983, J. Chem. Phys., 52, 374 Monaghan, J. J. & Lattanzio, J. C. 1985, A&A, 149, 135 Omang, M., Børve, S., & Trulsen, J. 2006, Journal of Computational Physics, 213, 391 R. J. Geretshauser et al.: Numerical Simulations of Highly Porous Dust Aggregates 19 Ormel, C. W., Paszun, D., Dominik, C., & Tielens, A. G. G. M. 2009, A&A, 502, 845 Ormel, C. W., Spaans, M., & Tielens, A. G. G. M. 2007, A&A, 461, 215 Paszun, D. & Dominik, C. 2006, Icarus, 182, 274 Paszun, D. & Dominik, C. 2008, A&A, 484, 859 Paszun, D. & Dominik, C. 2009, A&A, 507, 1023 Randles, P. W. & Libersky, L. D. 1996, Comp. Methods Appl. Mech. Engrg, 139, 375 Schafer, C., Speith, R., & Kley, W. 2007, A&A, 470, 733 Sirono, S.-I. 2004, Icarus, 167, 431 Teiser, J. & Wurm, G. 2009, MNRAS, 393, 1584 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2007, ApJ, 661, 320 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2008, ApJ, 677, 1296 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2009, ApJ, in press Weidenschilling, S. J. 1977, MNRAS, 180, 57 Weidenschilling, S. J. & Cuzzi, J. N. 1993, in Protostars and Planets III, ed. E. H. Levy & J. I. Lunine, 1031–1060 Weidenschilling, S. J., Spaute, D., Davis, D. R., Marzari, F., & Ohtsuki, K. 1997, Icarus, 128, 429 Weidling, R., Guttler, C., Blum, J., & Brauer, F. 2009, ApJ, 696, 2036 Wurm, G., Paraskov, G., & Krauss, O. 2005, Icarus, 178, 253 Zsom, A. & Dullemond, C. P. 2008, A&A, 489, 931 Zsom, A., Ormel, C. W., Guttler, C., Blum, J., & Dullemond, C. P. 2009, A&A, submitted
1602.02877
1
1602
2016-02-09T07:28:03
Applying Titius-Bode's Law on Exoplanetry Systems
[ "astro-ph.EP", "astro-ph.SR" ]
We report the application of Titius-Bode's law on 43 exoplanetary systems containing four or more planets. Due to the fact that most of these systems have their planets located within compact regions extending for less than the semi-major axis of Mercury we found the necessity to scale down the Titius-Bode law in each case. In this short article we present sample calculations for three systems out of the whole set. Results show that all systems studied are verifying the applicability of the law with high accuracy. Consequently our investigation verifies practically the scale invariance of Titius-Bode law. The results of this study buildup the confidence in predicting positions of the exoplanets according to Titius-Bode's law besides enabling diagnosing possible reasons of deviations.
astro-ph.EP
astro-ph
Applying Titius-Bode's Law on Exoplanetry Systems M. B. Altaie, Zahraa Yousef, A. I. Al-Sharif Department of Physics, Yarmouk University, 21163 Irbid, Jordan We report the application of Titius-Bode's law on 43 exoplanetary systems containing four or more planets. Due to the fact that most of these systems have their planets located within compact regions extending for less than the semi-major axis of Mercury we found the necessity to scale down the Titius- Bode law in each case. In this short article we present sample calculations for three systems out of the whole set. Results show that all systems studied are verifying the applicability of the law with high accuracy. Consequently our investigation verifies practically the scale invariance of Titius-Bode law. The results of this study buildup the confidence in predicting positions of the exoplanets according to Titius-Bode's law besides enabling diagnosing possible reasons of deviations. INTRODUCTION During the last two decades many exoplanetary systems have been discovered, most of these contain one or two planets, but a large number of systems contain more than three exoplanets. It would be worth to study the spacing between these planets in order to know whether the spacing satisfies the so-called Titius-Bode (hereafter TB) law or not. A positive result will provide support for this law and may inspire more studies to find its physical foundation. Historically the first support for the law came through its role in discovering Uranus where the position of the seventh planet in our solar system was predicted at 19.6 AU by this law. In 1781 William Herschel found that there is a planet at 19.218 AU which deviates only by 1.9% from the expected distance. Uranus could have been discovered earlier if the TB relation had been taken more seriously.1Although the TB law did not give an accurate prediction for the distance of Neptune from the Sun, it played an important role in discovering the Asteroid Belt, where the median of position was predicted with high accuracy. Many modified formulas were presented for the TB law because of the large deviation suffered in its prediction in the case of the outer planets. The first modification was suggested by Wurm in 1803 and the last to date was in 2009 presented by Pankovic and Radacovic.2 The configuration of the discovered exoplanetary systems can be compared with our solar system. In a given multiple exoplanetary system there could be planets that have not been discovered yet. That may occur because of the lack of sufficient data, such as planetary masses or radii that are below the detection limits.3 Here the TB law could be a useful tool for making predictions about the positions of exoplanets in multiple 1 exoplanetary systems. The aim of this study is to test this idea. We first try to check if the multiple exoplanetary systems fit well with the TB law, then we use the law to make predictions about the positions of undetected exoplanets in systems under our study. Several authors have considered applying the TB law on spacing between of the exoplanets in multiple planetary systems. These studies followed different approaches and have different strategies. Poveda and Lara4 tried to investigate if the five planets of the system 55 Cancri (also called 55Cnc) fit the TB law. They found that if the largest known semi-major axis to be allocated for the sixth planet then the exponential form of the TB law fits well with the five observed semi-major axes for that system. Hence they predict the existence of a planet at a=0.2 AU in the large gap between the fourth planet (at a=0.781 AU) and the sixth (at a=5.22 AU). Moreover, they predicted another new planet for the system at a=15.0 AU. After publishing their paper, the semi major axis for the most inner planet was updated from 0.038 AU to 0.015AU as reported by Winn.5 Raymond et al.6 mapped out the region in 55 Cancri between the known planets 55Cnc f and 55Cnc d where an additional planet, designated g, might exist. Because there is a wide region of stability between them, they found that this region can accommodate a Saturn-mass planet which is a giant planet so they suggested that there could be even two or three additional planets. Lovis et al.7 applied the generalized TB relation to multiple exoplanetary systems detected by the radial velocity method only, they found reasonable fits. But they did not make any predictions for positions of new planets. Cuntz8 used the TB relation to predict the existence of new planets in the 55 Cancri system. He predicted the existence of four planets at distances at 0.085 AU, 0.41 AU, 1.50 AU and 2.95 AU from their host star. He also predicted the distance of the next possible outer planet in the system to be between 10.9 AU to 12.2 AU. None of the above predictions is confirmed observationally. Bovaird and Lineweaver9 (2013) studied the applicability of TB relation on a sample of multiple exoplanetary systems containing four planets and more. They found that the majority of their sample adheres to the TB relation more than the solar system does. They inferred that if any system does not follow the TB relation as close as our Solar System there is a high probability that one or more exoplanets in this system are not detected so far. They used the TB relation to list their predictions for the existence 2 of 141 exoplanets in 68 multiple exoplanetary systems. Huang and Bakos (see ref. 3) used the Kepler mission data to search 97 planets of those studied by Bovaird and Lineweaver (ref. 9) in multi-planetary systems; they only found five planetary candidates around their predictions. They considered that the remaining predicted planets could not be discovered by Kepler mission yet. They conclude that the ability of the TB law at making predictions in extrasolar planetary systems is questionable. In this letter we present sample calculations for three exoplanetary systems and report a summary for our investigation of the 43 systems containing planets giving the results for the adherence to the TB law presented in terms of an average of the fitting percentage. Detailed calculations for the set of the 43 systems are to be given elsewhere.10 Method and Results The strategy of our study is based on the understanding that TB law is describing a configuration rather than a law with fixed numbers. This means that the numbers given by the sequence 0.4, 0.7…, etc. are only set as a scale for the configurations. The way that the TB law was deduced confirms this understanding. Accordingly, we feel that a scaling of the configuration is always possible, and therefore can be taken as a reference for the application of the TB law as long as the relative arrangement is preserved. From this point of view this study is a novel one that could open wide door for other studies in the field and motivating new constrains on the way exoplanetary systems are configured. We prepared a survey for the exoplanets reported by NASA Exoplanets Archive about the multiple exoplanetary systems under study. The confirmed exoplanets discovered up to date (21/1/2016) are 1932 planets. 259 of these exoplanets belong to 58 multiple exoplanetary systems (in which the same star hosts more than 3 planets). We have excluded 15 multiple exoplanetary systems with 66 exoplanets for lack of data about their semi major-axes. Our sample contains the remaining 193 exoplanets that are hosted by 43 stars. Thirty five systems host 153 of these planets were detected by the transit method, seven systems involves 36 planets were detected by the radial velocity method and only one system (HR8799) with 4 planets was detected by direct imaging. Our data about exoplanets are obtained from the NASA Exoplanets Archive.11 TB law scaling We noticed that most of the known exoplanetary systems under study have their 3 planets located within regions of distances less than the semi-major axis of Mercury. This means that we need to scale the TB law for each of those systems. For this purpose our method was to choose one of the planets of the given system to be satisfying the TB law at a given order of our choice, calculate the scaling factor which comes to be the ratio between the planet's semi-major axis and the corresponding distance in TB law, accordingly we go on scaling all the system with the same factor. Such a scaling of the distances may also be understood as a scaling of the measuring units as we can calculate this factor for any system by dividing the new astronomical unit by the corresponding distance in the TB law. Every individual system has its own scaling factor. The choice of the order of the planet to fit the TB law is done with the best fit and minimum deviation from the distance given by the law for that order. Once we get the scaling factor the configuration of the system becomes known to us. Accordingly, we can place the planets in their positions and the order of TB law which we have chosen up to 10 positions. This resulted, in many cases, in leaving some empty positions for planets that are not discovered yet, or that which even may not have been formed at all as it is the case with the Asteroid Belt. These empty positions are considered as predictions from our point of view. The justification for our scaling was theoretically validated by Garner and Dubrulle12 who had showed that the law is both scale and rotation invariant. This important work has established the theoretical bases for the invariance of TB law and it our work which we are presenting here is the practical verification for that theoretical work of Garner and Dubrulle, where the results shows that the scaled exoplanetary systems are satisfying the TB law with high accuracy, on average is 90.4%. Fitting results Out of the 43 systems that we have studied, we will choose only three systems. The first is the one with the best fitting percentage of 97.78% which is Kepler 215 system. The second is arbitrarily chosen from the systems which showed near the average fitting percent of about 90.4% and the last system is the one which showed the worst fitting percentage of 65.27%, this is Kepler 444 system. KEPLER-215 This is a G-type star similar to our Sun, its mass is about 0.98 solar mass with radius of about 1.0 solar radius and surface temperature of about 5739K. This exoplanetary system harbors 4 confirmed planets which were detected by the transit method. The 4 positions of the confirmed planets in the system are as shown in the second column of Table (1). The scaling factor for this system was calculated by positioning the second planet which has a semi-major axis of 0.311 AU in order n=3, accordingly the scaling factor should be 0.113. The planets configuration becomes as shown in Table (1). We predict an inner planet at order n=1 positioned at 0.0452 AU and the rest of the predicted planets are all in the outer region at 0.5876 AU and beyond. We have shown the absolute and the relative fitting errors marking the deviation from the standard TB law in order to expose the error contribution from each planet. This might be useful on analyzing the pathology of the system and may help designating some reasons for deviations from the TB law. With the suggested configuration the average fitting percentage for the system is calculated to be 97.78%. If we look at the configuration we see high similarity with the solar system apart from the scale factor. Table (1) System Kepler-215 ao - ap abs. error rel. error n 1 2 3 4 5 6 7 8 9 10 0.0452 0.0840 0.0791 0.1130 0.1130 0.1850 0.1808 0.3164 0.314 0.5876 1.1300 2.2148 4.3844 8.7236 - - - - - - - 0.0049 0.0000 0.0042 0.0024 0.0583 0.0000 0.0227 0.0076 - - - - - - - - - - 2.22% 97.78 Avg. rel. error Fitting percent The fitting of the planets for this system is shown in Fig. (1). Figure (1) The best TB fitting possible for system Kepler-215. This gives the highest fitting percentage among the set of exoplanetary systems considered in our study. The triangles mark the data from observations and the stars mark data from our predictions. 5 HD 160691 The star is again a G3 type that has a mass of about 1 solar mass with radius of 1.245 solar radius. The surface temperature of the star is about 2700K. The age of the star is not well defined but there are some references claiming it is about 6 Gyrs old. The system contains four confirmed planets that have been discovered by radial velocity method. The semi-major axes of the confirmed planets are as shown in the second column of Table (2). Clearly the system has a very large gap between the 3rd and the 4th planet where the distance is about 3.738 AU, a large distance in comparison with the size of the system. Consequently, the fitting allow us to predict the existence of several planets: an inner one at 0.0540 AU, and three intermediate ones at 0.1349 AU, 0.2159 AU and at 0.3778 AU. Two more planets are predicted in the outer region at 1.497 AU and at 2.6445 AU. Outer planets at 10.416 AU and more may also exist. The average fitting percentage for this system is 90.61% which is within the average obtained for the 43 systems which we have studied. Table (2) System HD 160691 ao - ap - abs. error rel. error n 1 2 3 4 5 6 7 8 9 10 - - - 0.0540 0.0909 0.0944 0.1349 0.2159 0.3778 0.9210 0.7016 1.497 1.3492 2.6445 5.2350 5.2350 10.416 - - - 0.0049 0.0583 - - - - - - 0.2194 0.1478 - 0.2382 0.0987 - 0.0000 0.0000 - - 9.39% 90.61 Avg. rel. error Fitting percent The fitting is chosen to have the fourth planet in position n=9. This will make the scaling factor become 0.13492. Accordingly the configuration of the planets for the best fitting is obtained as gives in Table (2). The fitting curve for the system is as depicted in Fig. (2). 6 Figure (2) The second best TB fitting possible for system HD 160691. This has been chosen for display because it has a fitting percentage of about the average for the set studied. The dark squares mark the data from observations and the stars mark data from our predictions. KEPLER-444 This is a system with extreme dynamical conditions. It harbors three stars (triple star system). There are 5 confirmed planets orbiting Kepler 444A which is a K-type star in a tightly packed region. The mass of this primary star is about 0.76 solar mass with a radius of about 0.75 solar radius. This star is believed to be 11.2 Gyrs old. This is the oldest known star to date harboring exoplanets. The other two stars are Kepler 444B and Kepler 444C are small M-type stars orbiting the center of mass of the whole system at a distance of about 40 AU. So the whole known system composed of the three stars and the five planets has a size less than our solar system. As for the Kepler 444A system the five planets orbit the host star at distances shown in Table (3). All the predicted planets are in the outer region at 0.2262 AU and beyond. n 1 2 3 4 5 6 7 8 9 10 Table (3) System Kepler-444 ao ap abs. error rel. error 0.0418 0.0174 0.0488 0.0304 0.0600 0.0435 0.0696 0.0696 0.0811 0.1218 0.2262 0.4350 0.8526 1.6878 3.3582 - - - - - 0.0244 0.0148 0.0165 0.0000 0.0407 0.5835 0.3762 0.2750 0.0000 0.5018 - - - - - - - - - - Avg. rel. error Fitting percent 34.73% 65.27 To obtain a fitting for this system we found that the best possible fit will be obtained 7 if we position the 4th planet in n=4 of the TB order. This gives a scaling factor of 0.04350. Consequently we obtain the configuration given in Table (3) where we see the confirmed planets are arranged sequentially. The fitting percentage is 65.27, which is very poor as compared to the other planetary systems in this study. The reason is clear; the dynamical setup of the system having three stars at the play makes the system highly unstable. The orbits of this system are coinciding. Orbits of planets c and d, e and f are coinciding. This causes instability in the system. The fitting curve of the system is shown in Fig (3). Figure (3) The best fitting possible for system Kepler-444. This gives the lowest fitting percentage among the set of exoplanetary systems considered in our study. Discussion and conclusions In this letter we have chosen to present three systems out of the set of 43 exoplanetary systems which have been fitted in accordance with the TB law. We find that the majority of our large set adheres to the TB law with an average adherence of about 90.40%. The examples given in this letter have been chosen to show the highest, the median and the lowest of the fitting percentages. Generally we notice that the adherence to the TB law for exoplanetary systems under our study has different percentages. There are several reasons for these differences; one of them is the age of the system indicated by the age of the hosting star. Young systems with age of less than 1 Gyr are normally under formation and might not have dynamically settled yet. Such systems suffer jostling during planetary system formation; this is a messy process it can include planetary scattering and migration (see reference 9). This 8 applies to exoplanetary system Kepler-90 (not presented here) which we found to adhere to TB low with fitting percentage of 72.22 and is estimated to have an age of 0.5 Gyr. The other reason for deviating from the TB law may happen if the system under consideration may be too old, and consequently may suffer from dynamical changes that affect the positions of the planets in the system. This applies to system Kepler 444 presented above which is thought to be 11.2 Gyr old. Similarly, the Kepler-11 system is found to have fitting percentage of 73.24 which is much below the average. To get an idea about the adherence of our solar system to TB law using the method presented here we note that the overall fitting percentage we get for it is about 85.80%. This is less than the average which we have obtained for the set of 43 systems that we have studied. However, it should be noted that the relative error we get for Neptune is about 27.46% whereas the relative errors for the other 7 planets inside the orbit of Neptune including the Asteroid Belt ranges from 1.0% to 5.26%. If we would include Pluto in these calculations we find that the relative error in its position is about 95.74%. But if we exclude Pluto and do not count it as a planet as it was decided by the IAU in 2006, then we obtain a fitting of 94.89%. The results presented in our study shows that, assuming high confidence in the TB law applicability and adherence, we can have systematic speculations in respect of diagnosing the status of some pathological exoplanetary system which shows low fitting percentage. Such speculative diagnosis can be tested by observations, including possible errors in observations or calculations. For example, the system Kepler-33 is found to have a fitting percentage of 78.09 which is again much below the average percentage for the set we considered, despite the fact that it is nearly an ideal system harboring 5 planets with the host star being Sun-like star with age of about 4.3 Gyr. This suggest that here we have a pathological case for further studies. Other systems are also analyzed in this respect and could prove that the degree of adherence to TB law might be very useful for analyzing exoplanetary systems. 1 Nieto, M., The Titius-Bode Law of Planetary Distances: Its History and Theory. Pergamon Press Oxford, New York, (1972). 2Flores-Guti'errez, J. et al., A variant of the Titius-Bode law, Revista Mexicana de Astronom'a y Astrof'sica, 47, 173-184, (2011). 3Huang C., Bakos G., Testing the Titius-Bode law predictions for Kepler, MNRAS 442(1), 674, (2014). 4Poveda A., Lara P., The exo-planetary system of 55 Cancri and the Titius-Bode law, Rev. Mex. Astron. Astrofis , 44, 243 (2008) 9 5Winn J. N. et al., The exo-planetary system of 55 Cancri and the Titius-Bode law, ApJ. 737, L18 (2011). 6Raymond S. N. Rory Barnes, and Noel Gorelick, A dynamical perspective on additional planets in 55 Cancri, ApJ, 689, 478 (2008). 7 Lovis C. et al., The HARPS search for southern extra-solar planets XXVIII. Up to seven planets orbiting HD 10180: probing the architecture of low-mass planetary systems, A&A 528, A112 (2011). 8 Cuntz M., Application of the Titius–Bode rule to the 55 Cancri system: tentative prediction of a possibly habitable planet, PASJ, 64, 73 (2011). 9Bovaird T., and Lineweaver, C. H., Exoplanet predictions based on the generalized Titius–Bode relation, MNRAS, 435, 1126 (2013) 10 Altaie M.B., Yousef, Z., and Al-Sharif, A.I .Verification the applicability and the scale invariance of Titius- Bode's law on Exoplanets Systems, in preparation (2016). 11 NASA Exoplanet Archive (http://exoplanetarchive.ipac.caltech.edu/). 12 Graner, F. and Dubrulle, B., Titius- Bode laws in the solar system. I: Scale invariance explains everything, Astron. Astrophys. 282, 262-268 (1994). 10
1505.02938
1
1505
2015-05-12T09:54:01
Tilting Jupiter (a bit) and Saturn (a lot) During Planetary Migration
[ "astro-ph.EP" ]
We study the effects of planetary late migration on the gas giants obliquities. We consider the planetary instability models from Nesvorny & Morbidelli (2012), in which the obliquities of Jupiter and Saturn can be excited when the spin-orbit resonances occur. The most notable resonances occur when the $s_7$ and $s_8$ frequencies, changing as a result of planetary migration, become commensurate with the precession frequencies of Jupiter's and Saturn's spin vectors. We show that Jupiter may have obtained its present obliquity by crossing of the $s_8$ resonance. This would set strict constrains on the character of migration during the early stage. Additional effects on Jupiter's obliquity are expected during the last gasp of migration when the $s_7$ resonance was approached. The magnitude of these effects depends on the precise value of the Jupiter's precession constant. Saturn's large obliquity was likely excited by capture into the $s_8$ resonance. This probably happened during the late stage of planetary migration when the evolution of the $s_8$ frequency was very slow, and the conditions for capture into the spin-orbit resonance with $s_8$ were satisfied. However, whether or not Saturn is in the spin-orbit resonance with $s_8$ at the present time is not clear, because the existing observations of Saturn's spin precession and internal structure models have significant uncertainties.
astro-ph.EP
astro-ph
To be submitted to ApJ Tilting Jupiter (a bit) and Saturn (a lot) During Planetary Migration David Vokrouhlick´y Institute of Astronomy, Charles University, V Holesovick´ach 2, Prague 8, CZ-180 00, Czech Republic [email protected] and David Nesvorn´y Department of Space Studies, Southwest Research Institute, 1050 Walnut Street, Suite 300, Boulder, CO 80302, USA [email protected] ABSTRACT We study the effects of planetary late migration on the gas giants obliqui- ties. We consider the planetary instability models from Nesvorn´y & Morbidelli (2012), in which the obliquities of Jupiter and Saturn can be excited when the spin-orbit resonances occur. The most notable resonances occur when the s7 and s8 frequencies, changing as a result of planetary migration, become commensu- rate with the precession frequencies of Jupiter's and Saturn's spin vectors. We show that Jupiter may have obtained its present obliquity by crossing of the s8 resonance. This would set strict constrains on the character of migration during the early stage. Additional effects on Jupiter's obliquity are expected during the last gasp of migration when the s7 resonance was approached. The magnitude of these effects depends on the precise value of the Jupiter's precession constant. Saturn's large obliquity was likely excited by capture into the s8 resonance. This probably happened during the late stage of planetary migration when the evo- lution of the s8 frequency was very slow, and the conditions for capture into the spin-orbit resonance with s8 were satisfied. However, whether or not Saturn is in the spin-orbit resonance with s8 at the present time is not clear, because the existing observations of Saturn's spin precession and internal structure models have significant uncertainties. -- 2 -- Subject headings: celestial mechanics -- planets and satellites: individual (Jupiter, Saturn) -- planets and satellites: dynamical evolution and stability 1. Introduction It is believed that the orbital architecture of the solar system was significantly altered from its initial state after the dissipation of the protosolar nebula. The present architecture is probably a result of complex dynamical interaction between planets, and between planets and planetesimals left behind by planet formation. This becomes apparent because much of what we see in the solar system today can be explained if planets radially migrated, and/or if they evolved through a dynamical instability and reconfigured to a new state (e.g., Malhotra 1995; Thommes et al. 1999; Tsiganis et al. 2005). 2008; Morbidelli et al. While details of this process are not known exactly, much has been learned about it over decade by testing various migration/instability models against various constraints. Some of the most important constraints are provided by the terrestrial planets and the populations of small bodies in the asteroid and Kuiper belts (e.g., Gomes et al. 2005; Minton & Malhotra 2009, 2011; Levison et al. 2010; Nesvorn´y 2015). Processes related to the giant planet instability/migration were also used to explain capture and orbital distribution of Jupiter and Neptune Trojans (e.g., Morbidelli et al. 2005; Nesvorn´y & Vokrouhlick´y 2009; Nesvorn´y et al. 2014a) and irregular satellites (e.g., Nesvorn´y et al. 2007, 2014b). Some of the most successful instability/migration models developed so far postulate that the outer solar system contained additional ice giant that was ejected into interstellar space by Jupiter (e.g., Nesvorn´y 2011; Nesvorn´y & Morbidelli 2012; Batygin et al. 2012). The orbits of the four surviving giant planets evolved in this model by planetesimal-driven migration and by scattering encounters with the ejected planet. In this work, we use this framework to investigate the behavior of Jupiter's and Saturn's obliquities. The obliquity, ε, is the angle between the spin axis of a planet and the normal to its orbital plane. The core accretion theory applied to Jupiter and Saturn implies that their primordial obliquities should be very small. This is because the angular momentum of the rotation of these planets is contained almost entirely in their massive hydrogen and helium envelopes. The stochastic accretion of solid cores should therefore be irrelevant for their current obliquity values (see Lissauer & Safronov 1991 for a discussion), and a symmetric inflow of gas on forming planets should lead to ε (cid:39) 0. The present obliquity of Jupiter is εJ = 3.1◦, which is small, but not quite small enough for these expectations, but that of Saturn is εS = 26.7◦, which is clearly not. -- 3 -- Ward & Hamilton (2004) and Hamilton & Ward (2004) noted that the precession frequency of Saturn's spin axis has a value close to s8 (cid:39) −0.692 arcsec yr−1, where s8 is the mean nodal regression of Neptune's orbit (or, equivalently, the eighth nodal eigenfrequency of the planetary system; e.g., Applegate et al. 1986; Laskar 1988). Similarly, Ward & Canup (2006) pointed out that the precession frequency of Jupiter's spin axis has a value close to s7 (cid:39) −2.985 arcsec yr−1, where s7 is the mean nodal regression of Uranus's orbit. These findings are significant because they raise the possibility that the current obliquities of Jupiter and Saturn have something to do with the precession of the giant planet orbits. Specifically, Ward & Hamilton (2004) and Hamilton & Ward (2004) suggested that the present value of Saturn's obliquity can be explained by capture of Saturn's spin vector in a resonance with s8. They proposed that the capture occurred when Saturn's spin vector precession increased as a result of Saturn's cooling and contraction, or because s8 decreased during the depletion of the primordial Kuiper belt. They showed that, if the post-capture evolution is conveniently slow, the spin-orbit resonance (also known as the Cassini resonance, see Section 2) is capable of increasing Saturn's obliquity to its current value. While changes of precession or s8 during the earliest epochs could have been important, it seems more likely that capture in the spin-orbit resonance occurred later, probably as a result of planetary migration (Bou´e et al. 2009). This is because both s7 and s8 signifi- cantly change during the instability and subsequent migration. Therefore, if the spin-orbit resonances had been established earlier, they would not survive to the present time. Bou´e et al. (2009) studied various scenarios for resonant tilting of Saturn's spin axis during the planetary migration and found that the present obliquity of Saturn can be explained only if the characteristic migration time scale was long and/or if Neptune's inclination was substantially excited during the instability. Since Neptune inclination is never large in the instability/migration models of Nesvorn´y & Morbidelli (2012; hereafter NM12), typically iN < 1◦, the migration timescales presumably need (note that Bou´e et al. (2009) did not investigated these low-i cases in detail) to be very long (see Section 3.3). Interestingly, these very long migration timescales are also required from other constraints (e.g., Morbidelli et al. 2014; Nesvorn´y 2015). They could be achieved in the Nesvorn´y & Morbidelli (2012) models if Neptune interacted with an already depleted planetesimal disk during the very last stages of the migration process. As for the obliquity of Jupiter, Ward & Canup (2006) suggested that the present value is due to the proximity of the spin precession rate to the s7 frequency. In fact, the obliquities of Jupiter and Saturn represent a much stronger constraint on the instability/migration models than was realized before. This is because the constraints from the present obliquities of Jupiter and Saturn must be satisfied simultaneously (McNeil & Lee 2010; Brasser & Lee 2015). For example, in the initial configuration of planets in the -- 4 -- NM12 models, the s8 frequency is much faster than the precession constants of both Jupiter and Saturn. This means that the s8 mode, before approaching Saturn's precession frequency and exciting its obliquity, must also have evolved over the precession frequency of Jupiter's spin vector. This leads to a conundrum, because if the crossing were slow, Jupiter's obliquity would increase as a result of capture into the spin-orbit resonance with s8. If, on the other hand, the general evolution at all stages were fast, the conditions for capture of Saturn into the spin-orbit resonance with s8 may not be be met (e.g., Bou´e et al. 2009), and Saturn's obliquity would stay small.1 A potential solution of this problem would be to invoke fast evolution of s8 early on, during the crossing of Jupiter's precession frequency, and slow evolution of s8 later on, such that Saturn's spin vector can be captured into the spin-orbit resonance with s8 during the late stages. This can be achieved, for example, if the migration of the outer planets was faster before the instability, and slowed down later, as the outer Solar System progressed toward a more relaxed state. As we show in Section 3.1, the jumping-Jupiter models developed in NM12 provide a natural quantitative framework to study this possibility. In Sec. 2 we first briefly review the general equations for the spin-orbit dynamics. Then, in Sec. 3, we investigate the behavior of the spin vectors of Jupiter and Saturn in the instability/migration models of NM12. We find that the constraints posed by Jupiter's and Saturn's obliquities can be satisfied simultaneously in this class of models, and derive detailed conditions on the migration timescales and precession constants that would provide a consistent solution. 2. Methods 2.1. Parametrization using non-singular spin vector components Consider a planet revolving about the Sun and rotating with angular velocity ω about an instantaneous spin axis characterized by a unit vector s. With solar gravitational torques applied on the planet, ω remains constant, but s evolves in the inertial space according to 1Note that the present obliquities of Uranus and Neptune are not an important constraint on planetary migration, because their spin precession rates are much slower than any secular eigenfrequencies of orbits (now or in the past). Therefore, the secular spin-orbit resonances should not occur for these planets, and giant impacts may be required to explain their obliquities (e.g., Morbidelli et al. 2012, and references therein). (e.g., Colombo 1966) -- 5 -- = −α (c · s) (c × s) , ds dt where c denotes a unit vector along the orbital angular momentum, and α = 3 2 GM(cid:12) ω b3 J2 + q λ + l (1) (2) √ is the precession constant of the planet. Here, G is the gravitational constant, M(cid:12) is the 1 − e2, where a and e are the orbital semimajor axis and ec- mass of the Sun, b = a centricity, J2 is the quadrupole coefficient of the planetary gravity field, and λ = C/M R2 is the planetary moment of inertia C normalized by a standard factor M R2, where M is planet's mass and R its reference radius (to be also used in the definition of J2). The term q = 1 j(mj/M )(aj/R)2 is an effective, long-term contribution to the quadrupole coefficient 2 due to the massive, close-in regular satellites with masses mj and planetocentric distances j nj)/(M R2ω) is the angular momentum content of the satellite orbits (nj denotes their planetocentric mean motion) normalized to the characteristic value of the planetary rotational angular momentum. aj, and l =(cid:80) (cid:80) j(mja2 For both Jupiter and Saturn, q is slightly dominant over J2 in the numerator of the last term in Eq. (2), while l is negligible in comparison to λ in the denominator. Spacecraft observations have been used to accurately determine the values of J2, q and l. On the other hand, λ cannot be derived in a straightforward manner from observations, because it depends on the structure of planetary interior. Using models of planetary interior, Helled et al. (2011) determined that Jupiter's λJ is somewhere in the range between 0.2513 and 0.2529 (here rescaled for the equatorial radius R = 71, 492 km of the planet). This would imply Jupiter's precession constant αJ to be in the range between 2.754 arcsec yr−1 and 2.772 arcsec yr−1. These values are smaller than the ones considered in Ward & Canup (2006), if their proposed small angular distance between Jupiter's pole and Cassini state C2 is correct. Similarly, Helled et al. (2009) determined Saturn's precession constant αS to be in the range between 0.8443 arcsec yr−1 and 0.8447 arcsec yr−1. Again, these values differ from those inferred by Ward & Hamilton (2004) and Hamilton & Ward (2004), if a resonant confinement of the Saturn's spin axis in their proposed scenario is true. This difference has been noted and discussed in Bou´e et al. (2009). If Helled's values are correct, Saturn's spin axis cannot be presently locked in the resonance with s8. However, it seems possible that of αS derived in Helled et al. (2009) may have somewhat larger uncertainty than reflected by the formal range of the inferred λ values. Also, the interpretation is complicated by the past orbital evolution of planetary satellites which may have also contributed to changes of αS (and αJ). For these reason, and in the spirit of previous studies (e.g., Ward & Hamilton -- 6 -- 2004; Ward & Canup 2006; Bou´e et al. precession constant values α for both Jupiter and Saturn. 2009), here we consider a wider range of the Assuming α constant for a moment, a difficult element preventing a simple solution of Eq. (1) is the time evolution of c. This is because mutual planetary interactions make their orbits precess in space on a characteristic timescale of tens of thousands of years and longer. In addition, during the early phase of planetary evolution, the precession rates of planetary orbits may have been faster due to the gravitational torques from a massive population of planetesimals in the trans-Neptunian disk. In the Keplerian parametrization of orbits, the unit vector c depends on the inclination I and longitude of node Ω, such that cT = (sin I sin Ω,− sin I cos Ω, cos I). Traditionally, the difficulties with the time dependence in Eq. (1) are resolved using a transformation to a reference frame fixed on an orbit, where cT = (0, 0, 1). The transformation from the inertial to orbit coordinate frames is achieved by applying a 3-1-3 rotation sequence with the Eulerian angles (Ω, I,−Ω). This transforms Eq. (1) to the following form = − [α (c · s) c + h] × s, ds dt (3) where now the planetary spin vector s is expressed with respect to the orbit frame, and c is now a unit vector along the z-axis. In effect, the time dependence has been moved to the vector quantity hT = (A,B,−2C) with A = cos Ω I − sin I sin Ω Ω, B = sin Ω I + sin I cos Ω Ω, C = sin2 I/2 Ω, (4a) (4b) (4c) and the over-dots mean time derivative. A further development consists in introducing complex and non-singular parameter √−1 is the complex unit) that replaces I and Ω. First order ζ = sin(I/2) exp(ıΩ) (ı = perturbation theory for quasi-circular and near-coplanar orbits indicates ζ for each planet can be expressed using a finite number of the Fourier terms with the si frequencies uniquely dependent on the orbital semimajor axes and masses of the planets, and amplitudes set by the initial conditions (e.g., Brouwer & Clemence In the models from non- linear theories or numerical integrations, ζ can still be represented by the Fourier expansion ζ =(cid:80) Ai exp[ı(sit + φi)] (e.g., Applegate et al. 1986; Laskar 1988), with the linear terms 1961). having typically the largest amplitudes Ai. As discussed in Section 1, terms with present frequencies s7 (cid:39) −2.985 arcsec yr−1 and s8 (cid:39) −0.692 arcsec yr−1 are of a particular importance in this work. The s7-term is the largest -- 7 -- in Uranus' ζ representation, and the s8-term is the largest in Neptune's ζ representation, and these terms also appear, though with smaller amplitudes, in the ζ variable of Jupiter and Saturn, because mutual planetary interactions enforce all fundamental frequency terms to appear in all planetary orbits. In terms of ζ, Eqs. (4a-c) become (cid:19) − ıζ C (cid:19) , , (5a) (5b) (cid:18)dζ dt − ζ d¯ζ dt 2(cid:112) (cid:18) 1 2ı 1 − ζ ¯ζ dζ ¯ζ dt A + ıB = C = where ¯ζ is complex conjugate to ζ. For small inclinations, relevant to this work, we therefore find that A + ıB (cid:39) 2 (dζ/dt), and C (cid:39) 0. Another important aspect is of the problem is that Eq. (3) derives from a Hamiltonian such that H(s; t) = α 2 (c · s)2 + h · s, = −∇sH × s. ds dt (6) (7) This allowed Breiter et al. (2005) to construct an efficient Lie-Poisson integrator for a fast propagation of the secular evolution of planetary spins. In Section 3, we will use the leap- frog variant of the Hamiltonian's LP2 splitting from Breiter et al. (2005). To propagate s through a single integration step, Breiter et al. (2005) method requires that the orbital semimajor axis, eccentricity and c are provided at times corresponding to the beginning and end of the step. These values can be supplied from an analytic model of orbit evolution, or be directly obtained from a previous numerical integrations of orbits where a, e and c were recorded with a conveniently dense sampling. 2.2. Parametrization using obliquity and precession angle The Hamiltonian formulation in Eq. (6) allows us to introduce several important con- cepts of the Cassini dynamics. A long tradition in astronomy is to represent s with obliquity ε and precession angle ψ such that sT = (sin ε sin ψ, sin ε cos ψ, cos ε). The benefit of this parametrization is that the unit spin vector is expressed using only two variables. The drawback is that resulting equations are singular when ε = 0. The conjugated momentum to ψ is X = cos ε. The Hamiltonian is then (e.g., Laskar & Robutel 1993) H(X, ψ; t) = X 2 − 2CX α 2 (8) -- 8 -- √ + 1 − X 2 (A sin ψ + B cos ψ) . For the low-inclination orbits, C is negligible, while A and B are expanded in the Fourier series with the same frequency terms as those appearing in ζ itself. A model of fundamental importance, introduced by G. Colombo (Colombo 1966; Henrard & Murigande 1987; Ward & Hamilton 2004), is obtained when only one Fourier term in A and B is considered. This Colombo model obviously serves only as an approximation of the complete spin axis evolution, since all other Fourier terms in A and B act as a perturbation. Nevertheless, the Colombo model allows us to introduce several important concepts that are the basis of the discussion in Section 3. In the Colombo model, the orbital inclination is fixed and the node precesses with a constant frequency. Put in a compact way, ζ = A exp[ı(st + φ)] is the single Fourier term, and A = sin I/2 and Ω = st + φ. Transformation to new canonical variables X(cid:48) = −X and ϕ = −(ψ + Ω), and scaling with the nodal frequency s, results in a time-independent Hamiltonian H(X(cid:48), ϕ) = κX(cid:48)2 − cos IX(cid:48) √ + sin I 1 − X(cid:48)2 cos ϕ, (9) where κ = α/(2s) is a dimensionless parameter. Note that the orbit-plane angle ϕ is mea- sured from a reference direction that is 90◦ ahead of the ascending node. The general structure of the phase flow of solutions H(X(cid:48), ϕ) = C, with C constant, derives from the location of the stationary points. Depending on the parameter values (κ, I), there are either two (κ < κ(cid:63)) or four (κ > κ(cid:63)) such stationary solutions (called the Cassini states). The critical value of κ reads (e.g., Henrard & Murigande 1987) (cid:0)sin2/3 I + cos2/3 I(cid:1)3/2 1 2 (10) κ(cid:63)(I) = Therefore, for small I, κ(cid:63) (cid:39) 1 2. The stationary solutions are located at ϕ = 0◦ or ϕ = 180◦ meridians in the orbital frame, and have obliquity values given by a transcendental equation . κ sin 2ε = − sin (ε ∓ I) , (11) with the upper sign for ϕ = 0◦, and the lower sign for ϕ = 180◦. In the present work, we are mainly interested in the Cassini state C2 located at ϕ = 0◦. For Jupiter, the C2 state related to frequency s = s7 is subcritical since κ < 1 2 for all estimates of the Jupiter's precession constant found in the literature. Only two Cassini states exist in this regime, and s must circulate about C2. In the case of Saturn, κ > κ(cid:63) (cid:39) 1 2 for s = s8, four Cassini states exist in this situation, and s was suggested to librate in the resonant zone about C2. The configuration of vectors in C2 can be inferred from Eq. (11). If the inclination is significantly smaller than the obliquity ε, we have that α cos ε (cid:39) −s. Since the term on the left hand side of this relation is the precession frequency of the planet's spin (see Eq. 1), we find that the spin and orbit vectors will co-precess with the same rate about the normal vector to the inertial frame. Small resonant librations about C2 would reveal themselves by small departures of the spin vector from this ideal state. The maximum width ∆ε of the resonant zone in obliquity at the ϕ = 0◦ meridian can be obtained using an analytic formula (e.g., Henrard & Murigande 1987) -- 9 -- (cid:114) sin 2I sin 2ε4 sin ∆ε 2 = 1 κ , (12) where ε4 is the obliquity of the unstable Cassini state C4 (a solution of Eq. (11) at ϕ = 180◦ having the intermediate value of the obliquity). Figure 1, top panel, shows how the location of the Cassini states and the resonance width ∆ε depend on κ, which is the fundamental parameter that changes during planetary migration. For sake of this example we assumed orbital inclination I = 0.5◦ (note that the overall structure remains similar for even smaller inclination values considered in the next section, but would be only less apparent in the Figure). The C1 and C4 stationary solution bifurcate when κ = κ(cid:63) at a non-zero critical obliquity value ε(cid:63) = atan(tan1/3 I) (e.g., Henrard & Murigande 1987; Ward & Hamilton 2004). Note that ∆ε is significant in spite of a very small value of the inclination, which manifests through its dependence on a square root of sin 2I in (12). The bottom panels show examples of the phase portraits H(X(cid:48), ϕ) = C for both sub-critical κ < κ(cid:63) and super-critical κ > κ(cid:63) cases. 3. Results We now turn our attention to the evolution of Jupiter's and Saturn's obliquities dur- ing planetary migration. We first discuss the orbital evolution of planets in the instabil- ity/migration simulations of NM12 (Section 3.1). We then parametrize the planetary mi- gration before (stage 1) and after the instability (stage 2), and use it to study the effects on Jupiter's and Saturn's obliquities. The two stages are considered separately in Sections 3.2 and 3.3. -- 10 -- 3.1. The orbital evolution of giant planets NM12 reported the results of nearly 104 numerical integrations of planetary instability, starting from hundreds of different initial configurations of planets that were obtained from previous hydrodynamical and N -body calculations. The initial configurations with the 3:2 Jupiter-Saturn mean motion resonance were given special attention, because Jupiter and Saturn, radially migrating in the gas disk before its dispersal, should have become trapped into their mutual 3:2 resonance (e.g., Masset & Snellgrove 2001; Morbidelli & Crida 2007; Pierens & Nelson 2008). They considered cases with four, five and six initial planets, where the additional planets were placed onto resonant orbits between Saturn and the inner ice giant, or beyond the orbit of the outer ice giant. The integrations included the effects of the transplanetary planetesimal disk. NM12 experimented with different disk masses, density profiles, instability triggers, etc., in an attempt to find solutions that satisfy several constraints, such as the orbital configuration of the outer planets, survival of the terrestrial planet, and the distribution of orbits in the asteroid belt. NM12 found the dynamical evolution in the four planet case was typically too violent if Jupiter and Saturn start in the 3:2 resonance, leading to ejection of at least one ice giant from the Solar System. Planet ejection can be avoided if the mass of the transplanetary disk of planetesimals was large (Mdisk (cid:38) 50 MEarth, where MEarth is the Earth mass), but such a massive disk would lead to excessive dynamical damping (e.g., the outer planet orbits become more circular then they are in reality) and to smooth migration that violates constraints from the survival of the terrestrial planets, and the asteroid belt. Better results were obtained when the Solar System was assumed to have five giant planets initially and one ice giant, with mass comparable to that of Uranus or Neptune, was ejected into interstellar space by Jupiter (Nesvorn´y 2011; Batygin et al. 2012). The best results were obtained when the ejected planet was placed into the external 3:2 or 4:3 resonances with Saturn and Mdisk (cid:39) 20 MEarth. The range of possible outcomes was rather broad in this case, indicating that the present Solar System is neither a typical nor expected result for a given initial state. The most relevant feature of the NM12 models for this work is that the planetary migration happens in two stages (see Fig. 2). During the first stage, that is before the instability happens, Neptune migrates into the outer disk at (cid:39) 20 − 30 au. The migration is relatively fast during this stage, because the outer disk still has a relatively large mass. We analyzed several simulations from NM12 and found that Neptune's migration can be approximately described by an exponential with the e-folding timescale τ (cid:39) 10 Myr for Mdisk = 20 MEarth and τ (cid:39) 20 Myr for Mdisk = 15 MEarth. The instability typically happens in the NM12 models when Neptune reaches (cid:39) 28 au. The main characteristic of the instability is that planetary encounters occur, mainly between the extra ice giant and all other planets. -- 11 -- The instability typically lasts ∼ 105 years and terminates when the extra ice giant is ejected from the solar system by Jupiter. The second stage of migration starts after that. The migration of Neptune is much slower during this period, because the outer disk is now very much depleted. From simulations in NM12 we find that τ (cid:39) 30 Myr for Mdisk = 20 MEarth and τ (cid:39) 50 Myr for Mdisk = 15 MEarth. Moreover, rather then being precisely exponential, the migration slows down relative to an exponential with fixed τ , such as, effectively, the very late stages have larger τ values (τ ∼ 100 Myr) than the ones immediately following the instability. Uranus accompanies the migration of Neptune on timescales similar to those mentioned above. The frequencies s7 and s8, which are the most relevant for this work, are initially some- what higher than the precession constants of Jupiter and Saturn, mainly because of the torques from the outer disk. The extra term from the third ice giant initially located at (cid:39) 10 au has much faster frequency then the precession constants and does not interfere with the obliquity of Jupiter and Saturn during the subsequent evolution. The s7 and s8 frequencies slowly decrease during both stages. Their e-folding timescales may slightly differ from the migration e-folding timescales mentioned above, due to the non-linearity of the dependence of the secular frequencies on the semimajor axis of planets. Our tests show that they are about (90 − 95)% of the e-folding timescales of planetary semimajor axes. It is not known exactly, however, how small. From analyzing the behavior of frequencies in different simulations we found that s8 should cross the value of Jupiter's precession constant during the first migration stage, that is before the instability. The main characteristic of this crossing is that the planetary orbits are very nearly coplanar during this stage. The amplitude I in Eq. (9) should thus be very small. In Section 3.2, we consider amplitudes down to 0.005◦ (about 10 times smaller than the present value of I58, where I58 is the amplitude of the 8th frequency term in Jupiter's orbit), and show that the effects on Jupiter's obliquity are negligible if the amplitudes were even lower. The second characteristic of the first migration stage is that the evolution of s8 happens on a characteristic timescale of (cid:39) 10− 20 Myr. Since the total change of s8 during this interval is several arcsec yr−1, and the first stage typically lasts ∼ 10 − 20 Myr, the average rate of change is very roughly, as an order of magnitude estimate, ds8/dt ∼ 0.1 arcsec yr−1 Myr−1. The actual value of ds8/dt during crossing depends on several unknowns, including when exactly the crossing happens during the first stage. Also, the changes of s8 could have been slower if the first stage lasted longer than in the NM12 simulations, as required if the instability occurred at the time of the Late Heavy Bombardment (e.g., Gomes et al. 2005). In Section 3.2, we will consider values in the range 0.005 < ds8/dt < 0.05 arcsec yr−1 Myr−1, and show that the obliquity of Jupiter cannot be pumped up to its current value if ds8/dt > 0.05 arcsec yr−1 Myr−1 (assuming that I58 (cid:46) 0.05◦). -- 12 -- Interesting effects on obliquities should happen during the second migration stage. First, the s8 frequency reaches the value of the precession constant of Saturn. There are several differences with respect to the s8 crossing of Jupiter's precession constant during the first stage (discussed above). The orbital inclinations of planets were presumably excited to their current values during the instability. Therefore, the amplitude I68 should be comparable to its current value, I68 (cid:39) 0.064◦, during the second stage. We see this happening in the NM12 simulations. First, there is a brief period during the instability, when the inclinations of all planets are excited by encounters with the ejected ice giant. The inclination of Neptune is modest, at most (cid:39) 2◦, and is rather quickly damped by the planetesimal disk. Also, the invariant plane of the solar system changes by (cid:39) 0.5◦ when the third ice giant is ejected during the instability. The final inclinations are of this order. The current amplitudes are I58 (cid:39) 0.066◦ (s8 term in Jupiter's orbit) and I68 (cid:39) 0.064◦ (s8 term in Saturn's orbit) (see e.g., Laskar 1988). Another difference with respect the first stage is that the evolution of s8 is much slower during the second stage. If, as indicated by the NM12 integration, s8 changes by ∼ 1 arc- sec yr−1 in 100 Myr, then the average rate of change is very roughly ds8/dt ∼ 0.01 arc- sec yr−1 Myr−1. The actual rate of change can be considerably lower than this during the very late times, when the effective τ was lower than during the initial stages. Finally, during the very last gasp of migration, the s7 frequency should have approached the precession con- stant of Jupiter. We study this case in an adiabatic approximation when the rate of change of s7 is much slower than any other relevant timescale. We find that the present obliquity of Jupiter can be excited by the interaction with the s7 term only if the precession constant of Jupiter is somewhat larger than inferred by Helled et al. (2011), in accord with the results of Ward & Canup (2006). 3.2. The effects on Jupiter's obliquity during stage 1 Since s8 remains larger than αS during the first stage, we do not expect any important effects on Saturn's obliquity during this stage. If Saturn's obliquity was low initially, it should have remained low in all times before the instability. We therefore focus on the case of Jupiter in this section. From the analysis of the NM12 numerical simulation in Section 3.1, we infer that the s8 frequency crossed αJ during the first stage. The values of ds8/dt and I58 during crossing are not known exactly from the NM12 simulations, because they depend on details of the initial conditions. We therefore consider a range of values and determine how Jupiter's obliquity excitation depends on them. The results can be used to constrain future simulations of the planetary instability/migration. -- 13 -- We consider Colombo's model with only one Fourier term in ζ of Jupiter, namely that of the s8 frequency.2 The inclination term I58 in Jupiter's orbit was treated as a free parameter. The range of values was set to be between zero and roughly the current value of (cid:39) 0.066◦. As we discussed in Section 3.1, it is reasonable to assume that I58 was smaller than the current value, because the orbital inclination of planets should have been low in times before the instability. The value of αJ was obtained by rescaling the present value to the the semimajor axis of Jupiter before the instability (aJ (cid:39) 5.45 au). To do so we used Eq. (2) and assumed αJ ∝ a−3 J . No additional modeling of possible past changes of αJ, for instance due to satellite system evolution or planetary contraction, was implemented. The s8 frequency was slowly decreased from a value larger than αJ to a value smaller than αJ. Motivated by the numerical simulations of the instability discussed in Section 3.1, we assumed the initial s8 value of −4 arcsec yr−1 and let it decrease to −1.2 arcsec yr−1 by the end of each test. The rate of change, ds8/dt, was treated as a free parameter. The integrations were carried for several tens of Myr for the highest assumed rates and up to several hundreds of Myr for the lowest rates. We recorded Jupiter's obliquity during the last 5 Myr of each run, and computed the mean value εfin. Jupiter's initial spin axis was oriented toward the pole of the Laplacian plane. [The value of εfin reported in Fig. 3 was averaged over all possible phases of the initial spin axis on the H(X(cid:48), ϕ) = C level curve, with C defined by s oriented toward the pole of the Laplacian plane] Figure 3 shows the results. For most parameter combinations shown here the s8 res- onance swept over αJ without having the ability to capture Jupiter's spin vector in the resonance. This happened because ds8/dt was relatively large and I58 was relatively small, thus implying that the s8 frequency crossed the resonant zone in a time interval that was shorter than the libration period. Captures occurred only in extreme cases (largest I58 and smallest ds8/dt). These cases ended up generating very large obliquity values of Jupiter and are clearly implausible. The plausible values of ds8/dt and I58 correspond to the cases where Jupiter's obliquity was not excited at all, thus assuming that Jupiter obtained its present obliquity later, or was excited by up to (cid:39) 3◦. To obtain εJ (cid:39) 3◦, ds8/dt and I58 would need to have values along the bold line labeled 3 in Fig. 3, which extends diagonally in ds8/dt and I58 space. An example of a case, where the obliquity of Jupiter was excited to near (cid:39) 3◦ value, is shown in Fig. 4. 2We found that adding higher frequency terms, such as s6 and/or s7, into our simulation does not change results. -- 14 -- 3.3. Behavior of obliquities during stage 2 We now turn our attention to the second stage, when the migration slowed down and the obliquities of Jupiter and Saturn should have suffered additional perturbations. At the beginning of stage, that is just after the time of the instability, the s8 frequency is already lower than αJ, but still higher than αS, while the s7 frequency is higher than αJ. Since the s7 and s8 frequencies are slowly decreasing during the second stage, a possibility arises that Jupiter's obliquity was (slightly) excited when s7 approached αJ (e.g., Ward & Canup 2006), and that Saturn's obliquity was strongly excited by capture into the spin-orbit resonance with s8 (e.g., Ward & Hamilton 2004; Hamilton & Ward 2004; Bou´e et al. 2009). 3.3.1. Jupiter Ward & Canup (2006) suggested a possibility that Jupiter's present obliquity may be explained by the proximity of αJ to the current value of the s7 frequency. They showed that, if κ = αJ/(2s7) is sufficiently close to the critical value from Eq. (10), namely (cid:39) 1 2 for small inclinations, the obliquity of the Cassini state C2 may be significant. Thus, as s7 adiabatically approached to αJ, Jupiter's obliquity may have been excited along. As a supportive argument for this scenario they pointed out that ϕJ (cid:39) 0◦, where the Cassini state 2 is located. To test this possibility we run a suite of simulations, assuming an exponential conver- gence to the current value s7 (cid:39) −2.985 arcsec yr−1. Specifically, we set s7(t) = s7(0) + [s7 − s7(0)] exp(−t/τ ), where τ and s7(0) are parameters. The initial value s7(0) at the beginning of the second stage was obtained from the numerical simulation discussed in Section 3.1. Here we chose to use s7(0) = −3.5 arcsec yr−1, however we verified that the results are insensitive to this choice. The e-folding timescale τ depends on how slow or fast planets migrate. Given that the planetary migration is slow during the second stage, we chose τ = 100 − 200 Myr. This assures that the approach of s7(t) to αJ is adiabatical. The amplitude I57 is assumed to be constant and equal to its current value (I57 (cid:39) 0.055◦). Two additional parameters need to be specified: (i) Jupiter's initial spin state, and (ii) αJ. As for (i), the results discussed in Section 3.2 indicate that Jupiter's obliquity may have remained near zero during the first stage, if I58 was too small and/or ds8/dt was too fast, or could have been potentially excited to (cid:39) 3◦, if I58 and ds8/dt combined in the right way. Therefore, here we treat the obliquity of Jupiter at the beginning of stage 2 as a free parameter. As for (ii), as we discussed in Section 1, the present value of αJ somewhat uncertain. We therefore performed various simulations, where αJ takes on a number of -- 15 -- different values between 2.75 arcsec yr−1 and 2.95 arcsec yr−1. A similar approach has also been adopted by Ward & Canup (2006). Figure 5 reports the results. The top panel shows how the obliquity εC2 of the Cassini state C2 depends on the assumed value of αJ. This is calculated from Eq. (11). The trend is that εC2 increases with αJ, because the larger values of αJ correspond to a situation where the system is closer to the exact resonance with s7. If αJ < 2.8 arcsec yr−1, as inferred from models in Helled et al. (2011), εC2 is too small to significantly contribute to εJ. This case would imply that Jupiter's present obliquity had to be acquired during the earlier stages and is possibly related to the non-adiabatic s8 resonance crossing discussed in Section 3.2. If, on the other hand, αJ (cid:39) 2.92 − 2.94 arcsec yr−1, Jupiter would owe its present obliquity to the proximity between αJ and s7. This would imply that the obliquity excitation during the first stage of planetary migration must have been minimal. Figures 3 and 5 express the joint constraint on the planetary migration also for the intermediate cases, where the present obliquity of Jupiter arose as a combination of both effects discussed here. Figure 6 illustrates the two limiting cases discussed above. In panel (a), we assumed that the parameters during the first migration stage were such that the obliquity of Jupiter was excited to its current value during the s8 crossing (such as shown in Fig. 4). Also, we set αJ = 2.77 arcsec yr−1, corresponding to the best theoretical value of Helled et al. (2011). The Cassini state C2 corresponding to the s7 term is only slightly displaced from the center of the plot, and does not significantly contribute to the present obliquity value. In panel (b) of Fig. 6, we set αJ = 2.93 arcsec yr−1. This value implies that εC2 (cid:39) 2.6◦. The present obliquity of Jupiter would then be in large part due to the "forced" term arising from the proximity of s7. The initial excitation of Jupiter's obliquity during the first stage would have to be minor in this case. 3.3.2. Saturn In the case of Saturn, all action is expected to take place during the second stage of plan- etary migration. Insights gleaned from the numerical simulations discussed in Section 3.1 show that the s8 frequency should have very slowly approached αS, thus providing a concep- tual basis for capture of Saturn's spin vector in a resonance with s8 (e.g., Ward & Hamilton 2004; Hamilton & Ward 2004; Bou´e et al. 2009). To study this possibility, we assume that the I68 amplitude was excited to its current value during the instability, and remained nearly constant during the second stage of planetary migration. This choice is motivated by the NM12 simulations, where the inclination of Neptune is never too large. Note that Bou´e et al. (2009) investigated the opposite case where Neptune's inclination was substantially -- 16 -- excited during the instability and remained high when the resonance with s8 occurred. This type of strong inclination excitation does not happen in the NM12 models. (2009), the best-modeling values of αS from Helled et al. Here we assume that the planetary migration was very slow during the second stage and parametrize s8(t) as s8(t) = s8(0) + [s8 − s8(0)] exp(−t/τ ) with τ ≥ 80 Myr. The initial frequency value at the beginning of stage 2, s8(0), is estimated from the NM12 simulations. We find that s8(0) (cid:39) −1.3 arcsec yr−1, and use this value to set up the evolution of s8(t). We also assume a range of αS values. This has the following significance. As already pointed out by Bou´e et al. (2009) are not compatible with a resonant location of Saturn's spin axis. This is because the Cassini state C2 would be moved to a significantly larger obliquity value (≥ 34◦). So these values of αS would imply that Saturn's spin circulates about the Cassini state C1. On the other hand, the significant obliquity of Saturn requires an increase when the s8 value was crossing αS value, as schematically shown in the left panel of Fig. 4 for Jupiter's obliquity during the phase 1. Bou´e et al. (2009) tested this scenario using numerical simulations and found it extremely unlikely: initial data of an insignificant measure have led this way to the current spin state of Saturn. Indeed, here we recover the same result with a less extensive set of numerical simulations. Given the arguments discussed above we therefore tend to believe that the precession constant of Saturn may be somewhat smaller than the one determined by Helled et al. (2009). For instance, R. A. Jacobson (personal communication) determined the Saturn precession from the Saturn's ring observations. The mean precession rate obtained by him is 0.725 arcsec yr−1 (formal uncertainty of about 6%). This value would indicate αS in the range between 0.769 arcsec yr−1 and 0.864 arcsec yr−1. Both Ward & Hamilton (2004) and Bou´e et al. (2009) report other observational constraints of Saturn's pole precession that have comparably large uncertainty. We therefore sampled a larger interval of the αS values to make sure that all interesting possibilities are accounted for. Our numerical simulations thus spanned a grid of two parameters: (i) αS discussed above, and (ii) τ , the e-folding timescale of the s8 frequency that slowly changes due to residual migration of Neptune and depletion of the outer disk. The amplitude related to the s8 term in the inclination vector ζ of Saturn is kept constant, namely I68 = 0.064◦. To keep number of tested free parameters low, we assumed initial orientation of Saturn's spin axis s to be near the pole of the invariable plane. Specifically, we set its obliquity to 0.1◦ in the reference frame of the s8-frequency Fourier term in ζ. To prevent fluke results, we sampled 36 values of longitude ϕ of the Saturn's pole in the same reference frame, incrementing it from 0◦ by 10◦. Each of the simulations covered a 1 Gyr timespan. We recorded Saturn's pole orientation during the last 150 Myr time interval. We specifically analysed if it passes -- 17 -- close to the current location of Saturn's pole, namely ε8 (cid:39) 27.4◦ and ϕ8 (cid:39) −31.4◦ in the s8-frequency reference frame (see Table 2 in Ward & Hamilton 2004). A numerical run was considered successful, if the simulated Saturn's pole passed through a box of ±0.2◦ in obliquity ε and ±3◦ in longitude ϕ around the planet's values (ε8, ϕ8) during the recorded 150 Myr time interval. Note that the libration period of Saturn's pole around Cassini state C2, if captured in the spin-orbit resonance, is ∼ (50 − 100) Myr, depending on the libration amplitude. This set our requirement for the timespan over which we monitored Saturn's pole position. Figure 7 shows the results from this suite of runs. The shaded region shows correlated αS-τ pairs that provided successful match to the Saturn's pole position. We note that no successful solutions were obtained for αS > 0.812 arcsec yr−1 and all successful solutions correspond to the capture in the resonance zone around the Cassini state C2. The absence of low-probability solutions in which Saturn's pole would circulate about the Cassini state C1 for larger αS may be related to the limited number of simulations performed. No solutions were also obtained for τ > 215 Myr. This is because for such long e-folding timescales the resonant capture process would be strictly adiabatic and the simulated spin would not meet the condition of at least 30◦ libration amplitude (see discussion in Ward & Hamilton 2004; Hamilton & Ward 2004). The area occupied by successful solutions splits into two branches for αS (cid:39) 0.78 arcsec yr−1. This is because the obliquity of the Cassini state C2 is εC2 (cid:39) 27.4◦ for the critical value of αS, and the solutions have the minimum libration amplitude of (cid:39) 31◦. Figure 8 shows two evolution paths of Saturn's spin vector obtained in two different simulations. As mentioned above, in both cases Saturn's spin state was captured into the spin-orbit resonance s8, and remained in the resonance for the full length of the simulation. The final states of both simulations are a good proxy for the Saturn's present spin state. These two cases differ from each other principally because the path in panel (b) shows li- brations with a larger amplitude than the path in panel (a). Note some of the solutions, such as (b) here, may attain a significant librations amplitude. This is because with the corresponding values of τ (cid:39) 100 Myr the evolution is not adiabatic and the librations am- plitude is excited immediately after capture. Therefore, the complicated evolution histories proposed in Hamilton & Ward (2004) may not be needed. 4. Conclusions We studied the behavior of Jupiter's and Saturn's obliquities in models of planetary instability and migration that were informed from NM12. Rather then investigating a few specific cases directly from NM12, we considered the general concept of a two-stage migration -- 18 -- from NM12, and studied a broad range of relevant parameter values. We found that, in general, the two stage migration provides the right framework for an adequate excitation of Jupiter's and Saturn's obliquities. Moreover, we found that certain conditions must be satisfied during the first and second stages of migration, if the final obliquity values are to match the present obliquities of these planets. Our results indicate that Jupiter spin axis could have been tilted either when (i) the s8 frequency swept over αJ during the first migration stage (that is before the instability happened), or when (ii) the s7 frequency approached αJ at the end of planetary migration. For (i) to work, the crossing of s8 must be fast, such that the capture into the resonance does not happen, but not too fast, such that some excitation is generated by the resonance crossing. To obtain full εJ = 3.1◦ during this stage, the rate of change of s8 during crossing, ds8/dt, must be smaller than 0.05 arcsec yr−1 Myr−1 (assuming that I58 < 0.05◦). Since the evolution of s8 mainly relates to the radial migration of Neptune and dispersal of the outer disk of planetesimals, this result implies that both these processes would need to occur relatively slowly. More specifically, parameters ds8/dt and I58 would have to have values along a diagonal line in the (ds8/dt, I58) plane with larger values of I58 requiring larger values of ds8/dt (see Fig. 3). Any model of planetary instability/migration can be tested against this constraint. The models where ds8/dt is too slow and/or I58 is too large, as specified in Fig. 3, can be ruled out, because Jupiter's obliquity would be excited too much by the s8 crossing. Not much excitation of εJ is expected during the s8 crossing if ds8/dt was relatively fast and/or if I58 was only a very small fraction of its current value. If that is the case, Jupiter's obliquity would probably need to be excited during the very last stages of migration by (ii). For that to work, Jupiter's precession constant αJ would need to be (cid:39) 2.95 arcsec yr−1 (assuming εJ = 0 initially), which is a value that is significantly larger than the one estimated by Helled et al. (2011). This means that Helled's model would need to be adjusted to fit within this picture. It is also possible, however, that Jupiter's present obliquity was contributed partly by (i) and partly by (ii). If so, Figs. 3 and 5 express the joint constraint on ds8/dt and I58 during the first stage, and αJ. As for Saturn, our results indicate that the capture into the spin-orbit resonance with s8 (Ward & Hamilton 2004; Hamilton & Ward 2004) is indeed possible during the late stages of planetary migration, assuming that the migration rate was slow enough. The exact constraint on the slowness of migration depends on I68, which in turn depends how much Neptune's inclination was excited by the instability and how long it remained elevated (Bou´e et al. 2009). Since in the NM12 models, Neptune's orbital inclination is never large, we have good reasons to believe that I68 was comparable to its current value when the crossing -- 19 -- of s8 occurred. Thus, using I68 (cid:39) 0.064◦, we find that the e-folding migration timescale τ would need to be τ (cid:38) 100 Myr. If τ > 200 Myr, however, the capture in the s8 resonance would be strictly adiabatic. This would imply, if εS was negligible before capture, that the resonant state should have a very small libration amplitude (see Ward & Hamilton 2004; Hamilton & Ward 2004). It would then be difficult to explain the current orientation of Saturn's spin axis, which indicates that the libration amplitude should be at least 30◦. A more satisfactory solution, however, can be obtained for τ (cid:39) 100 − 150 Myr, in which case the capture into the resonance was not strictly adiabatic. In this case the ≥ 30◦ libration amplitude is obtained during capture. While the capture conditions pose a strong constraint on the timescale of Neptune's ra- dial migration, as discussed above, and additional constraint on Saturn's precession constant αS derives from the present obliquity of Saturn. This is because, again assuming that Saturn spin vector is in the resonance with s8 today, the present obliquity of Saturn implies that the Cassini state C2 of this resonance would have to be located at ε ∼ (28 − 30)◦. This would require that αS (cid:39) 0.78−0.80 arcsec yr−1. The value derived by Helled et al. (2009) is larger, (cid:39) 0.8445 arcsec yr−1, and clearly incompatible with this assumption. Direct measurements of the mean precession rate of Saturn's spin axis suggest that αS (cid:39) 0.81 arcsec yr−1, which is still slightly larger than the range given above, but the uncertainty interval of this estimate includes values below 0.8 arcsec yr−1 (R. A. Jacobson, personal communication). Figuring out the exact value of Saturn's precession constant will therefore be important. Once αS is known, Fig. 7 could be used to precisely constrain the timescale of planetary migration. We are grateful to Robert Jacobson for sharing his latest estimate of the precession rate of Saturn's spin axis. We also thank Tristan Guillot and Alessandro Morbidelli for discussions and pointing to us the future capability of Juno mission to constrain Jupiter's precession constant. The work of DV was supported by the Czech Science Foundation (grant GA13-01308S). The work of DN was supported by NASA's Origin of Solar Systems (OSS) program. REFERENCES Applegate, J. H., Douglas, M. R., Gursel, Y., Sussman, G. J., & Wisdom, J. 1986, AJ, 92, 176 Batygin, K., Brown, M. E., & Betts, H. 2012, ApJ, 744, L3 Brasser, R., & Lee, M. H. 2015, AJ, submitted -- 20 -- Breiter, S., Nesvorn´y, D., & Vokrouhlick´y, D. 2005, AJ, 130, 1267 Brouwer, D., & Clemence, D. M. 1961, Methods of Celestial Mechanics, Academic Press, New York Bou´e, G., Laskar, J., & Kuchynka, P. 2009, ApJ, 702, L19 Colombo, G. 1966, AJ, 71, 891 Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005, Nature, 435, 466 Hamilton, D. P., & Ward, W. R. 2004, AJ, 128, 2510 Helled, R., Schubert, G., & Anderson, J. D. 2009, Icarus, 199, 368 Helled, R., Anderson, J. D., Schubert, G., & Stevenson, D. J. 2011, Icarus, 216, 440 Henrard, J., & Murigande, C. 1987, Celestial Mechanics, 40, 345 Laskar, J. 1988, A&A, 198, 341 Laskar, J., & Robutel, P. 1993, Nature, 361, 608 Le Maistre, S., Folkner, W. M., & Jacobson, R. A. 2014, presented at 46th DPS meeting, abstract 422.08 Levison, H. F., Morbidelli, A., Van Laerhoven, C., Gomes, R., & Tsiganis, K. 2008, Icarus, 196, 258 Lissauer, J. J., & Safronov, V. S. 1991, Icarus, 93, 288 Malhotra, R. 1995, AJ, 110, 420 Masset, F., & Snellgrove, M. 2001, MNRAS, 320, L55 McNeil, D. S., & Lee, M. H. 2010, Bull. Am. Astron. Soc., 42, 948 Minton, D. A., & Malhotra, R. 2009, Nature, 457, 1109 Minton, D. A., & Malhotra, R. 2011, ApJ, 732, 53 Morbidelli, A., & Crida, S. 2007, Icarus, 191, 158 Morbidelli, A., Levison, H. F., Tsiganis, K., & Gomes, R. 2005, Nature, 435, 462 Morbidelli, A., Brasser, R., Tsiganis, K., Gomes, R., & Levison, H. F. 2009, A&A, 507, 1041 -- 21 -- Morbidelli, A., Brasser, R., Gomes, R., Levison, H. F., & Tsiganis, K. 2010, AJ, 140, 1391 Morbidelli, A., Tsiganis, K., Batygin, K., Crida, A., & Gomes, R. 2012, Icarus, 219, 737 Morbidelli, A., Gaspar, H. S., & Nesvorn´y, D. 2014, Icarus, 232, 81 Nesvorn´y, D. 2011, ApJ, 742, L22 Nesvorn´y, D. 2015, AJ, submitted Nesvorn´y, D., & Vokrouhlick´y, D. 2009, AJ, 137, 5003 Nesvorn´y, D., & Morbidelli, A. 2012, AJ, 144, 117 Nesvorn´y, D., Vokrouhlick´y, D., & Morbidelli, A. 2007, AJ, 133, 1962 Nesvorn´y, D., Vokrouhlick´y, D., & Morbidelli, A. 2014a, ApJ, 768, 45 Nesvorn´y, D., Vokrouhlick´y, D., & Deienno, R. 2014b, ApJ, 784, 22 Pierens, A., & Nelson, R. P. 2008, A&A, 482, 333 Thommes, E. W., Duncan, M. J., & Levison, H. F. 1999, Nature, 402, 635 Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Nature, 435, 459 Ward, W. R., & Hamilton, D. P. 2004, AJ, 128, 2501 Ward, W. R., & Canup, R. M. 2006, ApJ, 640, L91 This preprint was prepared with the AAS LATEX macros v5.2. -- 22 -- Fig. 1. -- Top panel: Parametric dependence of Cassini state C1, C2 and C4 obliquity (ordinate) on the frequency-ratio parameter κ (abscissa) in the Colombo model (the C3 equilibirium has obliquity larger than 90◦ and it is not relevant for our discussion). The orbital inclination I has been set to 0.5◦ value. The dashed line indicates the critical value −κ(cid:63) at which C1 and C4 bifurcate (Eq. 10). The gray area for κ > κ(cid:63) shows maximum width of the resonant zone around C2. Bottom panels: Examples of phase portraits H(X(cid:48), ϕ) = C for two values of κ: (a) κ = −0.4 on the left, and (b) κ = −0.6 on the right. We use coordinates x = sin ε cos ϕ and y = sin ε sin ϕ with the origin at the north pole ε = 0. The gray symbols show location of the Cassini equilibria and the curves are isolines H(X(cid:48), ϕ) = C for suitably chosen C values. The bold line in (b) is the separatrix of the resonant zone around C2 stationary solution. -- 23 -- Fig. 2. -- An example of planetary migration and instability from Nesvorn´y & Morbidelli (2012). The plot shows the evolution of the semimajor axis (bold line), and the perihelion and aphelion distances (thin lines) of the giant planets. The initial orbits of Jupiter, Saturn and the inner ice giant were placed in the 3:2 resonant chain. The semimajor axes of the two outer ice giants were set to be 16 au and 22 au. The trans-Neptunian disk of planetesimals (not shown here) was resolved by 10,000 equal-mass particles. The disk originally extended from 23.5 au to 29 au and had the total mass of 20 MEarth. The instability happened at t (cid:39) 5.6 Myr after the start of the simulation. The third ice giant was subsequently ejected from the Solar system. Note that the migration rate before the instability (phase 1) is significantly larger than the migration rate after the instability (phase 2). -- 24 -- Fig. 3. -- The final obliquity εfin of Jupiter obtained in our integrations of crossing of the s8 spin-orbit resonance. Jupiter's obliquity is shown as a function of two parameters: (i) the amplitude I58 of the Fourier term in Jupiter's ζ variable (see Section 2), and (ii) the rate ds8/dt at the time when s8 became equal to αJ. The gray shading indicates the final obliquity (the scale bar on the right shows the corresponding numerical value in degrees). The three bold isolines correspond to εJ = 1◦, 2◦ and 3◦ (see the labels). -- 25 -- Fig. 4. -- An example demonstrating the effect of s8 frequency sweeping over αJ. Here we set I58 = 0.02◦ and ds8/dt = 0.01 arcsec yr−1 Myr−1. The left panel shows Jupiter's obliquity as a function of time. Obliquity εJ increases to (cid:39) 2.7◦ during the resonance crossing. The width of the Cassini resonance is small because I58 is small, and the assumed rate ds8/dt is too large in this case to lead to capture. The right panel shows the spin axis evolution projected onto the (x, y) plane, where x = sin ε cos ϕ and y = sin ε sin ϕ. The Cassini state C2 drifts along the x-axis during the integration (as indicated by the gray arrow), and reaches very large obliquity values. -- 26 -- Fig. 5. -- Final obliquity of Jupiter resulting from an adiabatic approach of the s7-frequency towards αJ (the obliquity has been computed in the reference frame associated with this frequency term in ζ). While today's value of s7 is known fairly accurately, αJ has a significant uncertainty. We therefore treat αJ as a free parameter. The initial obliquity of Jupiter is also treated as a free parameter, because it depends on the effects during the first migration stage (see Fig. 3). The key to the shading scale is provided by the vertical bar on the right (white region corresponds to the final maximum obliquity smaller than 2.5◦, incompatible with Jupiter's value). The bold curve corresponds to the 3.45◦ isoline, estimated obliquity value today in this frame (e.g., Ward & Canup 2006). The arrows indicate parametric location of the two examples shown on Fig. 6. The upper panel shows the obliquity value of the Cassini state C2 as a function of αJ. -- 27 -- Fig. 6. -- Two examples of Jupiter's spin state evolution during the 1 Gyr-long time interval after the planetary instability. The planet pole is shown in the Cartesian coordinates x = sin ε cos ϕ and y = sin ε sin ϕ. The arrow shows evolution of the Cassini state C2 over the interval of time covered by the simulation. The gray star shows the current location of Jupiter's pole in these coordinates. Gray dots show the pole position output every 5 kyr during the simulation, the black circles highlight the first and the last 30 Myr of the evolution. Two different parametric combination along the bold curve in Fig. 5 were chosen: (i) αJ = 2.77 arcsec yr−1 and εini = 3.1◦ in the left panel (a), and (ii) αJ = 2.93 arcsec yr−1 and εini = 1.3◦ in the right panel (b). -- 28 -- Fig. 7. -- Bottom panel: Distribution of solutions successfully matching Saturn's spin state in the parametric plane αS (abscissa) vs τ (ordinate), the e-folding time of s8-frequency evolution during the phase 2. No successful solutions were obtained in the white region of the plot. The darker the gray-scale in the given bin, the more robust the solution is. The maximum value 36 corresponds to 36 sampled initial conditions of the simulations for each (αS, τ ) pair (see the side bar). When αS (cid:39) 0.78 arcsec yr−1, the obliquity of the corresponding to Cassini state C2 is (cid:39) 27.4◦. Therefore the capture solution corresponds to the minimum needed libration amplitude of Saturn's spin in the s8 reference frame. From this value the larger libration-amplitude solutions bifurcate towards smaller/larger αS values. The arrows indicate parametric location of two particular examples shown on panels (a) and (b) of Fig. 8. Top panel: Obliquity of the Cassini state C2 (solid line) and maximum width of the associated resonant zone (gray area) as a function of αS. We assume inclination I = I68 = 0.064◦ and terminal value of the orbital precession frequency s8 (cid:39) −0.692 arcsec yr−1. -- 29 -- Fig. 8. -- Two examples Saturn's obliquity excitation by the capture and evolution in the resonance with s8. The spin axis s is projected onto the (x, y) plane, where x = sin ε cos ϕ and y = sin ε sin ϕ. The reference frame used here is the one associated with the s8 frequency term contributing to ζ of Saturn. The gray arrow shows the evolution of the Cassini state C2 over the full length of the simulation (1 Gyr). In panel (a), we assumed that αS = 0.785 arcsec yr−1, which means that the terminal obliquity of C2 is at (cid:39) 28.2◦, and τ = 150 Myr. In (b), we used αS = 0.8 arcsec yr−1, which means that the terminal obliquity of C2 is at (cid:39) 30.3◦, and τ = 135 My. These combinations were also highlighted by arrows on Fig. 7. The star in each panel shows the present orientation of the Saturn pole vector in the (x, y) coordinates (e.g., Ward & Hamilton 2004, Table 2).
1712.03746
1
1712
2017-12-11T12:24:14
Simultaneous, Multi-Wavelength Variability Characterization of the Free-Floating Planetary Mass Object PSO J318.5-22
[ "astro-ph.EP" ]
We present simultaneous HST WFC3 + Spitzer IRAC variability monitoring for the highly-variable young ($\sim$20 Myr) planetary-mass object PSO J318.5-22. Our simultaneous HST + Spitzer observations covered $\sim$2 rotation periods with Spitzer and most of a rotation period with HST. We derive a period of 8.6$\pm$0.1 hours from the Spitzer lightcurve. Combining this period with the measured $v sin i$ for this object, we find an inclination of 56.2$\pm 8.1^{\circ}$. We measure peak-to-trough variability amplitudes of 3.4$\pm$0.1$\%$ for Spitzer Channel 2 and 4.4 - 5.8$\%$ (typical 68$\%$ confidence errors of $\sim$0.3$\%$) in the near-IR bands (1.07-1.67 $\mu$m) covered by the WFC3 G141 prism -- the mid-IR variability amplitude for PSO J318.5-22 one of the highest variability amplitudes measured in the mid-IR for any brown dwarf or planetary mass object. Additionally, we detect phase offsets ranging from 200--210$^{\circ}$ (typical error of $\sim$4$^{\circ}$) between synthesized near-IR lightcurves and the Spitzer mid-IR lightcurve, likely indicating depth-dependent longitudinal atmospheric structure in this atmosphere. The detection of similar variability amplitudes in wide spectral bands relative to absorption features suggests that the driver of the variability may be inhomogeneous clouds (perhaps a patchy haze layer over thick clouds), as opposed to hot spots or compositional inhomogeneities at the top-of-atmosphere level.
astro-ph.EP
astro-ph
Draft version December 12, 2017 Typeset using LATEX modern style in AASTeX61 SIMULTANEOUS, MULTI-WAVELENGTH VARIABILITY CHARACTERIZATION OF THE FREE-FLOATING PLANETARY-MASS OBJECT PSO J318.5-22 Beth A. Biller,1, 2 Johanna Vos,1, 2 Esther Buenzli,3 Katelyn Allers,4 Mickael Bonnefoy,5 Benjamin Charnay,6 Bruno B´ezard,6 France Allard,7 Derek Homeier,8 Mariangela Bonavita,1, 2 Wolfgang Brandner,9 Ian Crossfield,10 Trent Dupuy,11 Thomas Henning,9 Taisiya Kopytova,12 Michael C. Liu,13 Elena Manjavacas,14 and Joshua Schlieder15 1SUPA, Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK 2Centre for Exoplanet Science, University of Edinburgh, Edinburgh, UK 3Institute for Particle Physics and Astrophysics, ETH Zurich, Wolfgang-Pauli-Strasse 27, 8093 Zurich, Switzerland 4Bucknell University; Department of Physics and Astronomy; Lewisburg, PA 17837, USA 5Institut de Plan´etologie et d'Astrophysique de Grenoble, Universit´e Grenoble Alpes, CS 40700, 38058 Grenoble C´edex 9, France 6LESIA, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Universit´es, UPMC Univ. Paris 6, Universit´e Paris-Diderot, Sorbonne Paris Cit´e, 5 place Jules Janssen, 92195 Meudon, France 7Univ Lyon, ENS de Lyon, Univ Lyon 1, CNRS, Centre de Recherche Astrophysique de Lyon UMR5574, F-69007 Lyon, France 8Zentrum fur Astronomie der Universitat Heidelberg, Landessternwarte, Konigstuhl 12, D-69117 Heidelberg, Germany 9Max-Planck-Institut fur Astronomie, K"onigstuhl 17, D-69117 Heidelberg, Germany 10Department of Physics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA 11Gemini Observatory, 670 N. Aohoku Pl., Hilo, HI 96720, USA 12School of Earth & Space Exploration, Arizona State University, Tempe AZ 85287, USA 13Institute for Astronomy, University of Hawaii at Manoa, 2680 Woodlawn Drive, Honolulu, HI, 96822, USA 14Steward Observatory, University of Arizona, 933 N. Cherry Ave, Tucson, AZ 85917, USA 15Exoplanets and Stellar Astrophysics Laboratory, Code 667, NASA Goddard Space Flight Center, Greenbelt, MD, USA (Accepted 10 December 2017) Corresponding author: Beth Biller [email protected] 2 Biller et al. Submitted to AJ ABSTRACT We present simultaneous HST WFC3 + Spitzer IRAC variability monitoring for the highly-variable young (∼20 Myr) planetary-mass object PSO J318.5-22. Our simultaneous HST + Spitzer observations covered ∼2 rotation periods with Spitzer and most of a rotation period with HST. We derive a period of 8.6±0.1 hours from the Spitzer lightcurve. Combining this period with the measured vsini for this object, we find an inclination of 56.2±8.1◦. We measure peak-to-trough variability amplitudes of 3.4±0.1% for Spitzer Channel 2 and 4.4 - 5.8% (typical 68% confidence errors of ∼0.3%) in the near-IR bands (1.07-1.67 µm) covered by the WFC3 G141 prism – the mid-IR variability amplitude for PSO J318.5-22 one of the highest variability amplitudes measured in the mid-IR for any brown dwarf or planetary mass object. Additionally, we detect phase offsets ranging from 200–210◦ (typical error of ∼4◦) between synthesized near-IR lightcurves and the Spitzer mid-IR lightcurve, likely indicating depth-dependent longitudinal atmospheric structure in this atmosphere. The detection of similar variability amplitudes in wide spectral bands relative to absorption features suggests that the driver of the variability may be inhomogeneous clouds (perhaps a patchy haze layer over thick clouds), as opposed to hot spots or compositional inhomogeneities at the top-of-atmosphere level. Keywords: planets and satellites: gaseous planets, planets and satel- lites: atmospheres,(stars:) brown dwarfs AASTEX variability characterization of PSO J318.5-22 3 1. INTRODUCTION Rotationally-modulated variability is a key probe of exoplanet and brown dwarf at- mospheric structure. Field brown dwarfs are generally rapid rotators, with periods of 3-20 hours (Zapatero Osorio et al. 2006). Several young planetary mass objects now have measured rotation periods <20 hours as well (Snellen et al. 2014; Biller et al. 2015; Lew et al. 2016; Zhou et al. 2016; Vos et al. 2017b). Any top-of-atmosphere inhomogeneity on rapid rotators (due e.g. to cloud structure, thermo-chemical in- stabilities, or auroral emission) may be detectable via the quasiperiodic photometric variability it produces. Variability is common in high surface gravity field brown dwarfs, with the maximum variability amplitude in the near-IR at the L/T spec- tral type transition (Radigan et al. 2014; Radigan 2014; Wilson et al. 2014; Metchev et al. 2015). Most efforts to model the mechanism driving the observed quasiperi- odic variability have utilized patchy thin and thick cloud cover (Apai et al. 2013), possibly related to the dissipation of clouds across the L/T transition. Variability studies enable a multi-dimensional view of these atmospheres – lightcurves at shorter wavelengths generally probe deeper atmospheric pressure levels compared to longer wavelengths (Buenzli et al. 2012; Biller et al. 2013a; Yang et al. 2016) and the rapid rotation of these objects move different parts of their atmospheres in and out of view (Apai et al. 2013). Due to the extreme contrast difference between host star and planet, very few exoplanet companions are amenable to high-precision variability searches. However, a wider class of "exoplanet analogues" can be defined with similar spectral types and surface gravities as young exoplanet companions. Several young planets and exoplanet analogues also have measured periods ≤20 hours (Snellen et al. 2014; Biller et al. 2015; Allers et al. 2016; Zhou et al. 2016), allowing variability searches with reasonable observation lengths from both ground and space. Known exoplanet analogues include bonafide free-floating planetary mass objects such as the 8.3±0.5 MJup VL-G L7±1 β Pic moving group member PSO J318.5-22 (Liu et al. 2013; Allers et al. 2016) as well as slightly higher mass (<25 MJup), young objects such as WISEP J004701.06+680352.1 (henceforth W0047), a ∼20 MJup, very red AB Dor (150 Myr) moving group member (Gizis et al. 2012, 2015; Filippazzo et al. 2015; Vos et al. 2017b). While they likely do not share a formation mechanism with exoplanet companions, they share similar masses and surface gravities (Faherty et al. 2016). Thus, they allow us to study similar atmospheres without first overcoming the light of a nearby star. In particular, PSO J318.5-22 and W0047 have spectra that are nearly identical to the inner two HR 8799 planets (Bonnefoy et al. 2016). The low surface gravity in young objects significantly affects the spectra of these objects and also the Teff at which these objects would transition between the L and T spectral types. Low surface gravity objects have redder colors compared to field dwarfs and potentially retain thick silicate clouds or small-scale turbulent energy transport via thermo-chemical instabilities down to 4 Biller et al. lower Teff than field counterparts with similar spectral types (Barman et al. 2011; Liu et al. 2016; Tremblin et al. 2016). Variability has recently been detected for the first time in young planetary mass objects. Vos et al. (in prep) surveyed ∼40 low surface gravity planetary mass ob- jects and very young brown dwarfs with L and T spectral types with NTT SofI and UKIRT WFCAM (sensitive to variability amplitudes >2%), detecting low-amplitude variability in 3 L1-L4 objects and high-amplitude variability (7-10%) in the mid-to- late L, low surface gravity objects PSO J318.5-22 (Biller et al. 2015) and 2MASS J2244316+204343 (henceforth 2M2244, Vos et al. 2017b), a ∼20 MJup, AB Dor (150 Myr) moving group member (Filippazzo et al. 2015; Vos et al. 2017b). Zhou et al. (2016) report lower-amplitude variability in the mid-L companion planetary mass ob- ject 2MASSW J1207334-393254b (henceforth 2MASS 1207b) and Lew et al. (2016) found high amplitude variability (8%) for WISE 0047. For young, T-spectral type objects, two detections have been reported to date. Recently, Gagn´e et al. (2017) identified the highly variable T2.5 dwarf SIMP 0136 (Artigau et al. 2009) as a likely member of the 200 Myr Carina-Near moving group. Naud et al. (2017) report a tentative J-band detection of variability for the wide T3.5 planetary mass companion GU Psc b, with 4±1% variability on a six-hour timescale detected in one epoch out of three of monitoring. The mid-to-late M, low-surface gravity objects PSO J318.5-22, WISE 0047, and 2M2244 have the highest near-IR amplitudes (>5%) measured for any L type object to date. Until recently few late L low surface gravity objects had been identified and only four such objects with spectral types between L6.5 and L9 have been monitored for variability in the near-IR (Morales-Calder´on et al. 2006; Lew et al. 2016; Biller et al. 2015; Vos et al. 2017b, Vos et al. in prep). From model predictions, these late- L objects are expected to have thick (and probably homogeneous) cloud cover (e.g. Madhusudhan et al. 2011), although some recent modeling efforts have posited that the red colors of late-L dwarfs could alternatively be produced as the result of con- vection driven by thermo-chemical instabilities (Tremblin et al. 2016, 2017). Late-L low surface gravity objects already clearly demonstrate differences in variability prop- erties compared to field brown dwarfs, which peak in variability amplitude at later spectral types (T0-T2) along the L/T spectral type transition (Radigan et al. 2014; Radigan 2014), potentially due to breakup or patchiness of clouds at this spectral type transition. These differences are likely a result of the lower surface gravities for the young, planetary mass late-L objects compared to field brown dwarfs. Studying the variability of young objects in detail will illuminate the role of surface gravity in determining atmospheric structure. Simultaneous multi-wavelength variability monitoring is a powerful tool to under- stand the atmospheres of these objects, allowing us to pinpoint the mechanism driv- ing the variability. Different wavelength regimes probe different atmospheric depths (Marley et al. 2012) and high spectroscopic resolution allows us to study variabil- AASTEX variability characterization of PSO J318.5-22 5 ity within individual spectral features. Morley et al. (2014) find that variability due to patchy clouds should drive high amplitude variability within wide spectral windows while variability due to hot spots (i.e. heating at a specific pressure level) should drive larger variability within absorption features. Based on this, Morley et al. (2014) suggest that simultaneous, multi-wavelength observations probing both inside and outside molecular absorption features will prove to be particularly valuable in understanding the physical processes driving this variability. Only ∼10 old field dwarfs have published HST spectroscopic variability monitoring (Buenzli et al. 2015b,a; Apai et al. 2013; Buenzli et al. 2012; Yang et al. 2015, 2016). SIMP 0136 (Apai et al. 2013) and W0047 (Lew et al. 2016) are the only potentially young objects with published HST spectroscopic variability monitoring. HST enables exceptionally high photometric precision as well as access to the 1.4 and 1.1 µm water absorption bands, which are unobservable from the ground due to telluric absorption. For L/T transition objects such as Luhman 16B, SIMP 0136, and 2M 2139 (Buenzli et al. 2015b; Apai et al. 2013), these studies have found correlated variability across the J and H band, with decreased variability in the 1.4 µm water absorption feature – consistent with variability due to inhomogeneous thick and thin cloud cover. For the T6.5 2M 2228, Buenzli et al. (2012) found significant phase shifts at different wave- lengths, including between broadband J and the 1.4 µm water feature – interpreted as differences in cloud properties at different atmospheric levels. For the mid-L dwarfs, 2M 1821 and 2M 1507, Yang et al. (2015) found correlated variability with similar amplitude in the 1.4 µm water band as in broadband J and H – this is interpreted as variability due to high level hazes (above significant water concentrations) in these atmospheres. Here we present simultaneous HST WFC3 + Spitzer IRAC variability monitoring for the variable planetary-mass object PSO J318.5-22. Of the current ensemble of free-floating young objects with estimated masses <30 MJup , PSO J318.5-22 (Liu et al. 2013) is the closest analogue in properties to imaged exoplanet companions. Gagn´e et al. (2014) and Liu et al. (2013) identified it as a β Pic moving group member and it possesses colors and magnitudes similar to the HR 8799 planets and 2M1207b (Bonnefoy et al. 2016). Using evolutionary models and adopting an age of 23±3 Myr, Allers et al. (2016) find an effective temperature of Tef f = 1127+24−26 K and a mass estimate of 8.3±0.5 MJup for PSO J318.5-22. Understanding the variability of this benchmark object will yield fundamental insights into its atmospheric properties, especially regarding the presence of clouds – and by proxy the expected properties of exoplanet companion atmospheres. 2. OBSERVATIONS Simultaneous HST + Spitzer observations of PSO J318.5-22 were acquired on 8-9 September 2016. Spitzer observations lasted from UTC 2016-09-08 09:01:14 to UTC 2016-09-09 02:18:27, with 5 HST orbits taken from UTC 2016-09-08 11:38:59 to UTC Biller et al. 6 2016-09-08 18:44:41, for a simultaneous monitoring period of ∼7 hours, and a Spitzer monitoring period of ∼17.2 hours. Spitzer observations were taken with IRAC in Channel 2 (4.5 µm) in staring mode, with 1940×30 s frames acquired (program id 12002). As Spitzer requires ∼30 minutes to settle after a target is acquired, a short dithered sky sequence (9×30 s frames taken at 5 dither positions) preceded the science sequence. A short sky sequence (1×30 s frame taken at 5 dither positions) was acquired after the science sequence as well. Following established procedures to ensure optimal photometric precision and correct for intrapixel sensitivity variations (Mighell et al. 2008), care was taken to place the target in the IRAC "sweet spot" during the science sequence, which lies in the upper left quadrant of the full detector. HST observations were taken with the infrared channel of WFC3 with the G141 grism (program ID 14188). The full 123×136 arcsec frame was used with the SPARS25 readout mode, enabling observation of 6 background stars as well as the target. Each 90-minute orbit yielded 59 minutes of usable exposure time, when the target was not occulted by the Earth. To determine object positions on the detector for each orbit, a direct image was taken in the F127M filter with exposure time of 53 s (NSAMP=3). Thereafter, a sequence of 9×278 s (NSAMP=12) exposures were taken with the G141 grism, which covers a wavelength range of 1.077 µm to 1.7 µm, with resolution R=130 at 1.4 µm. With the remaining orbital visibility, an additional 53 s (NSAMP=3) grism exposure was taken at the end of orbit 1 and an additional 153 s grism exposure (NSAMP=7) was taken in orbits 2-5. These final, shorter exposures were significantly noisier than the other exposures and were omitted from the final analysis. One 278 s exposure in orbit 4 suffered complete data loss, as the data were not fully read off the HST recorder before being overwritten, and is thus omitted in the following analysis. 3. SPITZER DATA REDUCTION AND LIGHTCURVE EXTRACTION No Spitzer [4.5 µ] magnitude has previously been reported in the literature for PSO J318.5-22. From the full-sequence Spitzer MOSAIC images, we derived a Spitzer [4.5 µ] magnitude of 12.541±0.017, using the code described in Dupuy & Kraus (2013). This is in good agreement with the WISE W2 magnitude for this object (Liu et al. 2013). To construct a lightcurve as a function of time, we extracted photometry from the corrected Basic Calibrated Data images from the Spitzer Science Center, processed with IRAC pipeline version 19.2.0. Times for each photometric point were taken from the M JD OBS header keyword, which provides the Modified Julian Date. After find- ing centroids for the target and a number of reference stars using box centroider.pro, we performed aperture photometry about these centroids. A range of apertures were tested – we adopt here an aperture of 2.4 pixels, which produced lightcurves with the lowest RMS. AASTEX variability characterization of PSO J318.5-22 7 To robustly remove outliers while avoiding subtracting out any astrophysical vari- ability, we followed the clipping procedure described in detail in Heinze et al. (2014). We median-smoothed each lightcurve with a sliding boxcar (width of 25 frames, cor- responding to 12.5 minutes). The smoothed lightcurve is subtracted from the original data, removing astrophysical and systematic signals with timescales longer than 12 minutes. Thus, any outliers remaining in the subtracted lightcurve must be artifacts and can be confidently removed, using a 6-σ clip. The flux of an object on a given point of the IRAC detector will vary depending on exactly where a point source falls with respect to the centre of a pixel – this is known as the 'pixel phase effect'. We correct for the pixel phase effect using the pixel phase correct gauss.pro routine from the Spitzer IRAC website, which models the pixel phase response as a double-gaussian, a summation of gaussians in the or- thogonal pixel directions. The pixel phase corrected flux is then binned into 2.5 min bins. 4. HST DATA REDUCTION AND LIGHTCURVE EXTRACTION We extracted spectra for PSO J318.5-22 and 6 background stars on the detector from FLT calibrated individual exposures downloaded from the MAST archive. Times for each exposure were taken from the mean of the EXP ST ART and EXP EN D header keywords, which are provided as Modified Julian Date. The FLT files have been processed using calwf c3, which performs basic data calibration including bad pixel flagging. We then corrected for bad pixels flagged by calwf c3, specifically pixels flagged with flag values 1, 4, 32, and 512. Bad pixel correction was performed by interpolating over the pixels on the left and the right of the flagged pixel. If the right-side pixel was also flagged, then only the left pixel is used to correct the original flagged pixel. We visually identified one bad pixel in the spectrum of PSO J318.5-22 that was not automatically flagged by calwf c3. This pixel was corrected manually. For the direct images taken at the beginning of each orbit, object positions on the chip were obtained using Source Extractor (SExtractor). For each object on the detector, the aXe pipeline was then used to extract slitless spectroscopy from each of the 9×278 s grism exposures per orbit, using the FLT grism files and the object positions obtained with SExtractor as inputs. Sky subtraction was performed using the aXeprep routine and spectrum extraction was performed using the aXecore routine. The usable spectral bandwidth runs from 1.07 to 1.67 µm, with a resolution of R=130. We extracted spectra with extraction widths ranging from 1 - 20 pixels. We found that a 7 pixel extraction width best balanced object signal against background noise. Thus, we adopt the 7-pixel extraction width for subsequent analysis. We flux-calibrated the extracted spectra using the G141 sensitivity curve. We extracted lightcurves from the spectra across a variety of spectral bandwidths. We integrated over the full 1.07-1.67 µm spectral range to generate a "white-light" lightcurve. In order to compare with ground-based studies, we also integrated over 8 Biller et al. the standard 2MASS J and H bandpasses. Note that the 2MASS H band extends to wavelengths longer than the 1.67 µm cutoff for the G141 grism, thus we have only integrated over the portion of the 2MASS H filter that falls within the G141 grism spectral range. We consider variability as well in two spectral features – integrating from 1.34 to 1.44 µm to capture variability in the 1.4 µm water absorption feature and from 1.60 to 1.67 µm to capture variability in the 1.6 µm methane absorption feature. As noted by previous studies (Buenzli et al. 2012; Apai et al. 2013; Buenzli et al. 2015b), WFC3 photometry displays a "ramp effect" – where the flux appears to in- crease with an exponential ramp at the beginning of each orbit. This is especially notable in the first orbit of a visit. Buenzli et al. (2012) find this effect to be inde- pendent of count rate and wavelength. Previous authors who have utilized a 256x256 pixel subarray instead of the 1048x1048 full frame have corrected this effect by us- ing an analytic function derived from a non-variable source in the field (Apai et al. 2013). Because we have 6 background stars in the field, we choose to build a correc- tion based solely on these background stars, without fitting an analytic function or a detector-based model as described in Zhou et al. (2017). Of the 6 background stars, one is considerably fainter than other objects on the detector and another appears to be somewhat variable itself. For the remaining 4 well-behaved stars (all 2-3× as bright as the target object), we combine their normalized white light lightcurves to produce a calibration curve (utilizing median combination, then taking the average of the two central values, as we use an even number of reference stars). We then divide both target and background star lightcurves and spectra by the calibration curve to correct for the ramp effect as well as other systematics which affect all objects on the detector (this is similar to the approach taken in ground based studies such as Radigan et al. 2014; Biller et al. 2015). 5. RESULTS Spitzer and HST lightcurves (after correction for the ramp effect) for PSO J318.5-22 are presented in Fig. 1. To increase the S/N ratio, the lightcurves have been binned by a factor of 5 for Spitzer, resulting in a 2.5-minute cadence, and by a factor of 3 for HST, resulting in a 14-minute cadence. Small colored points are the 6 background stars after being detrended by the calibration curve; PSO J318.5-22 is clearly variable compared to the reference stars. The mean and median spectra across the full 5 orbit HST observation, as well as the median spectrum per orbit, are presented in the top panel of Fig. 2. Similar spectra for one of the well-behaved reference stars is shown in the bottom panel of Fig. 2. 5.1. Period and amplitude from the Spitzer light curve The unbinned Spitzer lightcurve (30 s cadence) along with the best fit sinusoid using a Levenberg-Marquardt least-squares minimization algorithm is presented in Fig. 3. The sinusoidal model has four parameters: period (in hours), phase (in degrees), mean AASTEX variability characterization of PSO J318.5-22 9 lightcurve value (since we have divided the raw lightcurve by the median flux over the whole observation, this should tend towards unity), and amplitude (in percent vari- ation, peak to mean lightcurve value). The best fit model with gaussian noise added is also shown and provides a good match to the observed lightcurve. We also plot the periodogram in Fig. 3 of PSO J318.5-22 as well as a number of reference stars in the field to identify periodic variability in our targets. The 1% false-alarm probability (FAP, plotted in blue on the figure) is calculated from 1000 simulated lightcurves. These lightcurves are produced by randomly permuting the indices of reference star lightcurves (Radigan et al. 2014), producing lightcurves with Gaussian-distributed noise. The observed variability is reasonably well modeled with a sinusoidal model, although successive maxima and minima appear to be marginally increasing in flux. Thus we considered as well sinusoidal + linear models. To fully explore the parameter space of both sinusoidal and sinusoidal + linear fits, we used the emcee Markov-Chain Monte Carlo package (Foreman-Mackey et al. 2013) to determine the full posterior probability distribution. We ran an MCMC chain using a χ2 likelihood function with 1000 walkers for 2000 steps. The first 100 steps of each chain were thrown out as part of the "burn-in". Chains were checked by eye for convergence. We fit both a single sinusoid model (results shown in Fig. 4) and a sinusoid + linear model (shown in Fig. 5). We adopt the 50% quantile value as the best value for amplitude and period. Best values of amplitude (peak-to-median value) and period as well as 68% confidence-interval and 95% confidence-interval errors are presented in Table 1. Both provide reasonably good fits – in Fig. 6, we plot 1000 samples from our chain for both models. We calculated the Bayesian Information Criterion (BIC, Schwarz 1978) for the adopted best value parameters for both the single sinusoid and sinusoid + linear fit. The BIC is given by: BIC = ln(n) k − 2 ln(L) (1) where n is the number of data points in the lightcurve (1700 for the Spitzer curve), k is the number of model parameters, and L is the maximized value of the likelihood function of the model. We find BIC(cid:39)12 for the single sinusoid and BIC(cid:39)19 for the sinusoid + linear fit. The model with the lower value of BIC is preferred, thus we adopt the sinusoid-only model for further analysis and comparison to HST results. Apai et al. (2017) recently modeled the rapid brightness evolution found in brown dwarf lightcurves using beating patterns between multiple planetary-scale wave sur- face features that move at slightly different velocities due to zonal wind speed vari- ations. They considered both a simple, 3-sinusoid model and applied as well their Aeolus mapping package, with both spots and planetary-scale waves (Karalidi et al. 2016). We attempted to fit a similar, 3-sinusoid model to our Spitzer lightcurves, with the periods of the three sinusoids given as: Prot = (P1 + P2)/2 (wavenumber k=1 waves) and P3 = Prot/2 (k=2 wave), where Prot is the rotational period of PSO J318.5-22. In this case, the best fitting model always reverted to a single sinusoid, 10 Biller et al. Table 1. MCMC fit results for Spitzer variability monitoring. Peak-to-median variability amplitudes are reported here. model parameter best value 68% sine amplitude 1.69% 0.03% 95% 0.07% 8.61 hours 0.06 hours 0.11 hours 8.63 hours 0.05 hours 0.10 hours period phase sine+slope amplitude period phase slope -10.5◦ 1.81% -8.4◦ 0.05% 2.4◦ 0.04% 2.2◦ 0.005% 4.8◦ 0.07% 4.3◦ 0.01% with negligible amplitudes for the other two sinusoids relative to the uncertainties in the lightcurve. We only cover ∼2 apparent rotational periods for PSO J318.5-22, thus it is unclear from our Spitzer lightcurve alone whether the nearly sinusoidal variation observed is the fundamental lightcurve of this object or whether we have observed it during a period when multiple planetary-scale wave surface features happen to be in phase. This is qualitatively similar to the case of SIMP 0136 described in Apai et al. (2017), where the apparent periods of the two k=1 waves are expected to vary only by ∼1%. However, Allers et al. (2016) find a maximum period for PSO J318.5-22 of 10.2 hours from vsini measurements, ruling out the possibility of a double-peaked lightcurve, thus we expect the 8.6±0.1 hour rotational period derived from our single sinusoid MCMC fits to be accurate in this case. 5.2. Inclination Allers et al. (2016) constrain the inclination of PSO J318.5-22 to >29◦, based on their measured vsini, reasonable estimates of the radius of PSO J318.5-22 from on evolutionary models, and a lower limit on the period of ∼5 hours from Biller et al. (2015). With our high-precision measurement of the period from Spitzer observations, we can now directly measure the inclination for this object. Using Monte Carlo meth- ods to account for uncertainties in vsini, radius, and period, we drew 30000 samples from the vsini distribution found by Allers et al. (2016) and Gaussian distributions centered at radius = 1.4±0.08 RJup (the mean and standard deviation of the ra- dius values found in Allers et al. 2016) and period = 8.61±0.06 hours. The resulting distributions for equatorial velocity (derived from radius and our measured rotation period), vsini, sini, and inclination are presented in Fig. 7. We adopt a value for sini of 0.83±0.07, from the median and standard deviation of our sini distribution. Propagating errors in the normal way, we find i = 56.2 ± 7.2◦. Values of sini > 1 are unphysical and are a result of our uncertainties in measuring radius and vsini. We have tried to mitigate this issue in two manners: 1) discarding all values of sini > AASTEX variability characterization of PSO J318.5-22 11 Table 2. Peak-to-Median Variability Amplitudes, HST phases, and phase offsets relative to the Spitzer lightcurve for synthesized HST band lighcurves band amplitude white light 2.51% (1.07 - 1.67 µm) 2MASS J 2MASS H water band 2.92% 2.28% 2.38% (1.34 - 1.44 µm) methane band 2.20% (1.60 - 1.67 µm) 68% 0.11% 0.23% 196.4◦ 95% HST phase 68% 95% phase offset 68% 95% 6.1◦ 2.4◦ 4.8◦ 206.9◦ 3.1◦ 0.16% 0.32% 193.2◦ 0.13% 0.25% 198.1◦ 0.19% 0.37% 199.9◦ 2.9◦ 2.9◦ 3.9◦ 5.7◦ 5.7◦ 7.7◦ 203.7◦ 208.6◦ 210.4◦ 3.4◦ 3.4◦ 3.4◦ 6.8◦ 6.8◦ 8.7◦ 0.16% 0.31% 198.5◦ 3.6◦ 7.2◦ 209.0◦ 4.1◦ 8.1◦ 1, we find an inclination of 56.1±7.4◦ (median and standard deviation of remaining values) and 2) pinning all sini values greater than 1 to 1 (as such a value does imply a high inclination), we find an inclination of 56.2±8.1◦. All methods provide consistent values for inclination. 5.3. Amplitude and phase shifts from the HST light curves We adopted a similar Markov-Chain Monte Carlo methodology to interpret the HST light curves. Since the HST observation does not cover a full period, we fixed the period to 8.6 hours, as determined from the Spitzer light-curve. The MCMC code was run for each of the bands that we extracted lightcurves; posterior pdfs for each band are presented in Appendix A. Again, we adopt the 50%-quantile value as the best value for each parameter. For each synthesized light curve, a set of 100 samples drawn from the posterior pdf is overplotted on the unbinned HST lightcurve for that band in Fig. 8. We find very little covariance between phase and amplitude in the posterior pdfs, suggesting the fits are robust, at least for the purpose of estimating amplitudes and phase shifts. Amplitudes (peak to median value) and phases measured for each band from the MCMC fits, as well as phase offsets relative to the Spitzer lightcurve, are presented in Table 2. 5.4. Model fits to HST Spectroscopy We fit the median, orbit 1 (close to minimum), orbit 3 (bracketing maximum flux), and orbit 5 (close to minimum) spectra of PSO J318.5-22 with three sets of at- mospheric models utilizing very different treatments of cloud parameters: (1) the ExoREM models, which focus specifically on low-surface gravity atmospheres in the cloudy, clear, and partly cloudy cases (Baudino et al. 2015, Charnay et al. in prep), (2) the BT-Settl models, which explore a wide range of dust species grain formation 12 Biller et al. in the presence of hydrodynamical mixing (Allard et al. 2011), and (3) the thick cloud models of Madhusudhan et al. (2011) (henceforth M11). 5.4.1. ExoREM model fits We compared the grid of ExoREM models to the median spectrum of PSO J318.5- 22 through a χ2 minimization. Unlike most other 1-d models, ExoREM enables the modeling of clear, patchy, and fully cloudy atmospheres, parameterized according to the fc parameter, where fc = 0 is a clear atmosphere and fc = 1 is a fully cloud- covered atmosphere. As a consequence of the fits, we renormalized each synthetic spectrum by a dilution factor R2/d2 which minimizes the χ2, where R is the radius and d the distance to the target. The fit was performed independently for each value of fc and we compared the best χ2 a posteriori. We considered a fit with R left unconstrained, and another one with the radius varying in the interval R = 1.4±0.08 RJup (derived from evolutionary model fits assuming an age of 23±3 Myr, Allers et al. 2016). We adopt a distance of 22.2±0.8 pc (parallax of 45.1±1.7 mas) from Liu et al. (2016). The data are systematically best represented by the ExoREM models with full cloud cover. When R is left unconstrained, we find a best fit for Teff = 1250K, log g=3.4 dex, M/H=+0.5 dex, and R =1.12 RJup. When an a-priori range on the radius is given, the spectrum of PSO J318.5-22 is best reproduced for Teff=1150 K, log g=3.3 dex, M/H=0.0 dex, and R=1.39 RJup. We show the solution in Fig. 9 along with a χ2 map for the solar-metallicity models. The model reproduces the spectral slope and the object NIR brightness simultaneously, but it fails to reproduce the detailed morphology of the H band and the strength of the water-absorption at 1.3-1.5 µm. The variation of the spectral features of PSO J318.5-22 are within the error bars of the median spectrum, thus the fitting solutions are identical when our various HST spectra of PSO J318.5-22 are considered. 5.4.2. BT-Settl and M11 model fits For the BT-Settl and M11 models, fits were performed with a Levenberg-Marquardt least square fitter (numpy.optimize.curvef it), after binning the model spectra to the same resolution and sampling as the HST spectra. Best fit models are shown in Fig. 10. While both of these models produce spectra which are roughly qualitatively similar to our observed spectra, there are notable differences, namely, the models do not reproduce the steepness of the observed spectral slope from 1.2 to 1.35 µm or from 1.4 to 1.7 µm. As the difference between the model spectra and our observed spectra were larger than the difference between observed spectra from different orbits, we found that a single model spectrum fit best all the single orbit spectra as well as the median spectrum. In other words, we were unable to describe the changes seen in the observed spectra as a function of time by fitting different model spectra with varying Tef f and surface gravities. The 1D static models used here, however, are essentially averages over the entire visible surface area of the object and over multiple AASTEX variability characterization of PSO J318.5-22 13 rotational periods, so it is is not surprising that we find similar fits for both single orbit and median spectra. The best fit BT-Settl model had Teff=1600 K and log(g) = 3.5, although a range of models with Teff = 1500 − 1700K and log(g) = 3 − 5 fit the spectra nearly as well. The Teff=1600 K and log(g) = 3.5 best fit is driven by the fitting algo- rithm's attempt to fit the spectral slope in H band. In contrast, a Teff=1700 K, log(g) = 5.0 fits the J band slope better, at the expense of a very poor H band fit (similar to what was found for HR 8799de by Bonnefoy et al. 2016). These results are consistent with the best fits reported in Liu et al. (2013) for the IRTF SpeX spectrum (Teff = 1400 − 1600K and log(g) = 4.0 − 4.5 dex). As noted in Liu et al. (2013), this set of atmospheric model spectra fits yield Teff values that are significantly higher than those obtained from evolutionary model fits to photometry. For the Madhusud- han et al. (2011) models, the best fits were obtained for model A (thick clouds) with 60 µm grains, Teff = 1100 − 1200K and log(g) = 3.75 − 4.25 dex, consistent with the IRTF SpeX spectrum fits (model A, 60 µm grains,Teff = 1100K and log(g) = 4.0 dex) reported in Biller et al. (2015). As with the ExoREM model fits, we renormalized each synthetic spectrum by a dilution factor R2/d2 which minimizes the χ2, where R is the radius and d the distance to the target. We again adopt a distance of 22.2±0.8 pc (parallax of 45.1±1.7 mas) from Liu et al. (2016). Thus, from the dilution factor obtained from the model fit, we estimate the radius R of PSO J318.5-22. This results in an unphysically small radius estimate of ∼0.7-0.8 RJup using the BT-Settl models for this young, low surface gravity object. Liu et al. (2013) obtained a similar result fitting this same model set to a low-resolution near-IR spectrum; in contrast, adopting an age of 23±3 Myr, Allers et al. (2016) find Teff = 1100–1200 K and radius values of 1.34 – 1.46 RJup using a variety of evolutionary model grids with and without clouds. The observed luminosity of a (very roughly blackbody) object is governed by the temperature and the radius; the high temperature fit by the BT-Settl models has thus necessitated an unphysically small fit to the radius to produce the observed luminosity. For the M11 models, with a cooler fitted temperature of 1100-1200 K, we find radius estimates of ∼1.0-1.3 RJup, roughly consistent with our ExoREM model fits with radius left as a free parameter, but still somewhat smaller than estimates based on evolutionary models (Liu et al. 2013; Allers et al. 2016). 6. DISCUSSION 6.1. Amplitudes Over our 7 hours of HST monitoring, we only captured one clear extremum, a maxi- mum in brightness that occurs in orbit 3. The minimum value of brightness measured during our time series occurred in orbit 1. However, orbit 1 was the most affected by the ramp effect. Orbit 5 is also near a minimum of the lightcurve and should not be affected as strongly by the ramp effect. The ratios of maximum and minimum spectra 14 Biller et al. (taking both orbit 1 and orbit 3 as potential minima) are plotted in Fig. 11. The orbit 3 spectrum divided by the orbit 5 spectrum shows a monotonic and relatively small decrease in this ratio as a function of wavelength. In contrast, for the orbit 3 spectrum divided by the orbit 1 spectrum, the ratio of maximum to minimum spectral flux in the 1.4 µm water absorption feature is slightly smaller relative to that at ad- jacent shorter and longer wavelengths, superimposed on a monotonic, small decrease in variability amplitude as a function of increasing wavelength. The suppression of variability in the water band relative to the adjacent continuum in orbit 1 appears to be robust – from Fig. 8, all lightcurves except the water band lightcurve display a significant deviation from the sinusoidal MCMC fits during Orbit 1. This is likely not a result of the ramp effect, which should affect all wavelengths equally. For the L/T transition objects SIMP 0136 and 2M 2139, Apai et al. (2013) and Yang et al. (2016) found significantly smaller amplitudes in the water absorption feature relative to both the shorter and longer wavelength continuum, while for 2 mid-L dwarfs, Yang et al. (2015) found small but monotonic decreases of amplitude with increasing wavelength. Our results appear to be a hybrid of these cases, with different behavior observed in different orbits. As our HST observations did not cover a full period, we did not measure the full amplitude of variability. However, we did cover the majority of the period and have a robust period determination from the simultaneous Spitzer observations, so we can use the sinusoidal fits to the lightcurves to estimate amplitudes and phase shifts (relative to the Spitzer lightcurve) for the 5 broadband regions we have considered previously: the full white light spectrum from 1.1 to 1.67 µm, 2MASS J, 2MASS H, water, and methane. As noted previously, this method appears to actually slightly underestimate variability amplitude, as most of the synthesized lightcurves show some deviation from the sinusoidal fits during orbit 1. Measured amplitude and phase shift for each of the broadband regions we considered are plotted as a function of wavelength in Fig. 12. Similar to the divided spectra for orbit 3 and orbit 5, amplitude appears to generally decrease as a function of increasing wavelength, with the sharpest break between J and 1.4 µm water band. The mid-IR Spitzer [4.5 µm] lightcurve follows the same trend of decreasing am- plitude with longer wavelength, with a peak-to-trough amplitude of ∼3.4% vs. 4.4 - 5.8% for the near-IR bands. For field brown dwarfs, near-IR variability is gener- ally found to have a significantly higher amplitude than mid-IR variability if both are present. It is notable that the mid-IR variability amplitude for PSO J318.5-22 is so similar to its near-IR variability amplitude and is in fact one of the highest variability amplitudes ever measured in the mid-IR for a brown dwarf or planetary mass object! The highest amplitude variable from Metchev et al. (2015) is the T6 dwarf 2MASS J22282889-4310262, with a 3.6 µm variability amplitude of 4.6±0.2%, but most of the variables in their sample have amplitudes of <2%. This high ampli- tude variability in the mid-IR may be the effect of low surface gravity on the vertical AASTEX variability characterization of PSO J318.5-22 15 structure of such an atmosphere. Low surface gravity allows cloud species to poten- tially extend up to lower pressures and higher altitudes compared to the high surface gravity case (Marley et al. 2012). In general for brown dwarfs and free-floating plan- etary mass objects, the photosphere in the mid-IR is at lower pressures and higher amplitudes than the photosphere in the near-IR photosphere (see e.g. Marley et al. 2012; Biller et al. 2013a). Thus, the extension of clouds up to lower pressure regions increases the chance of heterogeneous cloud opacity (and hence variability) at the low pressure levels probed by mid-IR observations. The viewing inclination of a given object will significantly affect its variability prop- erties (see e.g. Vos et al. 2017a). Presuming that surface features that generate vari- ability are primarily equatorial, the same object observed at high inclination will appear to have a higher variability amplitude than if viewed at lower inclination. Additionally, Vos et al. (2017a) find that J-band variability is more affected by in- clination angle than mid-IR variability. As J-band observations generally probe a deeper part of the atmosphere than mid-IR observations (Biller et al. 2013b; Yang et al. 2016; Vos et al. 2017a), Vos et al. (2017a) propose that this effect may be due to the increased atmospheric path length of J-band flux at lower inclinations. We mea- sure a relatively high inclination for PSO J318.5-22 of 56±8◦, thus, we are observing close to the full amplitude in each band. 6.2. Phase Shifts We measure a phase shift between the Spitzer lightcurve and HST lightcurves of ∼200◦. At a lower significance, we find a ∼6-7◦ phase shift between the J-band lightcurve and the other near-IR narrow band lightcurves. Similar results have been found for older, high-surface gravity brown dwarfs – phase shifts between the near-IR and mid-IR lightcurves may be quite common for these objects. The first such phase shift was found by Buenzli et al. (2012) for the T6 2MASS J22282889-4310262 and Yang et al. (2016) recently found significant phase shifts between the mid- and near- IR for 4 high-surface gravity brown dwarfs (including 2MASS 2228). In 3 out of 4 cases from Yang et al. (2016), the measured phase shift between near-IR and mid-IR is nearly ∼180◦, ranging from 150-210◦ (see their Figs. 19-22), very similar to the ∼200◦ phase shift reported here for PSO J318.5-22. Phase shifts within the near-IR spectral bands are less common (Buenzli et al. 2012; Biller et al. 2013a; Yang et al. 2016) but have been reported now at multiple epochs for the T6 2MASS J22282889- 4310262 (Buenzli et al. 2012; Yang et al. 2016). For PSO J318.5-22, all the near-IR bands we consider agree in phase at the 2−σ level; at the 1−σ level, 2MASS J is shifted by ∼6 degrees relative to the other near-IR bands. However, given that we have monitored less than one rotation period with HST and have also assumed a sinusoidal light curve shape, this phase shift is still within the errors expected from our sinusoidal model fitting. Observed phase shifts have generally been interpreted as different "top-of-atmosphere" locations at different wavelengths – near-IR generally 16 Biller et al. probes deeper in the atmosphere than mid-IR (Buenzli et al. 2012; Biller et al. 2013a; Yang et al. 2016). In other words, the source of inhomogeneity which drives near-IR variability is located at a higher pressure level deeper in the atmosphere than the source of inhomogeneity driving mid-IR variability at a lower pressure level. 6.3. Principal Component Analysis of HST spectra Following the method of Apai et al. (2013), we perform principal component analysis (henceforth PCA) to determine how many spectral components drive the variability for PSO J318.5-22. We expect the observed variability to be driven by the rotation in and out of view of regions with differing spectra. PCA identifies the smallest set of independent spectra that account for the majority of the observed variability. Taking each of the 49 spectra in our spectral sequence as a dimension and subtracting off the mean value for each spectrum (as required for PCA), we calculate the 49×49 covariance matrix between each spectrum using the numpy.cov function in python. We then used numpy.linalg.eig to determine eigenspectra and eigenvalues. Sorting on eigenvalues, the principal spectral component accounted for 99.6% of the observed variability, with the second component contributing 0.1% and the third component contributing 0.07%. This is similar to results for other variable objects – Apai et al. (2013) find that the principal spectral component accounted for 99.6% and 99.7% of the variability for 2M2139 and SIMP 0136 respectively. They argue that this implies that only two types of surface patches are required to explain the observed variability for these objects. However, subtracting off the mean essentially means that we remove any gray variation – any shifting of the entire spectrum by a constant value. From Fig. 11 and Fig. 12, the observed variability does not possess a strong color component, as variability amplitude changes monotonically and slowly with wavelength over the 1.07-1.67 µm spectral range of the WFC3 G141 grism – for instance, the variability amplitude in the methane band is roughly 75% of that in the J band. Thus, removing the mean for the PCA actually removes most of the observed variability. Hence it is not surprising that the principal spectral component (which closely resembles the median spectrum) encompasses the vast majority of the power in the time-series. 6.4. Possible Sources of Variability The key variables for constraining variability properties in both field brown dwarfs and young exoplanets are spectral type (which correlates with Teff) and surface grav- ity. High temperature objects (spectral types from early to mid-L) may have vari- ability from cloud features as well as magnetic activity (e.g. star-spots and aurorae); lower temperature objects (>L5) are assumed to have atmospheres too cool to produce magnetic activity and are presumed to have entirely cloud-based variability (Gelino et al. 2002). The variability of older, high-surface gravity field dwarfs across the full L-T spectral has been studied in detail spectroscopically (Buenzli et al. 2015b,a; Apai et al. 2013; Buenzli et al. 2012; Yang et al. 2015). However, only three young, low- surface gravity objects have HST spectral variability monitoring to date – SIMP0136 AASTEX variability characterization of PSO J318.5-22 17 (Apai et al. 2013), W0047 (Lew et al. 2016), and the observations presented herein for PSO J318.5-22. Here we consider a number of drivers of variability suggested in the literature for both high and low surface gravity objects (specifically patchy cloud features, hot spots, high-level hazes, thermochemical instabilities, and magnetically- driven aurorae), to determine if they can describe the variability properties observed for PSO J318.5-22. Considering patchy salt and sulfide clouds as well as hot spot models (heating at a specific pressure level) for high-surface gravity objects with Teff >375 K, Morley et al. (2014) find that variability due to patchy clouds should drive high amplitude variabil- ity across a wide spectral range while variability due to hot spots should produce larger variability amplitudes within absorption features relative to continuum wavelengths. As noted in Section 6.2, for PSO J318.5-22, we find a smooth decrease of variability amplitude as a function of increasing wavelength across the 1.07-1.67 µm spectral range. For the 1.4 µm water absorption feature and the 1.6µm methane absorption feature covered by the HST WFC3 grism, we find similar or slightly smaller variabil- ity amplitudes relative to the adjacent continuum. The lack of stronger variability within absorption features compared to continuum wavelengths for PSO J318.5-22 implies that inhomogeneous cloud features (thick and thin clouds or a haze layer over a thick cloud surface) is likely to be a major driver of the observed variability. The 1.4 µm water absorption feature can potentially provide a useful diagnostic of cloud-driven variability mechanisms. This feature is available from space with HST but is very difficult to observe from the ground. In the case of variability from high- level hazes (as observed for two mid-L field dwarfs by Yang et al. 2015), we expect similar variability amplitudes both with and within the water absorption feature, if the high-level hazes driving the variability are located at a very low pressure, high altitude in the atmosphere where the water opacity is negligible. In the case of variability due to inhomogeneous thin and thick clouds, we expect the variability amplitude to be notably different in the water absorption feature relative to adjacent non-absorbed wavelengths. For instance, Apai et al. (2013) found variability to be suppressed at 1.4 µm relative to the adjacent continuum for two highly variable L/T transition brown dwarfs. However, as noted in Section 6.2 and Fig. 11, PSO J318.5-22 displays both behaviors in different orbits! Thus, in this case, the amplitude inside and outside the 1.4 µm water absorption feature does not clearly distinguish between the cases of high-level hazes vs. inhomogeneous thin and thick clouds. However, recent work has suggested that clouds may not be necessary to model the spectra Tremblin et al. (2016, 2017) – or variability – of L and T spectral type objects. Tremblin et al. (2016) have recently produced cloud-free models of L and T type brown dwarf atmospheres, successfully modeling the red colors of L dwarfs as well as the T dwarf J band brightening and re-emergence of the FeH absorption feature using additional convection from to thermo-chemical instabilities (in the CO / CH4 transition in the case of the L/T boundary). They suggest that turbulence 18 Biller et al. produced by CO or temperature fluctuations across the CO / CH4 may be a driver of the observed variability of brown dwarfs, in particular, that the inhomogeneous top- of-atmosphere structure mapped via Doppler imaging for the L/T transition brown dwarf Luhman 16B (Crossfield et al. 2014) may be explained by inhomogeneities in CO vs. CH4 abundance or temperature. We consider whether this mechanism might drive variability for PSO J318.5-22. Variability due to abundance variations should drive increased variability amplitudes in the absorption features produced by the species in question. For this reason, we produced synthesized lightcurves in the 1.6 µm methane absorption feature – at least the portion of it which lies within the 1.07-1.67 µm spectral range of the HST WFC3 G141 grism. We do not find variability amplitude in the methane band to be significantly enhanced or suppressed relative to the wider 2MASS H band (see Fig. 12). We tentatively suggest the observed variability is not driven by varying CH4 abundance here – however the full theoretical calculation of expected variability amplitude in methane absorption features due to thermochemical instabilities is not yet available, thus, we await more quantitative theoretical predictions here. While lower (>L5) are temperature objects commonly assumed to lack magnetically-driven variability (Gelino et al. 2002), Hallinan et al. (2015) suggest that this may not always be the case. Hallinan et al. (2015) find significant phase shifts between radio and various optical bands for the much-hotter nearby M8.5 ob- ject LSR J1835+3259, which they interpret as auroral heating. In particular, electron beams from global auroral current systems feeds energy from the magnetosphere into the atmosphere of this object. Hallinan et al. (2015) suggest that this mechanism may extend down even to very cool brown dwarfs, driving some of the more extreme ex- amples of weather phenomena in brown dwarfs. They note as well that radio emission has been detected from objects with spectral types as late as T6.5. From our current dataset, we cannot confirm or refute if this is the case for PSO J318.5-22 – radio and H-α monitoring would be necessary to do so. However, more generally, Miles-P´aez et al. (2017) searched for H-α emission for a sample of eight L3-T2 field brown dwarfs, six of which have detections of photometric variability. The only H-α detection in this sample was from a non-variable T2 dwarf, suggesting that aurorae and other chromospheric activity do not commonly drive variability for L and T spectral type objects. 6.5. Theoretical consideration of observed amplitudes and phase shifts in the framework of cloud-driven variability Assuming cloud-driven variability, what cloud species and cloud geometries are necessary to reproduce our observed amplitudes and phase shifts across the near and mid-IR? To try to quantify the expected pressure level at which the photosphere is found at a given wavelength for PSO J318.5-22, we considered the best-fit model to our HST median spectrum to identify at what pressure level flux is being emitted at each AASTEX variability characterization of PSO J318.5-22 19 wavelength. As our BT-Settl model fit in Section 5.4.2 yielded a higher temperature and smaller radius than is consistent with evolutionary model fits to the same object (Liu et al. 2013; Allers et al. 2016), we consider only our model fits using the M11 models and the ExoREM models for this analysis. We adopt the ExoREM fully cloudy model with Teff = 1150K, log g=3.3 dex, M/H=0.0 dex, and R=1.39RJup. In Biller et al. (2015), we found that the SpeX spectrum for PSO J318.5-22 is best fit with the A60, 1100 K, solar metallicity, log(g)=4 model from Madhusudhan et al. (2011) and found similar fits for our HST time-resolved spectra (See Section 5.4). As the SpeX spectrum fit is similar to what we find here and covers a wider wavelength range, we adopt this model fit for the M11 models. Photospheric pressures are provided with the publicly available M11 models; for the ExoREM models, a pressure spectrum was generated by combining the model spectrum with the pressure / temperature profile of the model. Flux at each wavelength was converted to the equivalent brightness temperature. We then interpolated using the corresponding pressure / temperature profile to obtain the photospheric pressure level. This method is correct if the source function varies linearly with optical depth. The resulting "pressure spectra" for both of these models are plotted in Fig. 13, with the bandwidth for each of our lightcurves overplotted. While the photospheric pressure level vs. wavelength varies between models, in both cases the mid-IR flux is generated higher in the atmosphere than the near-IR flux. What cloud species are expected to dominate at the respective higher and lower pressures probed by near-IR vs. mid-IR observations? In Fig. 14, we plot the pres- sure / temperature profile for both our best fit ExoREM cloudy model and an equiv- alent clear model with other parameters unchanged. Thick lines correspond to the photosphere (computed from 0.6 to 5 µm) and dashed lines are condensation tem- peratures for the different clouds present in the model. The presence of clouds (red curve) increases the temperature by around 200 K. In the cloudy case, the type of cloud which condenses varies according to the pressure / temperature profile and the condensation temperatures for different cloud species. Silicate and iron clouds form at around 1 bar and are optically thick up to around 0.3 bar at 1 µm. Thus, for these model atmospheres, silicate and iron clouds form below the photosphere pressures of the 1.4 µm water band and the 4.5 µm CO band. In contrast, Na2S clouds form in the upper atmosphere at around 0.06 bar, so above the photosphere pressures of all molecular bands except the 4.5 µm CO band. Longitudinal variations in cloud thickness can potentially produce anti-correlated in other words ∼180◦ phase shifts between the near-IR and mid-IR variability, lightcurves. In Fig. 15, we plot wavelength vs. brightness temperature, showing where in the spectrum the brightness temperature increases/decreases with clouds. In the cloudy case, the brightness temperature increases at longer wavelengths (e.g. 3 and 4.5 µm) by around 200 K relative to the clear case. The opposite is true at shorter wavelengths (∼1-2 µm), where the brightness temperature decreases by 200 20 Biller et al. K relative to the clear case. A hole in the cloud cover would thus produce a 180 phase shift between near-IR and mid-IR lightcurves, except for the 1.4 µm water band which would be correlated with the mid-IR lightcurve, contrary to the observations presented herein for PSO J318.5-22. However, we do not expect any fully clear patches on this object (and indeed previous work suggests this is the case for brown dwarfs in gen- eral, Apai et al. 2013), but rather longitudinal variations in the cloud thickness. The phase shift and the amplitude of light curves are then very dependent on the altitude, thickness, and placement of different cloud species. We consider a number of simple geometries for both silicate / iron clouds and sulfide clouds to model our observed amplitudes and phases. In Fig. 16, we computed the light curve amplitude that would be produced assuming: (1) a spot with optically thinner silicate and iron cloud thickness, covering 10% of the surface, with homoge- neous thick silicate and iron clouds over the rest of the surface and (2) one hemisphere covered by high-altitude sulfide clouds and no sulfide clouds on the other hemisphere, with homogeneous silicate/iron clouds below the sulfide cloud layer altitude for both hemispheres. Case (2a) was computed assuming no horizontal heat redistribution between the less cloudy spot and the rest of the brown dwarf. Case (2b) was com- puted with no horizontal heat redistribution (solid line) and with very efficient heat redistribution (dashed line). For longitudinal variations (Case 1) in silicate and iron cloud thickness, we predict large variations in the amplitude within the 1.07 - 1.67 µm spectral range of the HST WFC3 G141 grism and, additionally, the 1.4 µm water band lightcurve should be correlated with the 4.5 µm Spitzer Channel 2 lightcurve. For longitudinal variations in the sulfide cloud cover (Case 2a and b), the amplitude is quite constant in the HST bands and for a case intermediate between efficient heat re- distribution and no heat redistribution, the predicted amplitude and the phase shifts between different near-IR and mid-IR wavelengths could be compatible with our HST and Spitzer observations. This modelling remains very preliminary, but suggests that variations in the cover of high altitude clouds could begin to explain the observations presented herein. Na2S clouds are a good candidate for such high altitude clouds since they form at very low pressures high in the atmosphere (0.06 bar). Inhomogeneous Cr and MnS clouds also are potential candidates. An upper layer of silicate clouds could also produce such variability but it would require a mechanism for forming or transporting cloud particle higher than the cloud deck. Cloud convection triggered by latent heat release (Tan & Showman 2017) or radiative heating (Freytag et al. 2010) may produce vertically extended clouds and a detached silicate haze layer. Our measured phase shifts between the HST bands and the Spitzer 4.5 µm lightcurve are in fact somewhat more than 180◦, which is unsurprising, as the cloud geometry for PSO J318.5-22 is certainly more complicated than the simple geometries considered above. Modeling approaches that combine multiple 1-D models (such as the one pre- sented herein and Artigau et al. 2009; Apai et al. 2013; Karalidi et al. 2015, 2016) can reproduce correlated variability or 180◦ anti-correlated variability, but not other AASTEX variability characterization of PSO J318.5-22 21 phase shifts. As demonstrated above, where the "top-of-atmosphere" occurs varies depending on wavelength and the specific opacity sources that dominate at different atmospheric levels. The observed "phase shifts" may simply be heterogeneous and uncorrelated structure at different altitudes, which is still modulated by the rotation period of the object in question. Likely full 3-d models will be necessary to describe this structure, such as the C05BOLD model currently undergoing testing (Allard et al. in prep), especially as rotation probably plays a significant role in the appearance and features of these atmospheres (Showman & Kaspi 2013). 6.6. Variability in Low-Surface Gravity L dwarfs Metchev et al. (2015) find increased mid-IR variability amplitudes for 8 low-surface gravity L3 to L5 objects with respect to the rest of their older, high-surface gravity survey sample. We tentatively find that such a trend (in both the near and mid- IR) may continue for low-surface gravity mid-to-late L dwarfs. Only 4 such objects have been surveyed in either the near or mid-IR to date. Three out of the 4 have positive variability detections in both near- and mid-IR (Biller et al. 2015; Lew et al. 2016; Vos et al. 2017b; Morales-Calder´on et al. 2006); one is a non-detection in our ongoing SofI survey (Vos et al. in prep). The three low-surface gravity mid-to- late L dwarfs with positive variability detections are PSO J318.5-22, W0047, and 2M2244. For our HST+Spitzer monitoring of PSO J318.5-22, we found peak-to- trough amplitudes of ∼3.4% for Spitzer Channel 2 and 4.4 - 5.8% in the near-IR band (1.07-1.67 µm) covered by the WFC3 G141 prism. In the discovery epoch, Biller et al. (2015) found peak-to-trough variability amplitudes of 7-10% in the JS band and ∼3% in KS, indicating evolution of the variability between the discovery epoch and our HST observations. The lower amplitude in K vs. J during the discovery epoch is consistent with our finding in this work that variability amplitude decreases with increasing wavelength across the 1.1-1.7 µm spectral range of the HST WFC3 grism. Lew et al. (2016) find a similarly high near-IR variability amplitude for W0047, with the relative variability amplitude decreasing from 11% at 1.1 µm to 6.5% at 1.7 µm. Vos et al. (2017b) reported a mid-IR detection for this object with a relative variability amplitude of 1.07±0.04%. Morales-Calder´on et al. (2006) measured a Spitzer channel 1 peak-to-peak variability amplitude of 8 mmag for 2M2244, an L6.5 AB Dor member (Vos et al. 2017b). Variability in this object has recently been confirmed by (Vos et al. 2017b) who found a Spitzer channel 1 peak-to-peak variability amplitude of 0.8±0.2% and ≥3% variability amplitude in J band in a 4-hour long, J band UKIRT WFCAM observation of this object. All three of these objects have notably high near-IR amplitudes compared to field brown dwarfs with similar spectral types as well as planetary mass objects with earlier spectral types. For instance, the detection of variability in the L5 planetary mass object 2M1207b has a considerably lower near- IR amplitude of 1-2% (Zhou et al. 2016). Of the three, PSO J318.5 also has a notably 22 Biller et al. high mid-IR amplitude; mid-IR amplitudes for the other two objects are more in line with typical values for field brown dwarfs. With such a small number of low-surface gravity mid-to-late L variables to study, it is not clear whether these three objects are unusual – or if low-surface gravity objects are inherently more variable then their high surface gravity counterparts. The peak-of-variability for field brown dwarfs appears to be at the L/T transition (Radigan et al. 2014; Radigan 2014), commonly attributed to the breakup or at least thinning of silicate clouds at this spectral type transition Apai et al. (2013). It is hard to say if this is the case for low surface gravity objects, with three high-amplitude near-IR detections for mid-to-late L low surface gravity objects (Biller et al. 2015; Zhou et al. 2016; Vos et al. 2017b), one high-amplitude detection in a young T2.5 object (Artigau et al. 2009; Gagn´e et al. 2017), and one tentative detection in a young T3.5 object (Naud et al. 2017). Predominantly early L low-surface gravity objects have been surveyed to date (Vos et al. in prep), largely due to the current scarcity of late L, L/T transition, and T spectral type young, low-surface gravity objects. Nonetheless, the few mid-to-late L objects surveyed to date appear to be notably variable, which is surprising given that late-L objects are expected to have thick (and probably homogeneous) cloud cover. If late-L spectral type young objects are as a class highly variable, this may draw into question the interpretation of high amplitude variability as the breakup of silicate clouds between the L and T spectral type. Low-surface gravity mid-to-late L dwarfs are particularly interesting because these objects are excellent proxies for several known giant exoplanet companions. The spectra of PSO J318.5-22 and W0047 are nearly identical to those of inner two HR 8799 planets (Bonnefoy et al. 2016). De Rosa et al. (2016) find that the spectrum of the particularly red planet HIP 95086b (Rameau et al. 2013) closely matches that of 2M2244. The newly-discovered exoplanet companion HIP65426b also has an L5-L7 spectral type (Chauvin et al. 2017). Given the significant variability of PSO J318.5- 22, W0047, and 2M2244, we may expect exoplanet companions such as HR 8799bcde, HIP 95086b, and HIP 65426b to be similarly variable, although viewing angle (likely pole-on for the HR 8799 system) may render that variability hard to detect. 6.7. Are Young, Planetary Mass Objects Fast Rotators? Even if young, planetary mass objects have significant top-of-atmosphere inhomo- geneities, we will only be able to detect such features if these objects are relatively rapid rotators (periods <20 hours or so). Many old, field brown dwarfs are rapid rotators (Zapatero Osorio et al. 2006). From conservation of angular momentum, one might expect young objects to be predominantly slower rotators compared to old, field brown dwarfs, as they have somewhat inflated radii (e.g. ∼1.4 RJup for PSO J318.5-22 Allers et al. 2016) compared to older objects (radius ∼1 RJup) and will be expected to spin up with age as they contract. At least preliminarily, however, there is a small cohort of young (≤150 Myr), planetary mass objects with periods AASTEX variability characterization of PSO J318.5-22 23 <20 hours, including PSO J318.5-22, as well as the bonafide exoplanet β Pic b and 2M1207b. In Fig. 17, we plot estimated object mass vs. measured equatorial velocity for these objects, solar system objects, and field brown dwarfs with measured periods from Vos et al. (2017a). Planetary mass objects seem to encompass a similar range of equatorial velocities as older, field brown dwarfs, with both rapid rotators and no- table slow rotators such as the young, 30-40 MJup brown dwarf companion GQ Lup b (Schwarz et al. 2016). However, statistics are still too sparse for a robust comparison to the brown dwarf population in general. Preliminary analyses do suggest that the rotation rate between free-floating and companion objects is similar – Bryan et al. (2017) recently measured rotation rate for a number of companions with masses <20 MJup. Combining their measurements with others in the literature, they found no discernable difference in rotation speed between companions and free-floating objects with similar masses for a small sample of 11 objects. 7. CONCLUSIONS Here we present simultaneous HST WFC3 + Spitzer IRAC variability monitoring for the variable planetary mass object PSO J318.5-22. Our simultaneous HST + Spitzer observations covered >2 rotation periods with Spitzer and most of a rotation period with HST. The main results from these observations are: • Detection of high amplitude variability in both near-IR and mid-IR bands with a period of 8.6±0.1 hours. We estimate peak-to-trough variability amplitudes of 3.4±0.1% for Spitzer Channel 2 and 4.4 - 5.8% (typical uncertainty of ∼0.3%) in the near-IR bands (1.07-1.67 µm) covered by the WFC3 G141 prism. • a relatively high inclination for PSO J318.5-22 of 56±8◦, derived by combining our measured period with the measured vsini from (Allers et al. 2016) for this object. Thus, we are observing close to the full intrinsic variability amplitude in each band. • Detection of 200-210◦ (typical uncertainty of ∼4%) phase offsets between the near-IR and mid-IR lightcurves, likely indicating varying longitudinal atmo- spheric structure at different depths in this atmosphere • Tentative detection of a small ∼6◦ phase offset between the 2MASS J band and the rest of the near-IR bands, but this is at a considerably lower significance level than the mid-IR vs. near-IR phase shift • A decrease of variability amplitude as a function of increasing wavelength, as has previously been found for field brown dwarfs (c.f. among others Apai et al. 2013; Radigan et al. 2014; Yang et al. 2016). We tentatively find that the amplitude of variability in the 1.4 µm water absorption feature is slightly smaller than adjacent wavelengths in the first orbit of our observations, but similar to adjacent wavelengths in the final orbit of our observations. 24 Biller et al. • Detection of similar variability amplitudes in wide spectral bands relative to absorption features, suggesting that the driver of the variability may be inho- mogeneous clouds (perhaps variations in the cover of high altitude clouds over a homogeneous layer of thick clouds) as opposed to hot spots or compositional inhomogeneities at the top-of-atmosphere level. Na2S clouds are a good candi- date high altitude cloud species since they form at very low pressures high in the atmosphere (0.06 bar). Inhomogeneous Cr and MnS clouds also are potential candidates. Both mid-IR and near-IR variability amplitudes for PSO J318.5-22 are large – com- parable with that of high-amplitude L/T transition brown dwarfs and considerably larger than found for the early or mid-L dwarfs (Yang et al. 2015; Radigan et al. 2014; Metchev et al. 2015). Clearly, while low surface gravity late-L planetary ana- logues share some variability properties with field brown dwarfs, they are their own unique category of objects and merit the same in-depth observation and analysis. Given the significant variability of PSO J318.5-22 and other mid-to-late L low sur- face gravity objects, we may expect variability as well in exoplanet companions such as HR 8799bcde, HIP 95086b, and HIP 65426b which share similar spectral types and surface gravities. Based on observations made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with program # 14188. KNA acknowledges support for program #14188 provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. This work is based in part on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. BAB and JV also acknowledge support from STFC grant ST/J001422/1. We thank Jack Gallimore for providing the posterior vsini distribution for PSO J318.5-22 and Mike Cushing for a close reading of this manuscript and useful conversations. REFERENCES Allard, F., Homeier, D., & Freytag, B. 2011, in Astronomical Society of the Pacific Conference Series, Vol. 448, 16th Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, ed. C. Johns-Krull, M. K. Browning, & A. A. West, 91 Allers, K. N., Gallimore, J. F., Liu, M. C., & Dupuy, T. J. 2016, Astrophys. J, 819, 133 Apai, D., Radigan, J., Buenzli, E., et al. 2013, Astrophys. J, 768, 121 Apai, D., Karalidi, T., Marley, M. S., et al. 2017, Science, 357, 683 AASTEX variability characterization of PSO J318.5-22 25 Artigau, ´E., Bouchard, S., Doyon, R., & Lafreni`ere, D. 2009, Astrophys. J, 701, 1534 Barman, T. S., Macintosh, B., Konopacky, Q. M., & Marois, C. 2011, Astrophys. J, 733, 65 Baudino, J.-L., B´ezard, B., Boccaletti, A., et al. 2015, A&A, 582, A83 Biller, B. A., Crossfield, I. J. M., Mancini, L., et al. 2013a, Astrophys. J. Lett., 778, L10 Biller, B. A., Liu, M. C., Wahhaj, Z., et al. 2013b, Astrophys. J, 777, 160 Biller, B. A., Vos, J., Bonavita, M., et al. 2015, Astrophys. J. Lett., 813, L23 Bonnefoy, M., Zurlo, A., Baudino, J. L., et al. 2016, Astron. & Astrophys., 587, A58 Bryan, M. L., Benneke, B., Knutson, H. A., Batygin, K., & Bowler, B. P. 2017, ArXiv e-prints, arXiv:1712.00457 Buenzli, E., Marley, M. S., Apai, D., et al. 2015a, Astrophys. J, 812, 163 Buenzli, E., Saumon, D., Marley, M. S., Gagn´e, J., Faherty, J. K., Burgasser, A. J., et al. 2017, ApJL, 841, L1 Gelino, C. R., Marley, M. S., Holtzman, J. A., Ackerman, A. S., & Lodders, K. 2002, Astrophys. J, 577, 433 Gizis, J. E., Faherty, J. K., Liu, M. C., et al. 2012, Astron. J., 144, 94 Gizis, J. E., Dettman, K. G., Burgasser, A. J., et al. 2015, Astrophys. J, 813, 104 Hallinan, G., Littlefair, S. P., Cotter, G., et al. 2015, Nature, 523, 568 Karalidi, T., Apai, D., Marley, M. S., & Buenzli, E. 2016, Astrophys. J, 825, 90 Karalidi, T., Apai, D., Schneider, G., Hanson, J. R., & Pasachoff, J. M. 2015, Astrophys. J, 814, 65 Lew, B. W. P., Apai, D., Zhou, Y., et al. 2016, Astrophys. J. Lett., 829, L32 Liu, M. C., Dupuy, T. J., & Allers, K. N. 2016, Astrophys. J, 833, 96 Liu, M. C., Magnier, E. A., Deacon, N. R., et al. 2013, Astrophys. J. Lett., 777, L20 et al. 2015b, Astrophys. J, 798, 127 Madhusudhan, N., Burrows, A., & Currie, Buenzli, E., Apai, D., Morley, C. V., et al. T. 2011, Astrophys. J, 737, 34 2012, Astrophys. J. Lett., 760, L31 Chauvin, G., Desidera, S., Lagrange, A.-M., et al. 2017, ArXiv e-prints, arXiv:1707.01413 Crossfield, I. J. M., Biller, B., Schlieder, J. E., et al. 2014, Nature, 505, 654 De Rosa, R. J., Rameau, J., Patience, J., et al. 2016, Astrophys. J, 824, 121 Dupuy, T. J., & Kraus, A. L. 2013, 341, 1492 Faherty, J. K., Riedel, A. R., Cruz, K. L., et al. 2016, Astrophys. J. Supplement, 225, 10 Filippazzo, J. C., Rice, E. L., Faherty, J., et al. 2015, ApJ, 810, 158 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306 Freytag, B., Allard, F., Ludwig, H.-G., Homeier, D., & Steffen, M. 2010, Astron. & Astrophys., 513, A19 Gagn´e, J., Lafreni`ere, D., Doyon, R., Malo, L., & Artigau, ´E. 2014, Astrophys. J, 783, 121 Marley, M. S., Saumon, D., Cushing, M., et al. 2012, Astrophys. J, 754, 135 Metchev, S. A., Heinze, A., Apai, D., et al. 2015, Astrophys. J, 799, 154 Mighell, K. J., Glaccum, W., & Hoffmann, W. 2008, in Proc. SPIE, Vol. 7010, Space Telescopes and Instrumentation 2008: Optical, Infrared, and Millimeter, 70102W Miles-P´aez, P. A., Metchev, S. A., Heinze, A., & Apai, D. 2017, ApJ, 840, 83 Morales-Calder´on, M., Stauffer, J. R., Kirkpatrick, J. D., et al. 2006, Astrophys. J, 653, 1454 Morley, C. V., Marley, M. S., Fortney, J. J., & Lupu, R. 2014, Astrophys. J. Lett., 789, L14 Naud, M.-E., Artigau, ´E., Rowe, J. F., et al. 2017, AJ, 154, 138 Radigan, J. 2014, Astrophys. J, 797, 120 Radigan, J., Lafreni`ere, D., Jayawardhana, R., & Artigau, E. 2014, Astrophys. J, 793, 75 26 Biller et al. Rameau, J., Chauvin, G., Lagrange, Tremblin, P., Chabrier, G., Baraffe, I., A.-M., et al. 2013, Astrophys. J. Lett., 772, L15 Schwarz, G. 1978, Ann. Statist., 6, 461. https: //doi.org/10.1214/aos/1176344136 Schwarz, H., Ginski, C., de Kok, R. J., et al. 2016, ArXiv e-prints, arXiv:1607.00012 Showman, A. P., & Kaspi, Y. 2013, Astrophys. J, 776, 85 Snellen, I. A. G., Brandl, B. R., de Kok, R. J., et al. 2014, Nature, 509, 63 Tan, X., & Showman, A. P. 2017, ApJ, 835, 186 Tremblin, P., Amundsen, D. S., Chabrier, G., et al. 2016, Astrophys. J. Lett., 817, L19 et al. 2017, Astrophys. J, 850, 46 Vos, J. M., Allers, K. N., & Biller, B. A. 2017a, ApJ, 842, 78 Vos, J. M., Allers, K. N., Biller, B. A., et al. 2017b, ArXiv e-prints, arXiv:1710.07194 Wilson, P. A., Rajan, A., & Patience, J. 2014, Astron. & Astrophys., 566, A111 Yang, H., Apai, D., Marley, M. S., et al. 2015, Astrophys. J. Lett., 798, L13 -. 2016, Astrophys. J, 826, 8 Zapatero Osorio, M. R., Mart´ın, E. L., Bouy, H., et al. 2006, Astrophys. J, 647, 1405 Zhou, Y., Apai, D., Lew, B. W. P., & Schneider, G. 2017, AJ, 153, 243 Zhou, Y., Apai, D., Schneider, G. H., Marley, M. S., & Showman, A. P. 2016, Astrophys. J, 818, 176 AASTEX variability characterization of PSO J318.5-22 27 Figure 1. Spitzer (crosses) and HST lightcurves (filled circles) for PSO J318.5-22, after correction for the ramp effect. The lightcurves have been binned to increase S/N ratio, resulting in a 2.5 minute cadence for Spitzer and a 14 minute cadence for HST. The least- squares best fit to the Spitzer lightcurve is shown as a solid purple line. HST lightcurves are shown binned over 5 spectral bandwidths: the full usable 1.07 - 1.67 µm spectral bandwidth of the HST grism spectroscopy (white light, black circles), the 2MASS J band (green circles), the 2MASS H band up to the spectral cutoff at 1.67 µm (red circles), a band centered on the 1.4 µm water absorption feature (blue circles), and a band covering as much of the 1.6 µm methane absorption features as falls in the HST G141 grism spectral bandwidth (purple circles). Small colored points are the 6 background stars in the HST field after being detrended by the calibration curve; PSO J318.5-22 is clearly variable compared to the reference stars. The large 200-210◦ phase offsets between the near-IR and mid-IR lightcurves likely indicates varying longitudinal atmospheric structure at different depths in this atmosphere. 0.00.10.20.30.40.50.60.7time (frac. day)0.960.981.001.021.04normalized fluxwhite lightJHwatermethane 28 Biller et al. Figure 2. Top: The mean and median spectra across the full 5 orbit HST observation. Significant spectral variability is apparent. Bottom: Similar spectra for one of the well- behaved, non-variable reference stars in the HST field. The legend is shared between both panels. 1.11.21.31.41.51.61.7wavelength (micron)2•10-173•10-174•10-175•10-176•10-177•10-17flux (erg/s/cm2/angstrom)medianmeanorbit 1orbit 2orbit 3orbit 4orbit 51.11.21.31.41.51.61.7wavelength (micron)1.5•10-162.0•10-162.5•10-16flux (erg/s/cm2/angstrom) AASTEX variability characterization of PSO J318.5-22 29 Figure 3. The top panel shows the normalized, pixel phase corrected lightcurve of PSO J318.5-22 with best-fit sinusoidal function overplotted in red. The middle panel shows the best fit function with Gaussian noise added – this simulated lightcurve closely resembles the observed lightcurve. The bottom panel shows the periodogram of the target and the simulated curve, as well as the periodogram of several reference stars in the field. The blue dashed line shows the 1% false-alarm probability. 51015Elapsed Time (hr)0.960.981.001.021.04Relative FluxSinusoidSpitzer IRAC [4.5]51015Elapsed Time (hr)0.960.981.001.021.04Simulated Flux051015202530Period (hr)020406080100120140Power 30 Biller et al. Figure 4. Posterior probability distributions of parameters from sinusoid MCMC fits to our Spitzer Channel 2 lightcurve of PSO J318.5-22. The mean parameter is the mean value of the lightcurve – since we have divided the raw lightcurve by the median flux over the whole observation, this should tend towards unity. In the marginalized confidence interval plots, the middle dashed line gives the median, the two outer vertical dashed lines represent the 68% confidence interval. The contours show the 1, 1.5 and 2-σ levels. AASTEX variability characterization of PSO J318.5-22 31 Figure 5. Posterior probability distributions of parameters from sinusoid+linear MCMC fits to our Spitzer Channel 2 lightcurve of PSO J318.5-22. In the marginalized confidence interval plots, the middle dashed line gives the median, the two outer vertical dashed lines represent the 68% confidence interval. The contours show the 1, 1.5 and 2-σ levels. 32 Biller et al. Figure 6. 100 sample model lightcurves drawn respectively from our sinusoid (green) and sinusoid+linear (black) MCMC fits to the Spitzer Channel 2 lightcurve (plotted as filled red circles). AASTEX variability characterization of PSO J318.5-22 33 Figure 7. Left: Gaussian distribution in equatorial velocity derived from our measured period and radius estimates from (Allers et al. 2016). Center left: vsini distribution from fits of the high resolution spectrum of PSO J318.5-22 from (Allers et al. 2016). Center right: sini distribution. The shaded gray rectangle indicates values of sini above 1, which are unphysical and are a result of our adopted uncertainties in radius, period, and vsini for this object. Right: inclination distribution. Unphysical values of sini have been pinned to 1 here (i.e. 90◦ inclination). 34 Biller et al. Figure 8. 100 sample model lightcurves drawn from our sinusoid (black) MCMC fits to each synthesized HST lightcurve (white light, 2MASS J, 2MASS H, water, and methane). All lightcurves except the water lightcurve show a significant deviation from the sinusoidal fits in the first 30 minutes of the observation; given that the water lightcurve appears sinusoidal, this is not a result of the ramp effect, which should affect all wavelengths equally. AASTEX variability characterization of PSO J318.5-22 35 Figure 9. Top: Comparison of the HST median spectrum of PSO J318.5-22 to the ExoREM models when the radius is allowed to vary in the range 1.29 - 1.35 RJup. Bottom: χ2 map for the solar-metallicity models with a full cloud cover. The orange square corresponds to the χ2 minimum. 1.11.21.31.41.51.6l [mm]2•10-163•10-164•10-165•10-166•10-167•10-16Fl [W.m-2.mm-1]Median spectrumModel: Teff=1150K, log g=3.3 dex, Fe/H=0.0 dex, fc=1, R=1.39 RJup3.03.54.04.55.05.56.0logg500100015002000Teff3.03.54.04.55.05.56.0logg500100015002000Teff3.5•1041.6•1067.4•1077.4•1073.03.54.04.55.05.56.0logg500100015002000Teff3.03.54.04.55.05.56.0logg500100015002000Teff3.03.54.04.55.05.56.0logg500100015002000Teff 36 Biller et al. Figure 10. Best BT-Settl (red circles) and M11 model (blue crosses) fits overplotted with the HST median spectrum of PSO J318.5-22 (black stars). These models do not reproduce the steepness of the observed spectral slope from 1.2 to 1.35 µm or from 1.4 to 1.7 µm. The best fit BT-Settl model had Teff =1600 K and log(g) = 3.5, although a range of models with Teff = 1500− 1700K and log(g) = 3− 5 fit the spectra nearly as well. The Teff =1600 K and log(g) = 3.5 best fit is driven by the fitting algorithm's attempt to fit the spectral slope in H band. For the Madhusudhan et al. (2011) models, the best fits were obtained for the model A (thick clouds), 60 µm grains, Teff = 1100 − 1200K and log(g) = 3.75 − 4.25 dex. AASTEX variability characterization of PSO J318.5-22 37 Figure 11. Ratio of maximum and minimum PSO J318.5-22 HST spectra, binned by 0.05 µm. The minimum value of brightness measured during our time series occurred in orbit 1. However, orbit 1 was the most affected by the ramp effect. Orbit 5 is also near a minimum of the lightcurve and should not be affected as strongly by the ramp effect. Thus, we plot here two max / min spectral ratios: orbit 3 divided by orbit 1 and orbit 3 divided by orbit 5. The spectral ratio for orbit 3 divided by orbit 5 has been offset slightly in wavelength for clarity. As our HST observations did not cover a full period, these are lower limits on the full amplitude. 38 Biller et al. Figure 12. Top: Wavelength vs. measured variability amplitude for HST synthesized lightcurves. Shaded boxes gives the passband used on the wavelength axis and the 1- σ error on the amplitude axis. Amplitude appears to decrease from shorter to longer wavelengths. Bottom: Wavelength vs. measured phase relative to the Spitzer channel 2 lightcurve. Shaded boxes gives the passband used on the wavelength axis and the 1-σ error on the phase shift axis. Phase shifts across each of the synthesized bandpasses agree at the 2-σ level; J band is phase shifted by ∼6◦ relative to the other near-IR bands. AASTEX variability characterization of PSO J318.5-22 39 Figure 13. Pressure spectra for best-fit ExoREM (black, Teff =1150 K, log(g)=3.3 dex, M/H=0.0 dex, and R=1.39 RJup) and M11 models (blue, A60, Teff =1100 K, log(g)=4, solar metallicity). The bandpasses for the HST synthesized lightcurves and the Spitzer channel 2 lightcurve are shown as shaded boxes. For both models, mid-IR flux is generated higher in the atmosphere than near-IR flux. 40 Biller et al. Figure 14. Pressure / temperature profile for both our best fit ExoREM cloudy model (red curve) and an equivalent clear model (blue curve) with other parameters unchanged. Thick lines correspond to the photosphere (computed from 0.6 to 5 µm) and dashed lines are condensation temperatures for the different clouds present in the model. The presence of clouds increases the temperature by around 200 K in the photosphere region. 050010001500200025003000Temperature (K)10-510-410-310-210-1100101102Pressure (bar)FeMg2SiO4MgSiO3Na2SKCl AASTEX variability characterization of PSO J318.5-22 41 Figure 15. Wavelength vs. brightness temperature for our best-fit ExoREM cloudy model (red curve) and an equivalent clear model (blue curve) with other parameters unchanged, showing where in the spectrum the brightness temperature increases/decreases with clouds. In the cloudy case, the brightness temperature increases at longer wavelengths (e.g. 3 and 4.5 µm) by around 200 K relative to the clear case. The opposite is true at shorter wavelengths (∼1-2 µm), where the brightness temperature decreases by 200 K relative to the clear case. While we do not expect any fully clear patches on this object, longitudinal variations in the cloud thickness should produce similar trends and thus a ∼180◦ phase shift between near- and mid-IR lightcurves. 12345Wavelength (microns)60080010001200140016001800Brightness temperature (K)with cloudwithout cloud 42 Biller et al. Figure 16. Predicted light curve amplitude that would be produced assuming: (1) a spot with optically thinner silicate and iron cloud thickness, covering 10% of the surface, and homogeneous thick silicate / iron clouds on the remaining 90% (blue curve) and (2) one hemisphere covered by sulfide clouds and no sulfide clouds on the other hemisphere, with homogeneous silicate/iron clouds for both hemispheres (red curves). Case (2a) was computed assuming no horizontal heat redistribution between the less cloudy spot and the rest of the brown dwarf. Case (2b) was computed with no horizontal heat redistribution (solid line) and with very efficient heat redistribution (dashed line). 11.522.533.544.55Wavelength (microns)-0.15-0.1-0.0500.050.10.15Relative flux50% sulfide (no horizontal heat redistribution)50% sulfide (with efficient horizontal heat redistribution)10% thin silicate/iron (no horizontal heat redistribution) AASTEX variability characterization of PSO J318.5-22 43 Figure 17. Mass vs. equatorial velocity for young, planetary mass objects including PSO J318.5-22, as well as the exoplanet β Pic b, 2M1207b, and two ∼20 MJup members of AB Dor, 2M2244 and W0047. Solar system planets and brown dwarfs with measured periods from Vos et al. (2017a) are plotted as gray circles; young brown dwarfs are plotted as gray circles outlined in black. Planetary mass objects seem to encompass a similar range of equatorial velocities as older, field brown dwarfs, with both rapid rotators and notable slow rotators such as the young, 30-40 MJup brown dwarf companion GQ Lup b (Schwarz et al. 2016). 1x10−21x10−11x1001x10+11x10+21101001x10−21x10−11x1001x10+11x10+2Mass (MJup)110100Equatorial Velocity (km s−1)W00472M22442M1207bβ Pic bPSO 318.5−22JSUNGQ Lupi b 44 Biller et al. Figure 18. Posterior probability distributions of parameters from sinusoid MCMC fits to the HST "white light" lightcurve (full bandpass from 1.07 to 1.67 µm) for PSO J318.5-22. Since the HST observation does not cover a full rotation period, we have fixed the period to 8.6 hours, as found from the Spitzer lightcurve. In the marginalized confidence interval plots, the middle dashed line gives the median, the two outer vertical dashed lines represent the 68% confidence interval. The contours show the 1, 1.5 and 2-σ levels. APPENDIX A. HST LIGHTCURVE MCMC POSTERIORS MCMC posteriors for sinusoidal fits to HST lightcurves are presented in Fig. 18 through Fig. 22. AASTEX variability characterization of PSO J318.5-22 45 Figure 19. Posterior probability distributions of parameters from sinusoid MCMC fits to the HST synthesized 2MASS J lightcurve for PSO J318.5-22. Since the HST observation does not cover a full rotation period, we have fixed the period to 8.6 hours, as found from the Spitzer lightcurve. In the marginalized confidence interval plots, the middle dashed line gives the median, the two outer vertical dashed lines represent the 68% confidence interval. The contours show the 1, 1.5 and 2-σ levels. 46 Biller et al. Figure 20. Same as Fig. 19, for the HST synthesized 2MASS H lightcurve. AASTEX variability characterization of PSO J318.5-22 47 Figure 21. Same as Fig. 19, for the HST synthesized water band lightcurve (1.34 to 1.44 µm). Facilities: HST(WFC3), Spitzer(IRAC) Software: python, astropy, IDL, emcee 48 Biller et al. Figure 22. Same as Fig. 19, for the HST synthesized methane band lightcurve (1.60 to 1.67 µm).
1811.00352
1
1811
2018-11-01T13:11:16
The dynamical evolution of escaped Jupiter Trojan asteroids, link to other minor body populations
[ "astro-ph.EP" ]
Jupiter Trojan asteroids are located around L4 and L5 Lagrangian points on relatively stable orbits, in 1:1 MMR with Jupiter. However, not all of them lie in orbits that remain stable over the age of the Solar System. Unstable zones allow some Trojans to escape in time scales shorter than the Solar System age. This may contribute to populate other small body populations. In this paper, we study this process by performing long-term numerical simulations of the observed Trojans, focusing on the trajectories of those that leave the resonance. The orbits of current Trojans are taken as initial conditions and their evolution is followed under the gravitational action of the Sun and the planets. We find the rate of escape of Trojans from L5, ~1.1 times greater than from L4. The majority of escaped Trojans have encounters with Jupiter although they have encounters with the other planets too. Almost all escaped Trojans reach the comet zone, ~90% cross the Centaur zone and only L4 Trojans reach the transneptunian zone. Considering the real asymmetry between L4 and L5, we show that 18 L4 Trojans and 14 L5 Trojans with diameter D > 1 km are ejected from the resonance every Myr. The contribution of the escaped Trojans to other minor body populations would be negligible, being the contribution from L4 and L5 to JFCs and no-JFCs almost the same, and the L4 contribution to Centaurs and TNOs, orders of magnitude greater than that of L5. Considering the collisional removal, besides the dynamical one, and assuming that Trojans that escape due to collisions follow the same dynamical behavior that the ones removed by dynamics, we would have a minor contribution of Trojans to comets and Centaurs. However, there would be some specific regions were escaped Trojans could be important such as ACOs, Encke-type comets, S-L 9-type impacts on Jupiter and NEOs.
astro-ph.EP
astro-ph
The dynamical evolution of escaped Jupiter Trojan asteroids, link to other minor body populations Romina P. Di Sistoa,b, Ximena S. Ramosc, Tabar´e Gallardod aFacultad de Ciencias Astron´omicas y Geof´ısicas Universidad Nacional de La Plata bInstituto de Astrof´ısica de La Plata, CCT La Plata-CONICET-UNLP Paseo del Bosque S/N (1900), La Plata, Argentina cInstituto de Astronom´ıa Te´orica y Experimental (IATE), Observatorio Astron´omico, Universidad Nacional de C´ordoba, Laprida 854, X5000BGR C´ordoba, Argentina dDepartamento de Astronom´ıa, Facultad de Ciencias, Universidad de la Rep´ublica, Igu´a 4225, 11400 Montevideo, Uruguay Abstract The Jupiter Trojans constitute an important asteroidal population both in number and also in relation to their dynamical and physical properties. They are asteroids located around L4 and L5 Lagrangian points on relatively stable orbits, in 1 : 1 mean motion resonance with Jupiter. However, not all of them lie in orbits that remain stable over the age of the Solar System. Unstable zones allow some Trojans to escape in time scales shorter than the Solar System age. This may contribute to populate other small body populations. In this paper, we study this process by performing long-term numerical simulations of the observed Trojans, focusing on the trajectories of those that leave the resonance. The orbits of current Trojan asteroids are taken as initial conditions and their evolution is followed under the gravita- tional action of the Sun and the planets. We built "occupancy maps" that represent the zones in the Solar System where escaped Trojans should be found. We find the rate of escape of Trojans from L5, ∼ 1.1 times greater than from L4. The majority of escaped Trojans have encounters with Jupiter although they have encounters with the other planets too. The median life- time of escaped Trojans in the Solar System is ∼ 264000 years for L4 and ∼ 249000 years for L5. Almost all escaped Trojans reach the comet zone, ∼ 90% cross the Centaur zone and only L4 Trojans reach the transneptunian Email address: [email protected] (Romina P. Di Sisto) Preprint submitted to Icarus November 2, 2018 zone. Considering the real asymmetry between L4 and L5, we show that 18 L4 Trojans and 14 L5 Trojans with diameter D > 1 km are ejected from the resonance every Myr. The contribution of the escaped Trojans to other minor body populations would be negligible, being the contribution from L4 and L5 to Jupiter-family comets (JFCs) and no-JFCs almost the same, and the L4 contribution to Centaurs and TNOs, orders of magnitude greater than that of L5. Considering the collisional removal, besides the dynamical one, and assuming that Trojans that escape due to collisions follow the same dynamical behavior that the ones removed by dynamics, we would have a minor contribution of Trojans to comets and Centaurs. However, there would be some specific regions were escaped Trojans could be important such as Asteroids in Cometary Orbits (ACOs), Encke-type comets, Shoemaker-Levy 9-type impacts on Jupiter and Near-Earth objects (NEOs). Keywords: Jupiter; Trojan asteroids; numerical techniques 1. Introduction Jupiter Trojans are a population of asteroids in 1 : 1 mean motion reso- nance (MMR) with Jupiter and are located within the L4 and L5 Lagrange points. It is a significant asteroidal population, both in number and also in relation to their dynamical and physical properties. They form a key population for revealing the history of the Solar System since its existence and survival constrains the theories of formation and evolution of the Solar System as a whole (e.g. Marzari and Scholl, 1998; Morbidelli et al., 2005; Nesvorn´y et al., 2013). A distinctive feature of this population is an asymmetry in the number of bodies of the leading and trailing clouds. It is observed that the number of Trojans in L4 doubles the number in L5. However, dynamical studies of the Trojan region show that both L4 and L5 have the same structure and stability. It is unclear whether the observations could be biased. For example, Grav et al. (2011) estimate that the real asymmetry between L4 and L5 should be corrected to a factor 1.4. Nevertheless, even after correcting for biases, this asymmetry seems to be real and deserves attention. The general dynamical properties of Trojans are also key to understanding the Solar System process. They have been broadly studied in the past, both analytically and numerically (e.g. Erdi, 1996; Mikkola and Innanen, 1992; Milani, 1993). The first long-term numerical integration of Jupiter 2 Trojans was made by Levison et al. (1997). They numerically integrated the orbits of 270 fictitious L4 Trojans for 1 Gyr and also the orbits of 36 real Trojans for 4 Gyr detecting stability areas and the places occupied by the real Trojans. They also followed the evolution of the escaped Trojans and studied their relation with Jupiter-family comets (JFCs). A semi-analytical model to describe the long-term motion was developed by Beaug´e and Roig (2001) where they identified and confirmed the existence of the majority of the families previously detected. Thanks to the increasing computing power, long-term numerical simu- lations with a larger number of particles are now possible. Recently, new long-term numerical integrations were carried out for the Trojans, which al- lowed us to deeper characterize the dynamics within the resonance. For ex- ample, Marzari and Scholl (2000, 2002) and Marzari et al. (2003) performed a series of numerical simulations to study the role of secular resonances in the dynamical evolution of Trojans and explored their stability properties and destabilization mechanisms. They found that direct perturbations made by Saturn are the main source of instability on time scales of the order of 107 − 108 years, while secular resonances, in particular, ν16 contribute on longer timescales. This secular resonance raises the inclinations up to values greater than 20◦ on a time scale of 108 years. More recently, Robutel et al. (2005) and Robutel and Gabern (2006) performed long-term numerical simu- lations to study the global dynamical structure of the L4 region. They found that the inherent instability of the Trojans appears purely gravitational and caused by secondary and secular resonances within the tadpole regions. Tsiganis et al. (2005) studied the stability of Trojans to define the ef- fective stability of the region and compare it with the real distribution of Trojans. The effective stability region is defined in terms of the Lyapunov time and the escape time (time for an encounter with Jupiter) to study the regular and chaotic orbits at the border of the stability region. These orbits remain for the age of the Solar System. These authors numerically inte- grated real and fictitious L4 Trojans finding that 17% of the real (numbered) Trojans escaped from the swarm over the age of the Solar System and that chaotic diffusion is the origin of the unstable population. Studies of the physical properties of Trojan asteroids suggest that they contain water ice and organic material, similar to the cometary nuclei. Spec- troscopic studies derive mainly D taxonomic classes but also some P and C classes (Fornasier et al., 2007; Emery et al., 2011; Grav et al., 2012). Wa- ter ice content and their similarity with cometary nuclei seems to indicate 3 that Trojans could be formed in the outer Solar System. However, Emery et al. (2011) found two spectral groups which they attributed to different intrinsic compositions and suggested two distinct regions of origin. There are several studies of albedos that show values in the range from 0.025 to 0.2 (e.g Grav et al., 2011; Grav et al., 2012; Fern´andez et al., 2003) with a possible correlation between albedo and size, presenting lower values for smaller Trojans (Fern´andez et al., 2009). There are two space missions that plan to study in detail the physical properties of Trojans that will radically improve what we know about Trojans. "Lucy" NASA mission, which will be launched in 2021, will encounter one Main Asteroid Belt and six Trojans from both swarms after a 12-year journey. Its main objective is to study the Trojan surface compositions, the diversity of taxonomic classes and also the interior and bulk properties to link those results with the source Trojan regions. The Japan Aerospace Exploration Agency, JAXA, is planning a mission to Trojans based on propulsion by a solar power sail (already tested by IKAROS, the first deep space solar sail). They plan to arrive on a Trojan target for global remote observation, surface and sub-surface sampling by a lander, and a possible sample return option. In particular, Di Sisto et al. Although Trojan asteroids are librating about the Jupiter's L4 and L5 stable equilibrium points, there exist regions of partial instability from which Trojans can escape from the resonance (Levison et al., 1997; Di Sisto et al., 2014). (2014), hereinafter D14, studied the dynamical evolution of Jupiter Trojans by numerical integrations of observed Trojans under the gravitational action of the Sun and the four giant planets. They focused their study on the properties of the observed population and the escape/survive population. They found that the escape rate of L5 Trojans is greater than that of L4, and this fact could be responsible for ∼10 % of the total asymmetry. We will discuss those results later. In this paper, through numerical simulations, we study the escape of Tro- jan asteroids and follow their dynamical evolution over the age of the Solar System or until they physically collide with a planet or they completely es- cape the Solar System. The main objective is to evaluate the dynamical routes of escape and the contribution of Trojans to other minor body popu- lations, such as Comets, Centaurs, and NEOs. Also, we show the temporary captures of escaped Trojans in MMR, both inside and outside Jupiter's orbit. The paper is organized as follows. In Section 2, we review the main physical and dynamical characteristics of the Trojans, focusing on their number and size distribution. In Section 3, we describe the numerical simulations, while 4 in Section 4, we present the results of a long-term integration of the observed Trojans. The Trojan contribution to other small body populations is shown in Section 5. We finally present our discussion and conclusion in Section 6. 2. The Observed Population 2.1. Physical properties and size distribution Trojan asteroid spectral features are similar to those of cometary nuclei. Spectroscopic studies derive mainly D taxonomic classes and some P and C (Fornasier et al., 2007; Emery et al., 2011; Grav et al., 2012). Grav et al. (2011) derived thermal models for 1739 Jovian Trojans, observed by the WISE survey (Mainzer et al., 2011), and detected no differences for the leading and trailing cloud. They also found that the size distributions of the two swarms are very similar. Later, by recomputing thermal model fits derived from that sample, Grav et al. (2012) calculated visible albedos that vary from 0.025 to 0.2 for small Trojans, with a median value of 0.05 for D > 30 km and 0.07 for D < 30 km. In this paper we will adopt those results from Grav et al. (2011), Grav et al. (2012) since they are based on the largest sample of albedo measurements. The size distribution of Jovian Trojans has been studied from observa- tional surveys and measurements of albedos. Jewitt et al. (2000) carried out a survey in the L4 direction, detecting 93 Trojans with diameters of 4 km < D < 40 km, and obtained a Trojan cumulative size distribution (CSD) as a power law given by N(> D) ∝ D−s with an index s = 2.0 ± 0.3. But, by adding cataloged Trojans to the sample, they inferred that there must be a break in the CSD at diameters D ∼ 80 km toward an index s = 4.5. New surveys for small Trojans performed by Yoshida and Nakamura (2005), Yoshida and Nakamura (2008) found that the faint end of the CSD seems to have another break around D ∼ 4 − 5 km. Szab´o et al. (2007) analyzed the observations of more than 1000 Trojans and found that the CSD of L4 and L5 are virtually the same with a cumulative index s = 2.2 in the range 10 km . D . 80 km, but there are 1.6± 0.1 more objects in the leading swarm than in the trailing one. A new and complete analysis of the magnitude distribution of L4 Trojans has been recently done by Wong and Brown (2015). From a Subaru survey, they detected 557 small L4 Trojans, and, by combining these observations with the bright Trojans contained in the MPC catalog, they fit a complete magnitude distribution in 5 the range 7.2 < H < 16.4 given by: Σ(H) =   10α0(H−H0), 10α1(H−Hb′ )10α0(H b′ −H0), 10α2(H−Hb)10α1(H b−Hb′ )10α0(H b′ −H0), H0 ≤ H ≤ Hb′ Hb′ ≤ H ≤ Hb H ≥ Hb (1) The magnitude distribution begins at H0 = 7.22 and has two breaks at magnitudes Hb′ = 8.46 and Hb = 14.93. In the three regions defined by those limiting magnitudes the slopes are α0 = 0.91, α1 = 0.44 and α2 = 0.36 From a power-law magnitude distribution of the form of Eq. (1), the radii of Trojans follow a differential size distribution (DSD) given by N(R)dR = CR−qdr, where q = 5α + 1, and C is a constant. By converting Eq. (1) to the DSD in each region and integrating them, we can obtain the CSD of L4 Trojans as: N(> R) =   (R/R0)1−q0, (R/Rb′)1−q1(Rb′/R0)1−q0, (R/Rb)1−q2(Rb/Rb′)1−q1(Rb′/R0)1−q0, Rb′ ≤ R ≤ R0 Rb ≤ R ≤ Rb′ R ≤ Rb (2) where q0 = 5.55, q1 = 3.2 and q2 = 2.8. To convert magnitude to radius, we consider the results of Grav et al. (2012) that obtained an albedo equal to 0.05 for D > 30 km and 0.07 for D < 30 km. So, the limiting radii are R0 = 106.9 km, Rb′ = 60.4 km and Rb = 2.6 km, We can see that the indexes of the CSD in the different ranges of radius, and also the location of the breaks, are in agreement with the previous stud- ies. Then, we will adopt Eq. (2) as the CSD of L4 Trojans. There is no comprehensive study on the CSD of L5 Trojans, but some studies obtained very small (or no) differences in the CSD of L4 and L5 (Szab´o et al., 2007; Yoshida and Nakamura, 2008). Then, based on the study of Grav et al. (2011), we will consider that the number of L4 Trojans is 1.4 the number of L5 Trojans, and that this asymmetry does not depend on the size. There- fore, the CSD of L5 Trojans is given by Eq. (2) but offset by the asymmetry factor. Both CSDs are plotted in Fig. 1. Then, for example, there would be ∼ 265000 L4 Trojans and ∼ 190000 L5 Trojans with diameter greater than 1 km. 2.2. Dynamical properties An analysis of the observed population was made by Di Sisto et al. (2014), who found some differences in L4 and L5 swarms. While the mean values of 6 1x107 1x106 100000 10000 ) R > ( N 1000 100 10 1 0.1 0.1 L4 L5 1 10 R [km] 100 Figure 1: Cumulative size distribution of L4 (black) and L5 (red) Trojans. the semimajor axis and eccentricities are almost the same for L4 and L5, the mean inclination in L5 is 4◦ greater than that of L4. The relatively more ex- citation of the L5 population is also appreciated in the distribution of inclina- tions, which is broader than that of L4. Another point addressed by D14 what is related to the observed population is the calculation of proper elements and the determination of family members. They worked with numbered and multioppositional Trojans and concluded that only numbered asteroids have sufficiently well determined orbits to allow for detailed and long-term dynamical analysis. 3. The Numerical Simulation The initial conditions of our simulations are the orbits of all numbered Jupiter Trojan asteroids as of March 2013. Thereby, a numerical integration of 1975 L4 Trojans and 997 L5 Trojans were performed under the gravita- tional influence of the Sun and the planets from Venus to Neptune with the hybrid integrator EVORB (Fern´andez et al., 2002). The time step was set to 7.3 days, which is roughly 1/30 of Venus orbital period, and each Trojan 7 evolved for 4.5 Gyr, unless removed due to a collision with a planet or the Sun, or due to reaching a heliocentric distance r > 1000 au. The encounters at less than 2.1 Hill's Radius with the planets were registered to analyze them and to define an "escape time" for each "escaped Trojan". If a Trojan has an encounter with a planet, usually Jupiter, the time of the first encounter is considered the "escape time" and the Trojan would be an "escaped Trojan"; its subsequent evolution through the Solar System up to a collision or escape will be the objective of this paper. The initial orbital elements of all L4 and L5 Trojans are shown in Fig. 2. The black points represent the orbits of stable Trojans for 4.5 Gyr while escaped Trojans are represented in red. Given the small step used in the integrations, the number of particles, and the long time interval, the initial orbits were divided in groups of nearly 20 Trojans. They were integrated by using several computers under the same conditions for several months. We performed a total of 48 runs for L5 Trojans and 108 for L4 ones. 4. General Results 4.1. Escape from L4 and L5 We detect 466 (out of 1975) Trojans that escape from L4, this is 23.6%, and 250 (out of 997) Trojans that escape from L5, i.e. 25.1 %. Jupiter is the main cause for the escapes from both swarms, however, we have 1 L4 Trojan that has its first encounter with Mars and 5 L4 Trojans and 3 L5 Trojans with Saturn. The analysis of the dynamical evolution of these particular Trojans that encounter other planet than Jupiter first, reveals that a slow diffusion among resonances is at work before the escape, as already noted by Robutel and Gabern (2006). It is possible to see the typical behavior of objects going through secular and secondary resonances, slowly increasing the eccentricity and changing the inclination, which eventually favors close encounters with the planets. This behavior is also found in some other Trojans that encounter Jupiter first. However, the drastic change in the semimajor axis from which the escape of the resonance can be detected occurs after the encounter. The number of escapees from L5 is proportionally greater than that from L4, in agreement with the results of D14, although the difference is smaller. We find that the escape rate from both Lagrangian points follow a linear trend with the time given by aL4 = 7.0398 × 10−11 ± 8 × 10−14 and aL5 = 7.5590 × 10−11 ± 13 × 10−14. 8 ] g e d [ i 40 30 20 10 0 0.25 0.2 e 0.15 0.1 0.05 0 5 L4 L5 5.1 5.2 a [AU] 5.3 5 5.1 5.2 a [AU] 5.3 5.4 ] g e d [ i 40 30 20 10 0 0 0.05 0.1 0.15 0.2 0.25 0 0.05 0.1 0.15 0.2 0.25 0.3 e e the initial conditions of the Figure 2: Orbital elements of the numbered Trojans, i.e. simulation. Black points represent the Trojan that are stable for 4.5 Gyr and the red ones those that escape from the swarms. Following the same analysis as in D14, if the present unbiased asymmetry in the number of Trojans between L4 and L5 is Ns(L4)/Ns(L5) = 1.4 ± 0.2 (Grav et al., 2011), the original population of Trojans would have a 9 primordial asymmetry of N0(L4)/N0(L5) = 1.373 ± 0.204. (3) Therefore, the difference in the escape rate between L5 and L4, accounts for only ∼ 2% of the total asymmetry or, in other words, it has contributed to ∼ 7% of the present unbiased asymmetry. 4.2. Post-escape The evolution of escaped Trojans was followed up to a collision with a planet or the Sun, or until a heliocentric distance r > 1000 au (ejection was reached). From the 250 Trojans that escape from L5, 2 of them end their evolution due to a collision with the Sun, 1 with Jupiter and the remaining objects are ejected. From the 466 escaped L4 Trojans, 16 of them collide with an object: 2 with the Sun, 2 with Saturn and 12 with Jupiter, and the remaining 450 are ejected. The different intrinsic rate of collisions between both escapees is remarkable; the proportional number of collisions by L4 escapees is three times greater than that of L5 escapees. In Table 1, we show the fraction of escaped Trojans that have at least one encounter with each planet (NT ). From this set, we additionally calculate the quantity Ne, which is the fraction of encounters with each planet with respect to the total number of encounters. For example, 466 Trojans escape from L4, from which only 28 have at least one encounter with Venus, this is, NT = 6%. In addition, these 466 Trojans sum 229906 planetary encounters, from which 151 correspond to encounters with Venus, i.e. Ne = 0.07%. For L5, the 250 Trojans that escape undergo 131324 planetary encounters. We have found that most of the encounters occur with Jupiter and Saturn. Besides, the Trojans departing from L5 have a larger Ne for the inner Solar System including Jupiter than those departing from L4. The opposite is observed for the outer Solar System. The whole evolution of escaped Trojans out of the swarms can be seen in Fig. 3. Those plots show the normalized time fraction spent by escaped Trojans in the orbital element space. The color code is indicative of the permanence time spent in each zone (blue for the most visited regions, red for the least visited). Then, those plots form dynamical maps of "permanence" in the different zones of the Solar System and give a general idea of the regions visited by escaped Trojans. The regions of the Solar System occupied by escaped Trojans from L4 and L5 are similar but not equal. They cover similar ranges of orbital elements; 10 Table 1: Percentage of the escaped Trojans that have at least one encounter with each planet (NT ), and percentage of encounters with each planet with respect to the total number of planetary encounters (Ne) for L4 and L5 Lagrangian points. Planet Venus Earth Mars Jupiter Saturn Uranus Neptune L4 L5 Ne(%) NT (%) Ne(%) NT (%) 13 19 13 100 95 83 76 0.12 0.2 0.08 77.7 17.2 2.6 2.1 0.07 0.15 0.08 63.3 21.9 6.2 8.3 6 10 9 100 97 85 76 however, differences in semimajor axes and inclinations can be observed in Fig. 3. L4 escaped Trojans cover a wider range of semimajor axes than L5 escapees. The inner regions of Jupiter's orbit are preferred by L5 escaped Trojans; in Fig. 3, that zone is most visited by L5 escapees than L4 ones (i.e. there are blue strips in L5 maps but not in L4 maps). In this region, the densest zones correspond to the region near the 1 : 1 mean motion, the Hilda region and the outer zone of the asteroid Main Belt. In the region outside Jupiter's orbit, L5 escapees have perihelion distances near Saturn and there is a small structure near Neptune's perihelion whereas L4 escapees cover almost all the external region with perihelion near all the giant planets. There is also an island visited by L4 escapees with a semimajor axes between 40 and 100 au, and an inclination between 30 and 60 degrees. This structure is generated by large variations in inclination and eccentricity due to the Kozai mechanism inside exterior MMRs with Neptune. We identified objects following this dynamic up to resonances 1 : 5 with Neptune at a = 88.07 au and 1 : 6 with Neptune at a = 99.45 au. To detect captures in MMR, we compute the mean orbital elements of the escaped Trojans by means of a running window of 104 years every 103 years, according to the following formula: < E(t) >= 10−4 R t+5000 t−5000 E(t′)dt′, where E is the orbital element and 103 years is equivalent to one orbital state. We compute 1644535 (173937) total orbital states calculated for the 466 (250) escaped Trojans from the L4 (L5) point. Figures 4 and 5 show the orbital states up to 10 au in the (< a >, < e >) and (< a >, < i >) planes. 11 Figure 3: Normalized time-weighted distribution of the dynamical evolution of escaped L4 and L5 Trojans in the (a, e) plane (left) and (a, i) plane (right). The color zones of these maps are regions with different degrees of probability where escaped Trojans can be found (blue for most visited regions, red for least visited). The black curves correspond to constant perihelion values equal to the location of the giant planets. For a high resolution image, ask the authors. Several concentrations for the mean semimajor axis around nominal values of MMRs are observed. For some resonances, the aphelion of the escaped Trojans decouples from Jupiter's orbit and evolves to an inner region far from the curve of constant aphelion with Jupiter. A similar behavior was found by Fern´andez et al. (2018) for active Centaurs. In particular, some escaped Trojans are temporarily captured in the exterior 2 : 3 MMR with Jupiter and 1 : 1 with Saturn. In Table 2, we show the number of orbital states in the most important resonances. We also calculate the number of escaped Trojans that remain in resonance for more than 20000 (N20) and 100000 (N100) years. It is remarkable that the most populated resonances are first the co-orbital with Jupiter and then the co-orbital with Saturn. For the resonance 3 : 2, we compared the orbital properties of the Hildas with the orbital parameters of the objects temporarily captured in the reso- 12 nance 3 : 2. Fig. 6 shows the regions in the space (i, e) of the actual Hildas and the particles captured in the resonance 3 : 2. The escaped Trojans evolve to the population of the Hildas but in general with very large eccentricity and inclination. Nevertheless, some orbital states are perfectly compatible with the actual Hildas. Figure 4: Mean orbital elements of escaped Trojans from L4. Small arrows indicate the location of resonances. The points below the continuous curve correspond to orbital states completely inside the orbit of Jupiter. All the features observed in the orbital element distribution of escaped Trojans could be related to the different temporal evolution of both swarms. L5 escapees have shorter lifetimes than L4 ones as can be seen in Fig. 7, where the normalized distribution of lifetimes is plotted. We can also see that there are a few L4 escapees that reach lifetimes greater than 100 Myr while there are no L5 Trojans with these lifetime values. Another way of analysing see the differences in the temporal evolution 13 Figure 5: The same as Fig. 4 for escaped Trojans from L5. of escaped Trojans is shown in Fig. 8, where the mean lifetime versus the semimajor axis is plotted. We can see that for a < 10 au, the L5 Trojans have a greater lifetime whereas this is reversed for a > 10 au and the difference grows up and become significant. In particular, for 15 au< a < 20 au, the difference is notable. This general behavior is intrinsic to the different evolution of L5 and L4 escapees, i.e. L5 preferred the inner Solar System zones while L4 escapees preferred the outer ones, as can be seen in Fig. 3. The mean lifetime of escaped L5 Trojans in the Solar System up to ejec- tion or collision is 0.7 Myr while that of L4 is 3.5 Myr, i.e. five times greater. However, the different behavior of L4 and L5 escapees is biased by a statis- tic of few objects. Then, to characterize the temporal evolution of escaped Trojans, we chose to evaluate the "median lifetime" as a typical lifetime of Trojans outside the resonance. In this case, we obtain that the median life- time for L5 escapees is 264000 years, whereas that of L4 escapees is 249000, 14 ✂ ✁✡ ✁✠ ✁✟ ✁✞ ✁✝ ✁✆ ✁☎ ✁✄ ✁✂ ✙✚✖✕ ✛✒✕✜✖✢ ✣✆ ✛✒✕✜✖✢ ✣✝ ✛✒✕✜✖✢ ✂ ✄ ☎ ✆ ✝ ✞ ✟ ✠ ✒✓✔✕✒✓✖✗✒✘✓ Figure 6: Orbital elements for the Hildas and for the captured particles in the 3 : 2 resonance. i.e. both of the same order. To test if the results for the escaped Trojans from L4 and L5 are rep- resentative of the real behavior of these populations, we take proportional samples of the initial populations, extracting 20% of the objects from each sample. For each subsample, composed by 80% of the original sample, we performed all the calculations and statistics again. We repeated this proce- dure a few times. These experiments allowed us to better estimate the errors in our results. Since we obtain similar results for the different samples, we ca confirm that: • The proportion of escaped Trojans from L5 is 25.1%, which is slightly greater than that of L4 of 23.6%, with an error of 0.2%. • The difference in the escape rate is also significant, though small, i.e. aL4 = 7.04× 10−11± 1× 10−12 and aL5 = 7.56× 10−11± 1× 10−13. This implies a primordial asymmetry of N0(L4)/N0(L5) = 1.376 ± 0.204. • The encounters with the planets follow the same trend in all experi- ments. 15 ☛ ☞ ☞ ☛ ✌ ✍ ✎ ✏ ☞ ✏ ✍ ✑ Table 2: The number of orbital states (see text) in the most important resonances, and the number of escaped Trojans that remain in resonance for more than 20000 years (N20) and 100000 years (N100) for L4 and L5 Lagrangian points. escaped L4 escaped L5 Resonance a [au] 2.82 5:2 9:4 3.03 3.58 7:4 3.97 3:2 4.29 4:3 1:1 5.20 6.82 2:3 1:1S 9.55 states 1543 464 358 510 936 6467 516 1594 N20 N100 3 2 1 1 3 18 1 1 5 5 11 18 15 60 8 48 states 144 796 787 983 340 2535 413 852 N20 N100 1 2 1 2 1 8 1 2 2 7 11 14 7 46 8 19 L4 L5 0.7 0.6 0.5 0.4 0.3 0.2 0.1 n o i t c a r F 0 10000 100000 1x106 1x107 1x108 1x109 lifetime [yr] Figure 7: Normalized distribution of mean lifetimes of L4 (black) and L5 (red) escaped Trojans. • The mean lifetime of escaped L4 Trojans is greater than that of L5 due to the fact that a few L4 Trojans have very long lifetime after escape. 16 ] r y K [ e m i t e f i l 140 130 120 110 100 90 80 70 60 50 40 30 L4 L5 0 5 10 15 a [au] 20 25 30 Figure 8: Mean lifetime of L4 (black) and L5 (red) escaped Trojans vs semimajor axis. • The median lifetime of both escapees is of the same order. The different post-escape behaviour of L4 and L5 escapees can also be analytically investigated using the Opik theory (Opik, 1976; Valsecchi et al., 2000). The Opik method analyzes an encounter of a particle with a planet and gives a mean probability of collision with a planet and other parameters for an orbit with a given (a,e,i). Then, following Valsecchi et al. (2000) we consider the mean initial orbital elements of L4 and L5 escaped Trojans and calculate the relative encounter velocity, the impact parameter, the probability of collision, the mean time after which the object collides with the planet and the extreme changes in semimajor axis due to the encounter. We consider planet Jupiter for the calculations. Those values for both Trojan swarms are shown in Table 3. As can be seen, we obtain that the collision probability of L4 escapees is greater than for L5 escapees. This is related to the different initial inclination of Trojans in boths swarms. From Opik theory: Pcol = , Uσ2 Uxπsini 17 (4) Table 3: Opik theory for escaped Trojans. The input orbital elements (a, e, i) correspond to the mean initial orbital elements of L4 and L5 escaped Trojans. For those values, the relative encounter velocity (U), the impact parameter (σ), the probability of collision per orbital period (Pcol), the mean time after which the object collides with Jupiter(t(col)) and the extreme changes in semimajor axis due to the encounter (amax, amin) are shown. a [au] e i [degrees] U σ [Rp] Pcol t(col) [My] amax [au] amin [au] L4 5.2058 0.0842 10.6828 0.204 22.615 L5 5.2118 0.0813 14.464 0.264 17.473 1.717×10−5 0.692 9.457 3.807 1.019×10−5 1.167 12.958 3.567 where Ux = ±p2 − 1/a − a(1 − e2)) and [a] = aj. Then, since mean a and e are almost equal for both trojan swarms, the only variable that affects the result of an encounter is the different mean initial inclination. Orbits with greater initial inclinations give lower Tisserand constant (T ) and then greater relative velocities (U = √3 − T ). This is also transferred to the impact parameter σ and all together to the probability of collision, as can be seen from Eq. 4. This is consistent with our numerical results of the rate of collision of escaped Trojans with the planets, especially with Jupiter, i.e., the proportional number of collisions by L4 escapees is three times greater than that of L5 escapees. Another result from Opik theory shows that L5 escapees have greater changes in semimajor axis than L4 escapees and also higher speeds of encounter. In particular, L5 Trojans go further than L4 Trojans after an encounter. We think that this fact, together with L5s' higher speed of encounter, could make their evolution faster with respect to the L4 ones. That is, L5 Trojans would go further than the L4 Trojans in each encounter and therefore spread out faster. Then, they would have proportionally fewer encounters with the planets beyond Jupiter than the L4s (see Table 1) and their mean lifetime outside Jupiter would be lower than that of L4 escapees in this zone. This could explain the shorter mean lifetimes of L5 escapees with respect to the L4 ones, found in our simulation. Also, this could explain the preference for L4 escapees that almost cover all the external region with 18 perihelion near all the giant planets, in contrast to L5 escapees. Besides, the proportionally slower evolution of L4 escapees through the external region would allow the existence of very long-lived escaped Trojans as the few L4 escapees that reach lifetimes greater than 100 Myr, while there are no L5 Trojans with these lifetime values (see Fig. 7). We think that the application of Opik theory could help to explain those characteristics in the evolution of escaped Trojans. However, it has to be regarded as an approximation since it does not take into account other perturbers than the planet considered. 5. Contribution to other minor body populations To calculate the contribution of escaped Trojans to other minor body populations, we will define: • Comets: q < 5.2 au. -- JFCs: P < 20 yr and 2 < T < 3.15. -- no-JFCs: P > 20 yr or T < 2 or T > 3.15. • Centaurs: q > 5.2 au and a < 30 au. • TNOs: a > 30 au, where q is the perihelion distance, a is the semimajor axis, T is the Tisserand parameter with respect to Jupiter, and P is the period. We have considered the usual definition of JFCs but extended the limit of the Tisserand parameter to 3.15 taking into account that in fact, Jupiter's orbit is slightly elliptic and then, orbits with T values slightly above three allow close encounters with Jupiter (Di Sisto et al., 2009). The above defini- tions were taken in such a way that there is no overlap of populations. We have, according to perihelion distance, a comet population inside Jupiter's orbit, a Centaur population in the giant planet zone with semimajor axis inside Neptune's orbit, and a TN population beyond Neptune. The contribution of escaped Trojans to each of the above defined popula- tions is shown in Table 4. We can see that almost all escaped Trojans from L4 and L5 reach the comet's zone and ∼ 90% go through the Centaur zone. The proportion of contribution from L4 and L5 is similar in those zones, though the mean lifetime of L4 Trojans in a JFC zone is slightly smaller than that 19 Table 4: Percentages of escaped Trojans (with respect to the number of escapes) and mean lifetime (τ ) in each population. Population JFCs no-JFCs Centaurs TNOs N(%) 97 93 90 78 L4 τ [yr] 72300 87500 420000 3.9 × 106 L5 τ [yr] 88400 80900 272000 400000 N(%) 97 96 84 76 of L5 Trojans, and the opposite occurs in the Centaur and TN zones. This last topic is connected with what we have already mentioned about the 5 escaped L4 Trojans that have long lifetimes in the Centaur and TN region. The orbital state distribution of escaped Trojans in the JFC and Centaur zones are plotted in Figs. 9 and 10. The distribution of observed JFCs and Centaurs are also shown. The data were obtained from the JPL Small- Body Database Search Engine in September 2017. Fragments of disrupted comets were removed, leaving only one data point for each parent comet. To test the contribution of Trojans to JFCs, we consider the intrinsic real distribution of JFCs, i.e. a non-biased sample. We follow the reasoning proposed by Nesvorn´y et al. (2017) to select an unbiased sample of JFCs, and extract those JFCs with perihelion distances q < 2.5 au and absolute total magnitude HT < 10. We have 63 known JFCs that satisfy these criteria. The comparison of the orbital element distribution of escaped Trojans and this "complete" sample is shown in Fig. 9. We can see that the semimajor axis and aphelion distances are reasonably well fitted, although there are differences. However, escaped Trojans reach smaller perihelion distances than JFCs. Eccentricities and inclinations of Trojans are higher than those of JFC ones and the argument of perihelion shows the same typical distribution of a population dominated by encounters with Jupiter as the JFC one. In Fig. 10 the orbital state distributions of escaped Trojans in the Centaur zone and known Centaurs are shown. We can see that escaped Trojans are compatible with observed Centaurs. In fact, the distribution of both spatial orbital elements and angular ones of escaped Trojans are similar to the observed distribution. Only Centaurs in low eccentricity and low inclination orbits are not compatible with escaped Trojans. Two peaks are noticed in the inclination distribution of L4 escaped Trojans near 40◦ and 60◦. They 20 n o i t c a r F n o i t c a r F 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 0 1 2 3 4 5 6 7 8 a [au] 0 0.2 0.4 0.6 0.8 1 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.5 1 1.5 2 2.5 q [au] 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 0 10 20 30 40 50 60 70 80 90 0.25 0.2 0.15 0.1 0.05 0 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 0 2 4 6 8 Q [au] 10 12 14 0 50 100 150 200 250 300 350 e i [deg] w [deg] Figure 9: Distribution of orbital states of escaped Trojans (L4 (black) and L5 (red)) in the JFC zone with q < 2.5 au. The blue line corresponds to the observed "complete" sample of JFCs (q < 2.5 au and total absolute magnitude HT < 10). correspond to a few Trojans that remain for a long time on low eccentricity and high inclination orbits, in some cases due to Kozai mechanism inside MMRs. The previous analysis corresponds to our numerical simulation, which has aimed at evolving known Trojans. To evaluate the real contribution from Trojans, we have to take into account the real asymmetry between L4 and L5 as well as the number of Trojans in each swarm given by Eq. (2). Then, the number of Trojans ejected out of the resonance per year will be given by: Nejec(> R) = aLiN(> R). (5) For example, there will be 5 L4 Trojans and 4 L5 Trojans with radius R > 1 km ejected from the resonance every Myr. Or, for diameter D > 1 km, 2 L4 Trojans are ejected every 100000 yrs and 3 L5 Trojans every 200000 yrs. The number of escaped Trojans in each minor body population would be given by: and it is plotted in Fig. 11. We have added the contribution from both Nop(> R) = aLi N(> R) τ, (6) 21 0.1 0.09 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 0.14 0.12 0.1 0.08 0.06 0.04 0.02 n o i t c a r F n o i t c a r F 5 10 15 20 25 30 a [au] 0 0 0.2 0.4 0.6 0.8 1 0.12 0.1 0.08 0.06 0.04 0.02 0 5 10 15 20 25 30 q [au] 0.12 0.1 0.08 0.06 0.04 0.02 0 0 20 40 60 80 100 120 140 160 180 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 5 10 15 20 25 30 35 40 45 50 Q [au] 0 50 100 150 200 250 300 350 e i [deg] w [deg] Figure 10: Distribution of orbital states of escaped Trojans (L4 (black) and L5 (red)) in the Centaur zone. The blue line corresponds to the observed Centaurs obtained from the JPL Small-Body Database Search Engine. swarms and also the JFCs and no-JFCs into comets in the plot. However, we have noticed that considering the real asymmetry and the real number of Trojans given by Eq. (2), the contribution to JFCs and no-JFCs from L4 and L5 is almost the same, but the L4 contribution to Centaurs is two times greater than that of L5, and the L4 contribution to TNOs is ∼ 15 times that of L5. This difference, however, is based mostly in the five L4 Trojans that have long lifetime in the Centaur and TNO regions and then it should be taken with caution because of a low number statistic. Beyond this, the contribution of escaped Trojans to Comets, Centaurs and TNOs is negligible. In fact, for example, Di Sisto et al. (2009) estimate that there would be ∼ 450 JFCs with R > 1 km within Jupiter's orbit, and the contribution of Trojans would be only 1. In the case of Centaurs and TNOs, the number of Trojans with R > 1 km would be 3 and 20 respectively, which is orders of magnitude lower than the total number of both populations. Nevertheless, small Trojans in cometary orbits could resemble Asteroids in Cometary Orbits (ACOs). For example, we have ∼ 6 Trojans in cometary orbits with D > 1 km. Tancredi (2014) identified 203 ACOs belonging to the Jupiter-family group, thus, the contribution from Trojans would be minor. We have also tested whether there are Trojans in the NEO population 22 Comets Centaurs TNOs 1000 100 ) R > ( p o N 10 1 0.1 1 R [km] Figure 11: Cumulative number of escaped Trojans in each minor body population. (q < 1.3 au), and have we obtained that this contribution is also negligible. We could expect 2 escaped Trojans with D > 1 km in the NEO population with dynamical lifetimes of ∼ 30000 yrs. Almost all of those escaped Trojans with q < 1.3 au are in a JFC-orbit. Fern´andez and Sosa (2015) carried out orbital integrations of JFCs in NEO orbits and obtained that there is a fraction of them that show stable asteroidal orbits with lifetimes greater than 10000 yrs. They obtained that at least 8 JFCs in NEO orbits show this stable behavior and being them km- and sub-km size bodies, they attribute their long lifetime to a mostly rocky composition and might have a source region in the outer main asteroid belt. So, a fraction of those objects could come from the Trojan swarms. It is interesting that a fraction of escaped Trojans go through the zone of Encke type comets (ETC), defined as those with T > 3 and a < aJ up. We could expect 1 ETC with D > 1 km in this zone. However, the exact orbital elements of 2P/Encke are not reached by escaped Trojans. Another interesting case is the Shoemaker Levy 9 (SL9) impact with Jupiter. In our simulations we have a great number of collisions of escaped Trojans with Jupiter. Also, the number of close encounters with Jupiter 23 within the Roche limit is important: we have 5 L5 escaped Trojans and 17 L4 escaped Trojans that encounter Jupiter within its Roche limit. Within this radial distance to the planet, an object could be fragmented and end up impacting with the planet, as did the SL9. The rate of those encounters is, in fact, constant and then we could expect, for example, 5 "SL9 case"- Trojans with D > 1 km from L4 and 2 from L5 every 10 Myr, or in total one escaped Trojan with D > 1 km every 1.4 Myr would cross the Roche limit of Jupiter and then it would fragment and end up impacting Jupiter. Di Sisto et al. (2005) found that one escaped Hilda asteroid with D > 1 km would impact Jupiter every 65000 yrs, being this rate of collision much greater than our estimated rate for escaped Trojans. 6. Discussion and Conclusions Trojan asteroids are located in stable reservoirs and have long dynamical lifetimes. However, some of them are located in unstable zones and then are capable to escape from the swarms. In this paper, we have analyzed the observed L4 and L5 Trojan population trough numerical simulations in order to study the dynamical behavior of the escaped Trojans and their contribution to other minor body populations. We obtain that the number of Trojans that escape from L5 in the age of the Solar System is proportionally greater than that from L4. The difference is small, we have 25.1 ± 0.2% of Trojans from L5 and 23.6 ± 0.2% from L4. The escape rate from both swarms along the integration time can be fitted by a linear relation and the one from L5 is greater than that from L4 over time. Those results are qualitatively similar to D14. The dynamical evolution of escaped Trojans was studied up to their ejec- tion of the Solar System or collision with the Sun or a planet. The main general results of the simulation are the following: • The proportional number of collisions by L4 escapees is three times greater than that of L5 escapees, being the great majority of them with Jupiter although few collisions with the Sun and Saturn are also registered. This is a consequence of the different mean inclination of Trojans in both swarms, as we have demonstrated using Opik theory. • Most of the encounters are with Jupiter, but there are also with other planets. We found a greater relative proportion of encounters by L5 24 Trojans with Jupiter and the inner planets with respect to L4 Trojans, and the reverse is observed for the outer planets (Saturn, Uranus and Neptune). • L4 and L5 escaped Trojans cover similar regions of orbital elements; however, L5 escapees preferred the regions interior to Jupiter's orbit and L4 escapees cover almost all the external region with perihelion near all the giant planets. • L5 Trojans spent 0.7 Myr up to ejection or collision whereas L4 spent 3.5 Myr. But this difference is mainly due to five L4 Trojans that have long dynamical lifetime (i.e. > 100 Myr) once they escape, slowly evolving in the Centaur and TNO regions. A better characterization of the temporal evolution of escaped Trojans is then the "median lifetime" that is of the same order for both escapees, i.e. 264000 yrs for L4 and 249000 yrs for L5 escapees. The contribution of Trojans to other minor body populations was an- alyzed. We found that almost all escaped Trojans from L4 and L5 reach the comet's zone, ∼ 90% go through the Centaur zone and only L4 Trojans reach the transneptunian zone. In particular, we note that the distribution of both spatial orbital elements and angular ones of escaped Trojans are similar to the observed Centaur orbital element distribution. Considering the real asymmetry between L4 and L5 and the number of Trojans in each swarm given by Eq. (2), we obtained the number of Trojans ejected out of the resonance per year. Then, for example, there are 5 L4 Trojans and 4 L5 Trojans with radius R > 1 km ejected from the resonance every Myr. Or, for diameter D > 1 km, 18 L4 Trojans are ejected every Myr and 14 L5 Trojans every Myr. The results of the present paper are based on the dynamical evolution of Trojans, and then, the escape rate from the swarms is due only to dynamical instabilities on the resonance. However, the collisional evolu- tion of Trojans could be responsible for the escape of some of them, too. In fact, de El´ıa and Brunini (2007) analyze the collisional evolution of L4 Jovian Trojans. They obtained that most of the bodies ejected from the L4 swarm are small; one could expect up to ∼ 50 Trojans ejected from L4 swarm with D > 1 km per Myr. So, the escape rate by collisions would be roughly three times the dynamical rate of escape at least for small bodies, since large Trojans are unaffected by collisional evolution (de El´ıa and Brunini, 2007). 25 Then, the real number of escapes of small Trojans would be the sum of the collisional and dynamical removal. We calculated the number of escaped Trojans in each minor body popu- lation. Considering the real asymmetry and the real number of Trojans, the contribution to JFCs and no-JFCs from L4 and L5 is almost the same, but the L4 contribution to Centaurs and TNOs is orders of magnitude greater than that of L5. Considering the collisional removal, and assuming that Tro- jans that escape by collisions follow the same dynamical behavior that the ones removed by dynamics, we would have a minor contribution of Trojans to comets and Centaurs. For example, from our dynamical simulation, we could expect ∼ 20 Trojans in cometary orbits with D > 1 km or ∼ 7 in JFC-orbit. There are some specific regions where escaped Trojans could be important if considering dynamical plus collisional removal. • We could expect 8 escaped Trojans with D > 1 km in an NEO-JFC orbit with dynamical lifetimes of ∼ 30000 yrs. Fern´andez and Sosa (2015) obtained that at least 8 JFCs in NEO orbits show a stable behavior and their composition is mostly rocky; their source region is in the outer main asteroid belt. • A fraction of escaped Trojans go through the zone of Encke type comets. We could expect 4 ETC with D > 1 km in this zone. However the exact orbital elements of 2P/Encke are not reached by escaped Trojans. • We could expect that 1 escaped Trojan with D > 1 km every 350000 yrs would cross the Roche limit of Jupiter, then fragment and impact Jupiter as the SL9 case. This estimation is already smaller than the contribution of Hildas to this type of objects, but it is not negligible. Although the contribution of escaped Trojans to other minor body popu- lations would be minor, at least it could explain some peculiarities observed in some populations of our Solar System. Acknowledgments: The authors wish to express their gratitude to the FCAGLP for extensive use of their computing facilities and acknowledge the financial support by IALP, CONICET and Agencia de Promoci´on Cient´ıfica, through the grants PIP 0436 and PICT 2014-1292. We would like to thank Gabinete de Ingl´es de la Facultad de Ciencias Astron´omicas y Geof´ısicas de la 26 UNLP for a careful language revision. We also acknowledge the constructive comments from Christos Efthymiopoulos and an anonymous referee which helped us improve the article. References [1] Beaug´e, C., Roig, F.V. 2001. A Semianalytical Model for the Motion of the Trojan Asteroids: Proper Elements and Families. Icarus, 153, 391. [2] de El´ıa, G.C., Brunini, A. 2007. Collisional and dynamical evolution of the L4 Trojan asteroids. A&A, 475, 375. [3] Di Sisto, R.P., Brunini, A., Dirani, L.D., Orellana, R. B. 2005. Hilda asteroids among Jupiter family comets. Icarus, 174, 81-89. [4] Di Sisto, R.P., Fern´andez, J.A., Brunini, A. 2009. On the population, physical decay and orbital distribution of Jupiter Family Comets. Nu- merical Simulations. Icarus, 203, 140-154. [5] Di Sisto, R.P., Ramos, X.S., Beaug´e, C. 2014. Giga-year evolution of Jupiter Trojans and the asymmetry problem. Icarus 243, 287-295. [6] Emery, J.P., Burr, D. M., Cruikshank, D. P. 2011. Near-Infrared spec- troscopy of Trojan asteroids: evidence for two compositional groups. The Astronomical Journal, 141, 25. [7] ´Erdi, B. 1996. The Trojan problem. CeMDA, 65, 149. [8] Fern´andez, J.A., Gallardo, T., Brunini, A. 2002. Are There Many Inac- tive Jupiter-Family Comets among the Near-Earth Asteroid Population? Icarus, 159, 358. [9] Fern´andez, Y.R., Sheppard, S.S. & Jewitt, D.C. 2003. The Albedo Dis- tribution of Jovian Trojan Asteroids. AJ, 126, 1563. [10] Fern´andez, Y.R., Jewitt, D.C. & Ziffer, J.E. 2009. Albedos of Small Jovian Trojans. AJ, 138, 240. [11] Fern´andez, J.A., A. Sosa. 2015. Jupiter family comets in near-Earth orbits: Are some of them interlopers from the asteroid belt?. Planetary and Space Science, 118, 14-24. 27 [12] Fern´andez, J.A., M. Helal, T. Gallardo. 2018. Dynamical evolution and end states of active and inactive Centaurs. Planetary and Space Science, 158, 6-15. [13] Fornasier, S., Dotto, E., Hainaut, O., et al. 2007. Visible spectroscopic and photometric survey of Jupiter Trojans: Final results on dynamical families. Icarus, 190, Issue 2, 622-642. [14] Grav, T., Mainzer, A.K., Bauer, J., et al. 2011. WISE/NEOWISE ob- servations of the Jovian Trojans: preliminary results. ApJ, 742, 40. [15] Grav, T., Mainzer, A.K., Bauer, J., et al. 2012. WISE/NEOWISE ob- servations of the Jovian Trojan population: taxonomy. ApJ, 759, 49. [16] Jewitt, D.C., Trujillo, C.A. & Luu, J.X. 2000. Population and Size Dis- tribution of Small Jovian Trojan Asteroids. AJ, 120, 1140. [17] Levison, H.F., Shoemaker, E.M., Shoemaker, C.S. 1997. Dynamical evo- lution of Jupiter's Trojan asteroids. Nature, 385, 42. [18] Mainzer, A., Bauer, J., Grav, J. et al. 2011. Preliminary results from NEOWISE: an enhancement to the wide-field infrared survey explorer for solar system science. The Astrophysical Journal, 731, 53. [19] Marzari, F., Scholl, H. 1998. The growth of Jupiter and Saturn and the capture of Trojans. A&A, 339, 278-285. [20] Marzari, F., Scholl, H. 2000. The Role of Secular Resonances in the History of Trojans. Icarus 146, 232-239. [21] Marzari, F., Scholl, H. 2002. On the Instability of Jupiters Trojans. Icarus 159, 328-338. [22] Marzari, F., Tricarico, P. and Scholl, H. 2003. Stability of Jupiter Tro- jans investigated using frequency map analysis: the MATROS project. Mon. Not. R. Astron. Soc., 345, 1091-1100. [23] Mikkola, S. and Innanen, K. 1992. A numerical exploration of the evo- lution of Trojan-type asteroidal orbits. the Astronomical Journal, 104, N4, 1641-1649. 28 [24] Milani, A. 1993. The Trojan asteroid belt: Proper elements, stability, chaos and families. CeMDA, 57, 59. [25] Morbidelli, A., Levison, H.F., Tsiganis, K., Gomes, R.S. 2005. Origin of the orbital architecture of the giant planets of the Solar System. Nature, 435, 462. [26] Nesvorn´y, D., Vokrouhlick´y, D., Morbidelli, A. 2013. Capture of Trojans by jumping Jupiter. ApJ, 768, 45. [27] Nesvorn´y, D., Vokrouhlick´y, D., Dones, L., Levison, H., Kaib., N., Mor- bidelli, A. 2017. Origin and Evolution of Short-period Comets. ApJ, 845, 27. [28] Opik, E. J. 1976. Interplanetary Encounters, Elsevier, New York. [29] Robutel, P., Gabern, F., Jorba, A. 2005. The Observed Trojans and the Global Dynamics Around The Lagrangian Points of the Sun Jupiter System. CeMDA, 92, 55. [30] Robutel, P., Gabern, F. 2006. The resonant structure of Jupiter's Trojan asteroids - I. Long-term stability and diffusion. MNRAS, 372, 1463. [31] Szab´o, G.M., Ivezi´c, Z., Juri´c, M., Lupton, R. 2007. MNRAS, 337, 1393. [32] Tancredi, G. . 2014, Icarus, 234, 66. [33] Tsiganis, K., Varvoglis, H., Dvorak, R. 2005. Chaotic Diffusion And Effective Stability of Jupiter Trojans. CeMDA, 92, 71. [34] Valsecchi, G.B., Milani, A., Gronchi, G.F., Chesley, S.R. 2000. The dis- tribution of energy perturbations at planetary close encounters. Celestial Mechanics and Dynamical Astronomy 78, 83-91, 2000. [35] Wong, I and Brown, M.E. 2015. The color-magnitude distribution of small Jupiter Trojans. The Astronomical Journal, 150, 174. [36] Yoshida, F. & Nakamura, T., 2005. Size Distribution of faint L4 Trojan asteroids. The Astronomical Journal, 130, 2900-2911. [37] Yoshida, F. & Nakamura, T., 2008. A Comparative Study of Size Distri- bution for Small L4 and L5 Jovian Trojans. Publ. Astron. Soc. Japan, 60, 297. 29
1812.06993
2
1812
2019-02-14T14:03:12
A search for accreting young companions embedded in circumstellar disks: High-contrast H$\alpha$ imaging with VLT/SPHERE
[ "astro-ph.EP" ]
Aims: We want to detect and quantify observables related to accretion processes occurring locally in circumstellar disks, which could be attributed to young forming planets. We focus on objects known to host protoplanet candidates and/or disk structures thought to be the result of interactions with planets. Methods: We analyzed observations of 6 young stars (age $3.5-10$ Myr) and their surrounding environments with the SPHERE/ZIMPOL instrument on the VLT in the H$\alpha$ filter (656 nm) and a nearby continuum filter (644.9 nm). Results: We re-detect the known accreting M-star companion HD142527 B with the highest published signal to noise to date in both H$\alpha$ and the continuum. We derive new astrometry ($r = 62.8^{+2.1}_{-2.7}$ mas and $\text{PA} = (98.7\,\pm1.8)^\circ$) and photometry ($\Delta$N_Ha=$6.3^{+0.2}_{-0.3}$ mag, $\Delta$B_Ha=$6.7\pm0.2$ mag and $\Delta$Cnt_Ha=$7.3^{+0.3}_{-0.2}$ mag) for the companion in agreement with previous studies, and estimate its mass accretion rate ($\dot{M}\approx1-2\,\times10^{-10}\,M_\odot\text{ yr}^{-1}$). A faint point-like source around HD135344 B (SAO206462) is also investigated, but a second deeper observation is required to reveal its nature. No other companions are detected. In the framework of our assumptions we estimate detection limits at the locations of companion candidates around HD100546, HD169142 and MWC758 and calculate that processes involving H$\alpha$ fluxes larger than $\sim8\times10^{-14}-10^{-15}\,\text{erg/s/cm}^2$ ($\dot{M}>10^{-10}-10^{-12}\,M_\odot\text{ yr}^{-1}$) can be excluded. Furthermore, flux upper limits of $\sim10^{-14}-10^{-15}\,\text{erg/s/cm}^2$ ($\dot{M}<10^{-11}-10^{-12}\,M_\odot \text{ yr}^{-1}$) are estimated within the gaps identified in the disks surrounding HD135344B and TW Hya.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. Sphere_Halpha February 15, 2019 ©ESO 2019 A search for accreting young companions embedded in circumstellar disks: High-contrast Hα imaging with VLT/SPHERE(cid:63),(cid:63)(cid:63) G. Cugno1, S. P. Quanz1, 2, S. Hunziker1, T. Stolker1, H. M. Schmid1, H. Avenhaus3, P. Baudoz4, A. J. Bohn5, M. Bonnefoy6, E. Buenzli1, G. Chauvin6, 7, A. Cheetham8, S. Desidera9, C. Dominik10, P. Feautrier6, M. Feldt3, C. Ginski5, J. H. Girard11, 6, R. Gratton9, J. Hagelberg1, E. Hugot12, M. Janson13, A.-M. Lagrange6, M. Langlois12, 14, Y. Magnard6, A.-L. Maire3, F. Menard6, 15, M. Meyer16, 1, J. Milli17, C. Mordasini18, C. Pinte19, 6, J. Pragt20, R. Roelfsema20, F. Rigal20, J. Szulágyi21, R. van Boekel3, G. van der Plas6, A. Vigan12, Z. Wahhaj17, and A. Zurlo12, 22, 23 (Affiliations can be found after the references) Received -- ; accepted -- ABSTRACT Context. In recent years, our understanding of giant planet formation progressed substantially. There have even been detections of a few young protoplanet candidates still embedded in the circumstellar disks of their host stars. The exact physics that describes the accretion of material from the circumstellar disk onto the suspected circumplanetary disk and eventually onto the young, forming planet is still an open question. Aims. We seek to detect and quantify observables related to accretion processes occurring locally in circumstellar disks, which could be attributed to young forming planets. We focus on objects known to host protoplanet candidates and/or disk structures thought to be the result of interactions with planets. Methods. We analyzed observations of six young stars (age 3.5 − 10 Myr) and their surrounding environments with the SPHERE/ZIMPOL instrument on the Very Large Telescope (VLT) in the Hα filter (656 nm) and a nearby continuum filter (644.9 nm). We applied several point spread function (PSF) subtraction techniques to reach the highest possible contrast near the primary star, specifically investigating regions where forming companions were claimed or have been suggested based on observed disk morphology. Results. We redetect the known accreting M-star companion HD142527 B with the highest published signal to noise to date in both Hα and the continuum. We derive new astrometry (r = 62.8+2.1−2.7 mas and PA = (98.7 ± 1.8)◦) and photometry (∆N_Ha=6.3+0.2−0.3 mag, ∆B_Ha=6.7 ± 0.2 mag and ∆Cnt_Ha=7.3+0.3−0.2 mag) for the companion in agreement with previous studies, and estimate its mass accretion rate ( M ≈ 1 − 2 × 10−10 M(cid:12) yr−1). A faint point-like source around HD135344 B (SAO206462) is also investigated, but a second deeper observation is required to reveal its nature. No other companions are detected. In the framework of our assumptions we estimate detection limits at the locations of companion candidates around HD100546, HD169142, and MWC 758 and calculate that processes involving Hα fluxes larger than ∼ 8 × 10−14 − 10−15 erg/s/cm2 ( M > 10−10 − 10−12 M(cid:12) yr−1) can be excluded. Furthermore, flux upper limits of ∼ 10−14 − 10−15 erg/s/cm2 ( M < 10−11 − 10−12 M(cid:12) yr−1) are estimated within the gaps identified in the disks surrounding HD135344 B and TW Hya. The derived luminosity limits exclude Hα signatures at levels similar to those previously detected for the accreting planet candidate LkCa15 b. Key words. Planetary systems, Planet-disk interactions -- Techniques: high angular resolution -- Planets and satellites: detection, formation 1. Introduction Providing an empirical basis for gas giant planet formation mod- els and theories requires the detection of young objects in their natal environment, i.e., when they are still embedded in the gas and dust-rich circumstellar disk surrounding their host star. The primary scientific goals of studying planet formation are as follows: To understand where gas giant planet formation takes (cid:63) Based on observations collected at the Paranal Observatory, ESO (Chile). Program ID: 096.C-0248(B), 096.C-0267(A),096.C-0267(B), 095.C-0273(A), 095.C-0298(A) (cid:63)(cid:63) Figures 1, 6 and D.1 are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr(130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/ place, for example, at what separations from the host star and under which physical and chemical conditions in the disk; how formation occurs, i.e., via the classical core accretion process (Pollack et al. 1996) or a modified version of that process (e.g., pebble accretion, Lambrechts & Johansen 2012) or direct grav- itational collapse (Boss 1997)); and the properties of the sus- pected circumplanetary disks (CPDs). While in recent years high-contrast, high spatial resolution imag- ing observations of circumstellar disks have revealed an impres- sive diversity in circumstellar disk structure and morphology, the number of directly detected planet candidates embedded in those disks is still small (LkCa15 b, HD100546 b, HD169142 b, MWC 758 b, PDS 70 b; Kraus & Ireland 2012; Quanz et al. 2013a; Reggiani et al. 2014; Biller et al. 2014; Reggiani et al. Article number, page 1 of 20 9 1 0 2 b e F 4 1 . ] P E h p - o r t s a [ 2 v 3 9 9 6 0 . 2 1 8 1 : v i X r a A&A proofs: manuscript no. Sphere_Halpha 2018; Keppler et al. 2018). To identify these objects, high- contrast exoplanet imaging can be used. These observations are typically performed at near- to mid-infrared wavelengths using an adaptive optics-assisted high-resolution camera. In addition to the intrinsic luminosity of the still contracting young gas gi- ant planet, the surrounding CPD, if treated as a classical accre- tion disk, contributes significantly to fluxes beyond 3 µm wave- length (Zhu 2015; Eisner 2015), potentially easing the detection of young forming gas giants at these wavelengths. While the ma- jority of the forming planet candidates mentioned above were detected in this way, it has also been realized that the signature from a circumstellar disk itself can sometimes mimic that of a point source after PSF subtraction and image post-processing (e.g., Follette et al. 2017; Ligi et al. 2018). As a consequence, it is possible that some of the aforementioned candidates are false positives. Another approach is to look for direct signatures of the sus- pected CPDs, such as their dust continuum emission or their kinematic imprint in high-resolution molecular line data (Perez et al. 2015; Szulágyi et al. 2018). In one case, spectro-astrometry using CO line emission was used to constrain the existence and orbit of a young planet candidate (Brittain et al. 2013, 2014). Moreover, Pinte et al. (2018) and Teague et al. (2018) suggested the presence of embedded planets orbiting HD163296 from local deviations from Keplerian rotation in the protoplanetary disk. A further indirect way to infer the existence of a young, forming planet is to search for localized differences in the gas chemistry of the circumstellar disk, as the planet provides extra energy to the chemical network in its vicinity (Cleeves et al. 2015). Finally, it is possible to look for accretion signatures from gas falling onto the planet and its CPD. Accretion shocks are able to excite or ionize the hydrogen atoms, which then radi- ate recombination emission lines, such as Hα, when returning to lower energy states (e.g., Calvet & Gullbring 1998; Szulágyi & Mordasini 2017; Marleau et al. 2017). High-contrast imag- ing using Hα filters was already successfully applied in three cases. Using angular spectral differential imaging (ASDI) with the Magellan Adaptive Optics System (MagAO), Close et al. (2014a) detected Hα excess emission from the M-star compan- ion orbiting the Herbig Ae/Be star HD142527, and Sallum et al. (2015) also used MagAO to identify at least one accreting com- panion candidate located in the gap of the transition disk around LkCa15. The accretion signature was found at a position very similar to the predicted orbital position of one of the faint point sources detected by Kraus & Ireland (2012), attributed to a form- ing planetary system. Most recently, Wagner et al. (2018) have claimed the detection of Hα emission from the young planet PDS70 b using MagAO, albeit with comparatively low statisti- cal significance (3.9σ). In this paper we present a set of Hα high-contrast imaging data for six young stars, aiming at the detection of potential ac- cretion signatures from the (suspected) young planets embedded in the circumstellar disks of the stars. The paper is structured as follows: In Section 2 we discuss the observations and target stars. We explain the data reduction in Section 3 and present our analyses in Section 4. In Section 5 we discuss our results in a broader context and conclude in Section 6. 2. Observations and target sample 2.1. Observations The data were all obtained with the ZIMPOL sub-instrument of the adaptive optics (AO) assisted high-contrast imager SPHERE Article number, page 2 of 20 (Beuzit et al. 2008; Petit et al. 2008; Fusco et al. 2016), which is installed at the Very Large Telescope (VLT) of the European Southern Observatory (ESO) on Paranal in Chile. A detailed de- scription of ZIMPOL can be found in Schmid et al. (2018). Some of the data were collected within the context of the Guaran- teed Time Observations (GTO) program of the SPHERE consor- tium; others were obtained in other programs and downloaded from the ESO data archive (program IDs are listed in Table 1). We focused on objects that are known from other observa- tions to host forming planet candidates that still need to be con- firmed (HD100546, HD169142, and MWC 758)1, objects known to host accreting stellar companions (HD142527), and objects that have well-studied circumstellar disks with spatially resolved substructures (gaps, cavities, or spiral arms), possibly suggesting planet formation activities (HD135344 B and TW Hya). All data were taken in the noncoronagraphic imaging mode of ZIMPOL using an Hα filter in one camera arm and a nearby continuum filter simultaneously in the other arm (Cont_Ha; λc = 644.9 nm, ∆λ = 3.83 nm). As the data were observed in different programs, we sometimes used the narrow Hα filter (N_Ha; λc = 656.53 nm, ∆λ = 0.75 nm) and sometimes the broad Hα filter (B_Ha; λc = 655.6 nm, ∆λ = 5.35 nm). A more complete description of these filters can be found in Schmid et al. (2017). To establish which filter allows for the highest contrast performance, we used HD142527 and its accreting companion (Close et al. 2014a) as a test target and switched between the N_Ha and the B_Ha fil- ter every ten frames within the same observing sequence. All datasets were observed in pupil-stabilized mode to enable angu- lar differential imaging (ADI; Marois et al. 2006). The funda- mental properties of the target stars are given in Table 2, while a summary of the datasets is given in Table 1. We note that because of the intrinsic properties of the polariza- tion beam splitter used by ZIMPOL, polarized light might pref- erentially end up in one of the two arms, causing a systematic uncertainty in the relative photometry between the continuum and Hα frames. The inclined mirrors in the telescope and the instrument introduce di-attenuation (e.g., higher reflectivity for I⊥ than I(cid:107)) and polarization cross talks, so that the transmissions in imaging mode to the I⊥ and I(cid:107) arm depend on the telescope pointing direction. This effect is at the level of a few percent (about ±5 %), but unfortunately the dependence on the instru- ment configuration has not been determined yet. We discuss its potential impact on our analyses in Appendix A, even though we did not take this effect into account since it is small and could not be precisely quantified. 2.2. Target sample HD142527 HD142527 is known to have a prominent circumstellar disk (e.g., Fukagawa et al. 2006; Canovas et al. 2013; Avenhaus et al. 2014b) and a close-in M star companion (HD142527 B; Biller et al. 2012; Rodigas et al. 2014; Lacour et al. 2016; Christiaens et al. 2018; ?) that shows signatures of ongoing accretion in Hα emission (Close et al. 2014a). This companion orbits in a large, optically thin cavity within the circumstellar disk stretching from ∼ 0(cid:48)(cid:48).07 to ∼ 1(cid:48)(cid:48).0 (e.g., Fukagawa et al. 2013; Avenhaus et al. 2014b), and it is likely that this companion is at least partially responsible for clearing the gap by accretion 1 In the discussion (Section 5) we also include the analysis of a dataset of LkCa15 (PI: Huelamo) to set our results in context, but the data were poor in quality and hence not included in the main part of the paper. G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks Table 1. Summary of observations. Object HD142527 Hα Filtera B_Ha N_Ha HD135344 B N_Ha B_Ha TW Hya B_Ha HD100546 HD169142 B_Ha B_Ha MWC 758 Obs. date [dd.mm.yyyy] 31.03.2016 31.03.2016 31.03.2016 23.03.2016 23.04.2015 09.05.2015 30.12.2015 Prog. ID 096.C-0248(B) 096.C-0248(B) 096.C-0248(B) 096.C-0267(B) 095.C-0273(A) 095.C-0298(A) 096.C-0267(A) DITb [s] 30 30 50 80 10 50 60 # of DITs 70 70 107 131 1104e 90 194 Field rotation [◦] 47.8 48.6 71.7 134.1 68.3e 123.2 54.8 Mean airmass 1.06 1.05 1.04 1.16 1.46 1.01 1.63 c τ0 [ms] 2.7 ± 0.2 2.7 ± 0.3 4.4 ± 1.2 1.4 ± 0.4 1.7 ± 0.2 1.4 ± 0.1 3.2 ± 0.8 Mean seeingd [as] 0.71 ± 0.06 0.69 ± 0.07 0.47 ± 0.17 1.33 ± 0.53 0.98 ± 0.28 1.24 ± 0.04 1.39 ± 0.24 Notes. (a) Each dataset consists of data obtained in one of the two Hα filters and simultaneous data taken with the continuum filter inserted in the other ZIMPOL camera. (b) DIT = Detector integration time, i.e., exposure time per image frame. (c) Coherence time .(d) Mean DIMM seeing measured during the observation. (e) As we explain in Section 4.4 and Appendix E, for this dataset a frame selection was applied, which reduced the number of frames to 366 and the field rotation to 20.7◦. of disk material (Biller et al. 2012; Price et al. 2018). Avenhaus et al. (2017) obtained polarimetric differential imaging data with SPHERE/ZIMPOL in the very broad band (VBB, as defined in Schmid et al. 2018) optical filter, revealing new substructures, and resolving the innermost regions of the disk (down to 0(cid:48)(cid:48).025). In addition, extended polarized emission was detected at the position of HD142527 B, possibly due to dust in a circumsecondary disk. Christiaens et al. (2018) extracted a medium-resolution spectrum of the companion and suggested a mass of 0.34 ± 0.06 M(cid:12). This value is a factor of ∼ 3 larger than that estimated by spectral energy distribution (SED) fitting (Lacour et al. 2016, M = 0.13 ± 0.03 M(cid:12)). Thanks to the accreting close-in companion, this system is the ideal target to optimize the Hα observing strategy with SPHERE/ZIMPOL and also the data reduction. HD135344 B HD135344 B (SAO206462) is surrounded by a transition disk that was spatially resolved at various wavelengths. Con- tinuum (sub-)millimeter images presented by Andrews et al. (2011) and van der Marel et al. (2016) revealed a disk cavity with an outer radius of 0(cid:48)(cid:48).32. In polarimetric differential imaging (PDI) observations in the near-infrared (NIR), the outer radius of the cavity appears to be at 0(cid:48)(cid:48).18, and the difference in apparent size was interpreted as a potential indication for a companion orbiting in the cavity (Garufi et al. 2013). Data obtained in PDI mode also revealed two prominent, symmetric spiral arms (Muto et al. 2012; Garufi et al. 2013; Stolker et al. 2016). Vicente et al. (2011) and Maire et al. (2017) searched for planets in the system using NIR NACO and SPHERE high-contrast imaging data, but did not find any. Using hot start evolutionary models these authors derived upper limits for the mass of potential giant planets around HD135344 B (3 MJ beyond 0(cid:48)(cid:48).7). TW Hya TW Hya is the nearest T Tauri star to Earth. Its almost face-on transitional disk (i ∼ 7 ± 1◦; Qi et al. 2004) shows multiple rings and gaps in both dust continuum and scattered light data. Hubble Space Telescope (HST) scattered light images from Debes et al. (2013) first allowed the identification of a gap at ∼ 1(cid:48)(cid:48).48. Later, Akiyama et al. (2015) observed in H-band polarized images a gap at a separation of ∼ 0(cid:48)(cid:48).41. Using Atacama Large Millimeter Array (ALMA), Andrews et al. (2016) identified gaps from the radial profile of the 870 µm continuum emission at 0(cid:48)(cid:48).41, 0(cid:48)(cid:48).68 and 0(cid:48)(cid:48).80. Finally, van Boekel et al. (2017) obtained SPHERE images in PDI and ADI modes at optical and NIR wavelengths, and identified three gaps at 0(cid:48)(cid:48).11, 0(cid:48)(cid:48).39, and 1(cid:48)(cid:48).57 from the central star. A clear gap was also identified by Rapson et al. (2015) at a separation of 0(cid:48)(cid:48).43 in Gemini/GPI polarimetric images and the largest gap at r (cid:39) 1(cid:48)(cid:48).52 has also been observed in CO emission with ALMA (Huang et al. 2018). HD100546 The disk around HD100546 was also spatially resolved in scattered light and dust continuum emission in different bands (e.g., Augereau et al. 2001; Quanz et al. 2011; Avenhaus et al. 2014a; Walsh et al. 2014; Pineda et al. 2014). The disk appears to be almost, but not completely, devoid of dusty material at radii between a few and 13 AU. This gap could be due to the interaction with a young forming planet, and Brittain et al. (2013, 2014) suggested the presence of a companion orbiting the star at 0(cid:48)(cid:48).13, based on high-resolution NIR spectro-astrometry of CO emission lines. Another protoplanet candidate was claimed by Quanz et al. (2013a) using L(cid:48) band high-contrast imaging data. The object was found at 0(cid:48)(cid:48).48 ± 0(cid:48)(cid:48).04 from the central star, at a position angle (PA) of (8.9 ± 0.9)◦, with an apparent magnitude of L(cid:48)=13.2 ± 0.4 mag. Quanz et al. (2015) reobserved HD100546 in different bands (L(cid:48), M(cid:48), Ks) and detected the object again in the first two filters. Based on the colors and observed morphology these authors suggested that the data are best explained by a forming planet surrounded by a circumplanetary disk. Later, Currie et al. (2015) recovered HD100546 b from H-band integral field spectroscopy (IFS) with the Gemini Planet Imager (GPI; Macintosh et al. 2006) and identified a second putative point source c closer to the star (rproj ∼ 0(cid:48)(cid:48).14) potentially related to the candidate identified by Brittain et al. (2013, 2014). More recently, Rameau et al. (2017) demonstrated that the emission related to HD100546 b appears to be stationary and its spectrum is inconsistent with any type of low temperature objects. Furthermore, they obtained Hα images with the MagAO instrument to search for accretion signatures, but no point source was detected at either the b or c position, and they placed upper limits on the accretion lumi- nosity (Lacc < 1.7 × 10−4 L(cid:12)). The same data were analyzed by Article number, page 3 of 20 Table 2. Target sample. A&A proofs: manuscript no. Sphere_Halpha RA Object 15h56m41.89s HD142527 HD135344 B 15h15m48.44s 11h01m51.90s TW Hya 11h33m25.44s HD100546 18h24m29.78s HD169142 05h30m27.53s MWC 758 Spec. type mR [mag] DEC -42◦19(cid:48)23(cid:48)(cid:48).27 F6III -37◦09(cid:48)16(cid:48)(cid:48).03 F8V -34◦42(cid:48)17(cid:48)(cid:48).03 K6Ve -70◦11(cid:48)41(cid:48)(cid:48).24 B9Vne -29◦46(cid:48)49(cid:48)(cid:48).32 B9V -25◦19(cid:48)57(cid:48)(cid:48).08 A8Ve 7.91 8.45 10.43 ± 0.1 8.78 8.0 9.20 ± 0.01 Distance [pc] Age [Myr] 157.3 ± 1.2 135.9 ± 1.4 60.1 ± 0.1 110.0 ± 0.6 114.0 ± 0.8 160.3 ± 1.7 8.1+1.9−1.6 9 ± 2 ∼ 10 7 ± 1.5 ∼ 6 3.5 ± 2 Notes. Coordinates and spectral types are taken from SIMBAD, R-magnitudes are taken from the NOMAD catalog (Zacharias et al. 2004) for HD142527 and HD169142, from the APASS catalog (Henden et al. 2016) for HD135344 B, and from the UCAC4 catalog (Zacharias et al. 2012) for the other targets. Distances are from GAIA data release 2 (Gaia Collaboration et al. 2018). The ages -- from top to bottom -- are taken from Fairlamb et al. (2015), Müller et al. (2011), Weinberger et al. (2013), Fairlamb et al. (2015), Grady et al. (2007), and Meeus et al. (2012). Follette et al. (2017), together with other Hα images (MagAO), H band spectra (GPI), and Y band polarimetric images (GPI). Their data exclude that HD100546 c is emitting in Hα with LHα > 1.57 × 10−4L(cid:12). HD169142 HD169142 is surrounded by a nearly face-on pre-transitional disk. Using PDI images, Quanz et al. (2013b) found an unre- solved disk rim at 0(cid:48)(cid:48).17 and an annular gap between 0(cid:48)(cid:48).28 and 0(cid:48)(cid:48).49. These results were confirmed by Osorio et al. (2014), who investigated the thermal emission (λ = 7 mm) of large dust grains in the HD169142 disk, identifying two annular cavities (∼ 0(cid:48)(cid:48).16 − 0(cid:48)(cid:48).21 and ∼ 0(cid:48)(cid:48).28 − 0(cid:48)(cid:48).48). The latter authors also identified a point source candidate in the middle of the outer cavity at a distance of 0(cid:48)(cid:48).34 and PA ∼ 175◦. Biller et al. (2014) and Reggiani et al. (2014) observed a point-like feature in NaCo L(cid:48) data at the outer edge of the inner cavity (separation = 0(cid:48)(cid:48).11 − 0(cid:48)(cid:48).16 and PA=0◦ − 7.4◦). Observations in other bands (H, KS , zp) with the Magellan Clay Telescope (MagAO/MCT) and with GPI in the J band failed to confirm the detection (Biller et al. 2014; Reggiani et al. 2014), but revealed another candidate point source albeit with low signal-to-noise ratio (S/N; Biller et al. 2014). In a recent paper, Ligi et al. (2018) explained the latter Biller et al. (2014) detection with a bright spot in the ring of scattered light from the disk rim, potentially following Keplerian motion. Pohl et al. (2017) and Bertrang et al. (2018) compared different disk and dust evolutionary models to SPHERE J-band and VBB PDI observations. Both works tried to reproduce and explain the complex morphological structures observed in the disk and conclude that planet-disk interaction is occurring in the system, even though there is no clearly confirmed protoplanet identified to date. MWC 758 MWC 758 is surrounded by a pre-transitional disk (e.g., Grady et al. 2013). Andrews et al. (2011) found an inner cavity of ∼55 AU based on dust continuum observations, which was, however, not observed in scattered light (Grady et al. 2013; Benisty et al. 2015). Nevertheless, PDI and direct imaging from the latter studies revealed two large spiral arms. A third spiral arm has been suggested based on VLT/NaCo L(cid:48) data by Reggiani et al. (2018), together with the claim of the detection of a point-like source embedded in the disk at (111 ± 4) mas. This object was observed in two separate datasets from 2015 Article number, page 4 of 20 and 2016 at comparable separations from the star, but different PAs, which was possibly due to orbital motion. The contrast of this object relative to the central star in the L(cid:48) band is ∼ 7 mag, which, according to the BT-Settl atmospheric models (Allard et al. 2012), corresponds to the photospheric emission of a 41-64 MJ object for the age of the star. More recently, ALMA observations from Boehler et al. (2018) traced the large dust continuum emission from the disk. Two rings at 0(cid:48)(cid:48).37 and 0(cid:48)(cid:48).53 were discovered that are probably related to two clumps with large surface density of millimeter dust and a large cavity of ∼ 0(cid:48)(cid:48).26 in radius. Finally, Huélamo et al. (2018) observed MWC 758 in Hα with SPHERE/ZIMPOL, reaching an upper limit (cid:46) 5 × 10−5L(cid:12) (corresponding for the line luminosity of LHα to a contrast of 7.6 mag) at the separation of the protoplanet candidate. No other point-like features were detected. 3. Data reduction The basic data reduction steps were carried out with the ZIM- POL pipeline developed and maintained at ETH Zürich. The pipeline remapped the original 7.2 mas / pixel × 3.6 mas / pixel onto a square grid with an effective pixel scale of 3.6 mas / pixel × 3.6 mas / pixel (1024 × 1024 pixels). Afterward, the bias was subtracted and a flat-field correction was applied. We then aligned the individual images by fitting a Moffat profile to the stellar point spread functions (PSFs) and shifting the images using bilinear interpolation. The pipeline also calculated the par- allactic angle for each individual frame and added the informa- tion to the image header. Finally, we split up the image stacks into individual frames and grouped them together according to their filter, resulting in two image stacks for each object: one for an Hα filter and one for the continuum filter2. In general, all images were included in the analysis if not specifically men- tioned in the individual subsections. The images in these stacks were cropped to a size of 1(cid:48)(cid:48).08 × 1 (cid:48)(cid:48).08 centered on the star. This allowed us to focus our PSF subtraction efforts on the contrast dominated regime of the images. The removal of the stellar PSF was performed in three different ways: ADI, spectral differential imaging (SDI), and ASDI (a two-step combination of SDI and ADI). To perform ADI, we fed the stacks into our PynPoint pipeline (Amara & Quanz 2012; Amara et al. 2015; Stolker et al. 2 For HD142527 we have four image stacks as we used both the N_Ha and the B_Ha filter during the observing sequence. G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks 2018). The PynPoint package uses principal component analy- sis (PCA) to model and subtract the stellar PSF in all individual images before they are derotated to a common field orientation and mean-combined. To investigate the impact on the final con- trast performance for all objects, we varied the number of prin- cipal components (PCs) used to fit the stellar PSF and the size of the inner mask that is used to cover the central core of the stel- lar PSF prior to the PCA. No frame selection based on the field rotation was applied, meaning that all the images were consid- ered for the analysis, regardless of the difference in parallactic angle. The SDI approach aims at reducing the stellar PSF us- ing the fact that all features arising from the parent star (such as Airy pattern and speckles) scale spatially with wavelength λ, while the position of a physical object on the detector is inde- pendent of λ. The underlying assumption is that, given that λc is similar in all filters, the continuum flux density is the same at all wavelengths. To this end, modified versions of the continuum images were created. First, they were multiplied with the ratio of the effective filter widths to normalize the throughput of the con- tinuum filter relative to the Hα filter3. Then, they were spatially stretched using spline interpolation in radial direction, going out from the image center, by the ratio of the central wavelengths of the filters to align the speckle patterns. Because of the possibly different SED shapes of our objects with respect to the standard calibration star used in Schmid et al. (2017) to determine the cen- tral wavelengths λc of the filters, it is possible that λc is slightly shifted for each object. This effect, however, is expected to alter the upscaling factor by at most 0.4% for B_Ha (assuming the unrealistic case in which λc is at the edge of the filter), which is the broadest filter we used. This is negligible at very small separations from the star, where speckles dominate the noise. Values for filter central wavelengths and filter equivalent widths can be found in Table 5 of Schmid et al. (2017). The modified continuum images were then subtracted from the images taken simultaneously with the Hα filter, leaving only Hα line flux emit- ted from the primary star and potential companions. As a final step, the images resulting from the subtraction are derotated to a common field orientation and mean-combined. It is worth not- ing that if, as a result of the stretching, a potential point-source emitting a significant amount of continuum flux moves by more than λ/D, the signal strength in the Hα image is only marginally changed in the SDI subtraction step, and only the speckle noise is reduced. If this is not the case, this subtraction step yields a significant reduction of the source signal in addition to the reduc- tion of the speckle noise. For SPHERE/ZIMPOL Hα imaging, a conservative SDI subtraction without substantial signal removal is achieved for angular separations (cid:38) 0(cid:48)(cid:48).90 (∼ 250 pixels). Nev- ertheless, this technique is expected to enhance the S/N of ac- creting planetary companions even at smaller separations, since young planets are not expected to emit a considerable amount of optical radiation in the continuum. In this case, the absence of a continuum signal guarantees that the image subtraction leaves the Hα signal of the companion unchanged and only reduces the speckle residuals. Therefore, for this science case, there is no penalty for using SDI. To perform ASDI, the SDI (Hα-Cnt_Hα) subtracted images are fed into the PCA pipeline to subtract any remaining residuals. During the analysis we varied the same parameters as described for simple ADI. The HD142527 dataset was used to compare the different sensitivities achieved when applying ADI, SDI, and 3 This approach ignores any potential color effects between the filters, which, given their narrow band widths, should, however, not cause any significant systematic offsets. ASDI. The results are discussed in Section 4.1.1 and Appendix B. With ZIMPOL in imaging mode, there is a constant offset of (135.99 ± 0.11)◦ between the parallactic angle and the PA of the camera in sky coordinates (Maire et al. 2016). A preliminary as- trometric calibration showed, however, that this reference frame has to be rotated by (−2.0 ± 0.5)◦ to align images with north pointing to the top (Ginski et al., in preparation). This means that overall, for every PSF subtraction technique, the final im- ages have to be rotated by (134 ± 0.5)◦ in the counterclockwise direction. 4. Analysis and results 4.1. HD142527 B: The accreting M-star companion 4.1.1. Comparing the performance of multiple observational setups In this section, we quantitatively compare the detection perfor- mance for multiple filter combinations and PSF subtraction tech- niques and establish the best strategy for future high-contrast Hα observations with SPHERE/ZIMPOL. For the analysis, the HD142527 dataset was used; during the data reduction, no fur- ther frame selection was applied. The final images of HD142527 clearly show the presence of the M-star companion east of the central star. The signal is detected in all filters with ADI (B_Ha, N_Ha, and Cnt_Ha) and ASDI (in both continuum-subtracted B_Ha and N_Ha images) over a broad range of PCs and also for different image and inner mask sizes (see Figure 1). We used the prescription from Mawet et al. (2014) to compute the false positive fraction (FPF) as a metric to quantify the con- fidence in the detection. The flux is measured in apertures of di- ameter λ/D (16.5 mas) at the position of the signal and in equally spaced reference apertures placed at the same separation but with different PAs, so that there is no overlap between these angles and the remaining azimuthal space is filled. These apertures sam- ple the noise at the separation of the companion. Since the aper- tures closest to the signal are dominated by negative wings from the PSF subtraction process, they were ignored. Then, we used Equation 9 and Equation 10 from Mawet et al. (2014) to cal- culate S/N and FPF from these apertures. This calculation takes into account the small number of apertures that sample the noise and uses the Student t-distribution to calculate the confidence of a detection. The wider wings of the t-distribution enable a better match to a non-Gaussian residual speckle noise than the normal distribution. However, the true FPF values could be higher if the wings of the true noise distribution are higher than those of the t-distribution4. The narrow N_Ha filter delivers a significantly lower FPF than the broader B_Ha filter over a wide range of PCs (see Fig- ure B.1 in Appendix B). Figure B.1 also shows that the com- bination of SDI and ADI yields lower FPF values than only ADI for both filters. Applying ASDI on N_Ha images is hence the preferred choice for future high-contrast imaging programs with SPHERE/ZIMPOL in the speckle-limited regime close to the star. Furthermore, as shown in Figure C.1 and explained in Appendix C, it is crucial to plan observations maximizing the 4 As an example, Figure 7 of Mawet et al. (2014) shows how the t- distribution produces lower FPF values than the case where speckle noise follows more closely a modified Rician distribution. Nevertheless, it has been shown that applying ADI removes the correlated component of the noise leaving quasi-Gaussian residuals (Marois et al. 2008). Article number, page 5 of 20 A&A proofs: manuscript no. Sphere_Halpha Fig. 1. Final ADI and ASDI reduced images of HD142527. Top row: B_Ha, Cnt_Ha, and N_Ha filter images resulting in the lowest FPFs (1.5 × 10−11, 2.2 × 10−9, and < 10−17, corresponding to S/Ns of 13.1, 9.8, and 26.6, respectively). Bottom row: final images after ASDI reduction for B_Ha-Cnt_Ha and N_Ha-Cnt_Ha frames (4.4× 10−16 and < 10−17, corresponding to S/Ns of 22.7 and 27.6). We give the number of subtracted PCs and the radius of the central mask in milliarcseconds in the top left corner of each image. The color scales are different for the two rows. Because all images of the top row have the same color stretch, the detection appears weaker in the continuum band. In Figure 2 we show the resulting contrast curves for the three filters for a confidence level (CL) of 99.99995%. For each dataset (B_Ha, N_Ha, and Cnt_Ha) and technique (ADI and ASDI), we calculated the contrast curves for different numbers of PCs (between 10 and 30 in steps of 5) after removing the companion (see Section 4.1.2). From each set of curves, we only considered the best achievable contrast at each separation from the central star. The presence of Hα line emission from the cen- tral star made SDI an inefficient technique to search for faint objects at small angular separations. To derive the contrast curves, artificial companions with varying contrast were inserted at six different PAs (separated by 60◦) and in steps of 0(cid:48)(cid:48).03 in the radial direction. As the stellar PSF was unsaturated in all individual frames, the artificial com- panions were obtained by shifting and flux-scaling the stellar PSFs and then adding these companions to the original frames. Also for the calculation of the ASDI contrast curves, the origi- nal Hα filter images, containing underlying continuum and Hα line emission, were used to create artificial secondary signals. For each reduction run only one artificial companion was in- serted at a time to keep the PCs as similar as possible to the original reduction. The brightness of the artificial signals was re- duced/increased until their FPF corresponded to a detection with a CL of 99.99995% (i.e., a FPF of 2.5×10−7), corresponding to ≈5σ whether Gaussian noise was assumed. An inner mask with Fig. 2. Contrast curves for HD142527. The colored shaded regions around each curve represent the standard deviation of the achieved con- trast at the 6 azimuthal positions considered at each separation. The markers (red diamond, orange circle, and violet star) represent the con- trast of HD142527 B. field rotation to best modulate and subtract the stellar PSF and to achieve higher sensitivities. Article number, page 6 of 20 -0.2-0.100.10.2Arcseconds-0.20.100.10.2ArcsecondsPC = 15r = 10.8 masNEBroad Band Ha (655.6 nm)-0.2-0.100.10.2Arcseconds-0.2-0.100.10.2ArcsecondsPC = 17r = 32.4 masNEContinuum Band (644.9 nm)-0.2-0.100.10.2Arcseconds-0.2-0.100.10.2ArcsecondsPC = 14r = 32.4 masNENarrow Band Ha (656.53 nm)-0.2-0.100.10.2Arcseconds-0.2-0.100.10.2ArcsecondsPC = 27r = 21.6 masNEASDI B_Ha-Cnt_Ha-0.2-0.100.10.2Arcseconds-0.2-0.100.10.2ArcsecondsPC = 17r = 32.4 masNEASDI N_Ha-Cnt_Ha0.00.20.40.60.81.00.00.20.40.60.81.0-0.200.20.40.60.81Flux (arbitrary linear scale)-0.300.30.61flux (arbitrary linear scale) G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks a radius of 0(cid:48)(cid:48).02 was used to exclude the central parts dominated by the stellar signal. The colored shaded regions around each curve represent the standard deviation of the contrast achieved at that specific separation within the six PAs. It is important to note that, while in Figure B.1 the N_Ha fil- ter provides the lowest FPF for the companion, Figure 2 seems to suggest that the B_Ha filter provides a better contrast perfor- mance. However, this is an effect from the way the contrast anal- ysis is performed. As described above, the stellar PSF was used as a template for the artificial planets, as it is usually done in high-contrast imaging data analysis. The flux distribution within a given filter can vary significantly depending on the object. In this specific case, HD142527 B is known to have Hα excess emission, hence the flux within either Hα filter is strongly domi- nated by line emission (∼50% in B_Ha and ∼83% in N_Ha filter) and a contribution from the optical continuum can be neglected. The primary shows, however, strong and non-negligible optical continuum emission that contributes to the flux observed in the Hα filters. Indeed, for the primary, only 10% and 56% of the flux in the B_Ha and N_Ha filters are attributable to line emis- sion. Hence, when using the stellar PSF as template for artificial planets, we obtain a better contrast performance for the B_Ha filter as it contains overall more flux. In reality, however, if the goal is to detect Hα line emission from low-mass accreting com- panions, the N_Ha filter is to be preferred. Finally, as found by Sallum et al. (2015) for the planet candidate LkCa15 b, the fact that ASDI curves reach a deeper contrast confirms that this tech- nique, in particular close to the star, is more effective and should be preferred to search for Hα accretion signals. 4.1.2. Quantifying the Hα detection The clear detection of the M-star companion in our images al- lows us to determine its contrast in all the filters and its posi- tion relative to the primary at the epoch of observation. For this purpose, we applied the Hessian matrix approach (Quanz et al. 2015) and calculated the sum of the absolute values of the de- terminants of Hessian matrices in the vicinity of the compan- ion's signal. The Hessian matrix represents the second deriva- tive of an n-dimensional function and its determinant is a mea- sure for the curvature of the surface described by the function. This method allows for a simultaneous determination of the po- sition and the flux contrast of the companion and we applied a Nelder-Mead (Nelder & Mead 1965) simplex algorithm to mini- mize the curvature, i.e., the determinants of the Hessian matrices. We inserted negative, flux-rescaled stellar PSFs at different loca- tions and with varying brightness in the input images and com- puted the resulting curvature within a region of interest (ROI) around the companion after PSF subtraction5. To reduce pixel- to-pixel variations after the PSF-subtraction step and allow for a more robust determination of the curvature, we convolved the images with a Gaussian kernel with a full width at half maxi- mum (FWHM) of 8.3 mas (≈ 0.35 of the FWHM of the stel- lar PSF, which was calculated to be 23.7 mas on average). To fully include the companion's signal, the ROI was chosen to be (43.2 × 43.2) mas around the peak flux detected in the original set of PSF subtracted images. Within the ROI, the determinants of the Hessian matrices in 10,000 evenly spaced positions on a fixed grid (every 0.43 mas) were calculated and summed up. For the optimization algorithm to converge, we need to pro- vide a threshold criterion: if the change in the parameters (posi- 5 For this analysis we used an image size of 0(cid:48)(cid:48).36 × 0(cid:48)(cid:48).36 to speed up the computation and an inner mask of 10.8 mas (radius). tion and contrast) between two consecutive iterations is less than a given tolerance, the algorithm has converged and the optimiza- tion returns those values for contrast and position. The absolute tolerance for the convergence was set to be 0.16, as this value is the precision to which artificial signals can be inserted into the image grid. This value applies for all the investigated parameters (position and contrast). Errors in the separation and PA measure- ments take into account the tolerance given for the converging al- gorithm and the finite grid. Errors in the contrast magnitude only consider the uncertainty due to the tolerance of the optimization. To account for systematic uncertainties in the companion's loca- tion and contrast resulting from varying self-subtraction effects in reductions with different numbers of PCs, we ran the Hessian matrix algorithm for reductions with PCs in the range between 13 and 29 and considered the average of each parameter as fi- nal result. This range of PCs corresponds to FPF values below 2.5× 10−7 (see Figure B.1). To quantify the overall uncertainties in separation, PA, and contrast in a conservative way, we con- sidered the maximum/minimum value (including measurement errors) among the set of results for the specific parameter and computed its difference from the mean. In Figure 3, we present the results from this approach for the N_Ha dataset and show the comparison between the original residual image and the image with the companion successfully removed. 4.1.3. Astrometry The previously described algorithm was used to determine the best combination of separation, PA, and magnitude contrast for HD142527 B. In the N_Ha data the companion is located at 63.3+1.3−1.0 mas from the primary star, in the B_Ha dataset at 62.3+1.7−2.2 mas, and in the Cnt_Ha data at 62.8+2.1−1.9 mas. The cor- responding PAs are (97.8 ± 0.9)◦, (99.4+1.1−1.5)◦ and (99.0+1.5−1.6)◦, re- spectively. Errors in the PA measurements also take into account the above mentioned uncertainty in the astrometric calibration of the instrument, which was added in quadrature to the PA error bars. As within the error bars all filters gave the same results, we combined them and found that HD142527 B is located at a pro- jected separation of 62.8+2.1−2.7 mas from the primary star (9.9+0.3−0.4 AU at 157.3 ± 1.2 pc) and has a PA of (98.7 ± 1.8)◦. The final values result from calculating the arithmetic mean of all the val- ues obtained from the three different datasets, while their errors are calculated identically to those for each single dataset. In Figure 4 we compare the positions previously estimated (Close et al. 2014a; Rodigas et al. 2014; Lacour et al. 2016; Christiaens et al. 2018) and that resulting from our analysis. Lacour et al. (2016) used a Markov chain Monte Carlo analy- sis to infer the orbital parameters of HD142527 B. Because the past detections were distributed over a relatively small orbital arc (∼ 15◦), it was difficult to constrain the parameters precisely. The high precision measurement added by our SPHERE/ZIMPOL data extends the arc to a range of ∼ 30◦. An updated orbital analysis is provided in (?). Figure 4 shows that HD142527 B is clearly approaching the primary in the plane of the sky. 4.1.4. Photometry The Hessian matrix approach yields the contrasts between HD142527 A and B in every filter: ∆N_Ha = 6.3+0.2−0.3 mag in the 6 This is an absolute value, meaning that if the sum of the determinants can be lowered only using steps in pixels and contrast lower than 0.1, then the algorithm stops. Article number, page 7 of 20 Table 3. Summary of the stellar fluxes measured in the different filters in our ZIMPOL data and the derived Hα line fluxes for our targets (last column). The extinction values AHα were estimated as described in Section 4.1.4 from AV. A&A proofs: manuscript no. Sphere_Halpha AV [mag] Object <0.05a HD142527 (N_Ha) <0.05a HD142527 (B_Ha) HD142527 B (N_Ha) <0.05a HD142527 B (B_Ha) <0.05a 0.23a HD135344 B 0.0b TW Hya <0.05a HD100546 0.43c HD169142 0.22d MWC758 AHα [mag] 0.04 0.04 0.04 0.04 0.19 0.0 0.04 0.35 0.18 F∗ F_Hα [erg/s/cm2] 3.0 ± 0.8 × 10−11 9.7 ± 0.8 × 10−11 9.1+3.5−2.9 × 10−14 2.0 ± 0.4 × 10−13 3.1 ± 1.0 × 10−11 9.9 ± 0.4 × 10−11 4.2 ± 0.2 × 10−10 1.1 ± 0.1 × 10−10 8.1 ± 0.7 × 10−11 F∗ Cnt_Hα [erg/s/cm2] 6.1 ± 0.2 × 10−11 6.1 ± 0.2 × 10−11 7.4+1.4−2.1 × 10−14 7.4+1.4−2.1 × 10−14 4.9 ± 0.6 × 10−11 1.5 ± 0.05 × 10−11 1.6 ± 0.1 × 10−10 7.4 ± 0.2 × 10−11 5.3 ± 0.2 × 10−11 F∗ Hα [erg/s/cm2] 1.7 ± 0.8 × 10−11 1.0 ± 0.5 × 10−11 7.6+3.5−2.9 × 10−14 1.0+0.5−0.4 × 10−13 1.8 ± 0.8 × 10−11 7.8 ± 0.3 × 10−11 1.7 ± 0.2 × 10−10 3.2 ± 4.4 × 10−12 6.3 ± 3.7 × 10−12 References. (a) Fairlamb et al. (2015). (b) Uyama et al. (2017). (c) Fedele et al. (2017). (d) van den Ancker et al. (1998). Table 4. Summary of our detection limits for each target. While for HD100546, HD169142, and MWC 758 we consider the specific locations (separation and PA) of previously claimed companion candidates, we focused our analyses for HD135344B and TW Hya on separations related to disk gaps (hence no specific PA). Columns 5 and 6 give the mass and radius assumed for the accretion rate calculations, column 7 gives the contrast magnitude at the specific location and columns 8 -- 11 report the values for the Hα line flux, Hα line luminosity, accretion luminosity, and mass accretion rate ignoring any possible dust around the companion. Target Sep. [mas] PA [◦] HD100546 HD135344B 180 390 TW Hya 480 ± 4 ∼ 140 ∼ 340 156 ± 32 111 ± 4 HD169142 MWC 758 8.9 ± 0.9 ∼ 133 ∼ 175 7.4 ± 11.3 162 ± 5 Ref. Mass [MJ] 10.2(h) 2(k) 15(c) 15(l) 0.6(e) 10(f) 5.5(m) (a) (b) (c) (d) (e) (f) (g) Radius [RJ] 1.6(j) 1.3(j) 2(j) 2(j) 1.4(j) 1.7(j) 1.7(n) ∆Hα [mag] > 9.8 > 9.3 > 11.4 > 9.3 > 10.7 > 9.9 > 9.4 F p Hα [erg/s/cm2] < 3.8 × 10−15 < 1.9 × 10−14 < 1.1 × 10−14 < 7.9 × 10−14 < 5.7 × 10−15 < 1.2 × 10−14 < 1.4 × 10−14 LHα [L(cid:12)] < 2.0 × 10−6 < 2.2 × 10−6 < 4.7 × 10−6 < 3.3 × 10−5 < 2.5 × 10−6 < 5.2 × 10−6 < 1.2 × 10−5 Lacc [L(cid:12)] < 3.7 × 10−6 < 3.5 × 10−6 < 1.1 × 10−5 < 2.0 × 10−4 < 4.3 × 10−6 < 1.3 × 10−5 < 4.3 × 10−5 M [M(cid:12) yr−1] < 2.4 × 10−12 < 1.0 × 10−11 < 6.4 × 10−12 < 1.1 × 10−10 < 4.4 × 10−11 < 7.6 × 10−11 < 5.5 × 10−11 References. (a) Andrews et al. (2011) ; (b) Garufi et al. (2013) ; (c) Quanz et al. (2015) ; (d) Brittain et al. (2014) ; (e) Osorio et al. (2014) ; (f) Reggiani et al. (2014) ; (g) Reggiani et al. (2018) ; (h) Maire et al. (2017) , (j) AMES-Cond (Allard et al. 2001; Baraffe et al. 2003) , (k) Ruane et al. (2017) , (l) Mendigutía et al. (2017) , (m) Pinilla et al. (2015) , (n) BT-Settl (Allard et al. 2012) . narrow band, ∆B_Ha = 6.7 ± 0.2mag in the broad band, and ∆Cnt_Ha= 7.3+0.3−0.2 mag in the continuum filter. To quantify the brightness of the companion and not only its contrast with re- spect to the central star, we determined the flux of the primary in the multiple filters. We measured the count rate (cts) in the cen- tral circular region with radius ∼ 1(cid:48)(cid:48).5 in all frames of each stack √ n, where σ is and computed the mean and its uncertainty σ/ the standard deviation of the count rate within the dataset and n is the number of frames. No aperture correction was required because the same aperture size was used by Schmid et al. (2017) to determine the zero points for the flux density for the three fil- ters from photometric standard star calibrations. To estimate the continuum flux density we used their Equation 4 λ(Cnt_Ha) = cts · 100.4 (am·k1+mmode) · ccont F∗ where ccont zp (Cnt_Ha) is the zero point of the Cnt_Ha filter, cts = 1.105 (±0.001) × 105 ct/s is the count rate measured from our data, am = 1.06 is the average airmass, k1 is the atmospheric extinction at Paranal (k1(λ) = 0.085 mag/airmass for Cnt_Ha, k1(λ) = 0.082 mag/airmass for B_Ha and N_Ha; cf. Patat et al. zp (Cnt_Ha), (1) Article number, page 8 of 20 2011), and mmode = −0.23 mag is the mode dependent transmis- sion offset, which takes into account the enhanced throughput of the R-band dichroic with respect to the standard gray beam split- ter. The flux density of the primary star in the continuum filter F∗ λ(Cnt_Ha) was then used to estimate the fraction of counts in the line filters due to continuum emission via × 10−0.4(am·k1+mmode), cts(F_Ha) = (2) F∗ λ(Cnt_Ha) ccont zp (F_Ha) where ccont zp (F_Ha) is the continuum zero point of the Hα fil- ter used in the observations (cf. Schmid et al. 2017). During this step, we assumed that the continuum flux density was the same in the three filters. The continuum count rate was sub- tracted from the total count rate in B_Ha and N_Ha, cts(B_Ha) = 1.631 (±0.001)× 105 ct/s and cts(N_Ha) = 3.903 (±0.003)× 104 ct/s, leaving only the flux due to pure Hα emission. These were used, together with Equation (1) with line zero points, to deter- mine the pure Hα line fluxes (see fifth column in Table 3). For each filter, the continuum flux density was multiplied by the fil- ter equivalent width, and the flux contribution from line emission was added for the line filters. As in Sallum et al. (2015), we as- G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks Cnt_Ha of the primary in each individual ZIMPOL filter led to a larger difference when comparing the companion's apparent magnitude = 15.4±0.2 mag) with that from Close et al. in our work (mB = 15.8± 0.3 mag). Such values are possibly con- (2014a) (mB sistent within the typical variability of accretion of the primary and secondary at these ages. However, given the different pho- tometry sources and filters used for the estimation of the stellar flux densities in the two works, the results cannot be easily com- pared. Close = 1.0+0.5−0.4 × 10−13 erg/s/cm2 and f line 4.1.5. Accretion rate estimates The difference between the flux in the line filters and the con- tinuum filter (normalized to the Hα filter widths) represents the pure Hα line emission for which we find for HD142527 B = 7.6+3.5−2.9 × 10−14 f line B_Ha erg/s/cm2, respectively. The line flux is then converted into a line luminosity multiplying it by the GAIA distance squared (see Table 2), yielding LB_Ha = 7.7+4.0−3.6 × 10−5 L(cid:12) and LN_Ha = 6.0+2.8−2.4 × 10−5 L(cid:12). We then estimated the accretion luminosity with the classical T Tauri stars (CTTS) relationship from Rigli- aco et al. (2012), in which the logarithmic accretion luminosity grows linearly with the logarithmic Hα luminosity N_Ha log(Lacc) = b + a log(LHα), (3) and a = 1.49 ± 0.05 and b = 2.99 ± 0.16 are empirically deter- mined. We calculated the accretion luminosity for both datasets, = 5.0+4.4−4.0× 10−4L(cid:12). yielding Lacc Following Gullbring et al. (1998) we finally used N_Ha B_Ha (cid:32) = 7.3+6.8−6.4× 10−4L(cid:12) and Lacc (cid:33)−1 LaccRc ∼ 1.25 LaccRc GMc GMc Macc = 1 − Rc Rin (4) to constrain the mass accretion rate. The quantity G is the uni- versal gravitational constant, and Rc and Mc are the radius and mass of the companion, respectively. Assuming that the trun- cation radius of the accretion disk Rin is ∼ 5Rc, we obtain Fig. 3. Image of HD142527 before (top panel) and after (bottom panel) the insertion of the negative companion resulting from the Hessian ma- trix algorithm. The image flux scale is the same in both images. In this case 14 PCs were subtracted and a mask of 10.8 mas (radius) was ap- plied on the 101× 101 pixels images of the N_Ha stack. sumed the B object to have the same extinction as A, ignoring additional absorption from the disk. Indeed, we considered an extinction of AV = 0.05 mag (Fairlamb et al. 2015) and, interpo- lating the standard reddening law of Mathis (1990) for RV = 3.1, we estimated the extinction at ∼ 650 nm to be AHα = 0.04 mag. The stellar flux was found to be 6.1 ± 0.2 × 10−11 erg/s/cm2 in the Cnt_Ha filter, 9.7 ± 0.8 × 10−11 erg/s/cm2 in the B_Ha filter and 3.0 ± 0.8 × 10−11 erg/s/cm2 in the N_Ha filter (see Table 3). With the empirically estimated contrasts, we calculated the com- panion flux, i.e., line plus continuum emission or continuum only emission, in the three filters as follows: F p Cnt_Ha F p B_Ha F p N_Ha = 7.4+1.4−2.1 × 10−14 erg/s/cm2, = 2.0 ± 0.4 × 10−13 erg/s/cm2, = 9.1+3.5−2.9 × 10−14 erg/s/cm2. We note that the contrast we calculated in the continuum fil- ter is very similar to that obtained by Close et al. (2014a) of ∆mag = 7.5 ± 0.25 mag. The direct estimation of the brightness Fig. 4. Position of HD142527 B based on NaCo sparse aperture mask- ing (red pentagons), MagAO (cyan triangles), GPI non-redundant mask- ing (dark green diamonds) and VLT/SINFONI (blue circle) data from Rodigas et al. (2014), Close et al. (2014a), Lacour et al. (2016), and Christiaens et al. (2018), together with the SPHERE/ZIMPOL obser- vation presented in this work (light green square). The position of HD142527 A is shown with the yellow star at coordinates (0,0). Article number, page 9 of 20 -0.18-0.090.00.090.18Arcseconds-0.18-0.090.00.090.18ArcsecondsNEOriginal Image (N_Ha)-0.18-0.090.00.090.18Arcseconds-0.18-0.090.00.090.18ArcsecondsNEAfter subtraction (N_Ha)-0.200.20.40.60.81Flux (arbitrary linear scale)806040200 RA [mas]70605040302010010 DEC [mas]March 2012March 2013April 2013July 2013April 2014April 2014May 2014March 2016NaCo SAMMagAOGPI NMRVLT/SINFONISPHERE/ZIMPOL (cid:16) Rin A&A proofs: manuscript no. Sphere_Halpha (cid:17)−1 ∼ 1.25. For the companion mass and radius, two dif- 1 − Rc ferent sets of values were considered: Lacour et al. (2016) fit- ted the SED of HD142527 B with evolutionary models (Baraffe et al. 2003) and calculated Mc = 0.13 ± 0.03 M(cid:12) and Rc = 0.9 ± 0.15 R(cid:12), while Christiaens et al. (2018) estimated from H+K band VLT/SINFONI spectra Mc = 0.34 ± 0.06 M(cid:12) and Rc = 1.37 ± 0.05 R(cid:12), in the presence of a hot circumstellar en- vironment7. The accretion rates obtained from the Hα emission line are MB_Ha = 2.0+2.0−1.9 × 10−10M(cid:12)/yr and MN_Ha = 1.4+1.3−1.2 × 10−10M(cid:12)/yr in the first case and MB_Ha = 1.2±1.1×10−10M(cid:12)/yr and MN_Ha = 0.8 ± 0.7 × 10−10M(cid:12)/yr in the second case. Some Hα flux loss from the instrument when the N_Ha filter is used MN_Ha compared to MB_Ha. might explain the lower value of Indeed, according to Figure 2 and Table 5 from Schmid et al. (2017), the N_Ha filter is not perfectly centered on the Hα rest wavelength, implying that a fraction of the flux could be lost, in particular if the line profile is asymmetric. Moreover, high tem- perature and high velocities of infalling material cause Hα emis- sion profiles of CTTS to be broad (Hartmann et al. 1994; White & Basri 2003). Also, line broadening from the rotation and line shift of the object due to possible radial motion might be im- portant, even though it is not expected to justify the ∼40% Hα flux difference of HD142527B. We argue, therefore, that with the available data it is very difficult to estimate the amount of line flux lost by the N_Ha filter, and that the value given by the B_Ha filter is expected to be more reliable, since all line emis- sion from the accreting companion is included. As shown in PDI images from Avenhaus et al. (2017), dust is present at the separation of the secondary possibly fully em- bedding the companion or in form of a circumsecondary disk. During our calculations, we neglected any local extinction ef- fects due to disk material. It is therefore possible that on the one hand some of the intrinsic Hα flux gets absorbed/scattered and the actual mass accretion rate is higher than that estimated in this work; on the other hand, the material may also scatter some Hα (or continuum) emission from the central star, possibly con- tributing in very small amounts to the total detected flux. Although the results obtained in this work are on the same or- der of magnitude as those obtained by Close et al. (2014a), who derived a rate of 6 × 10−10 M(cid:12) yr−1, it is important to point out some differences in the applied methods. Specifically, Close et al. (2014a) used the flux estimated in the Hα filter to calcu- late LHα, while we subtracted the continuum flux and considered only the Hα line emission. Moreover, we combined the derived contrast with the stellar flux in the Hα filters obtained from our data, while Close et al. (2014a) used the R-band magnitude of the star. As HD142527 A is also accreting and therefore emitting Hα line emission, this leads to a systematic offset. Finally, Close et al. (2014a) used the relationship found by Fang et al. (2009) and not that from Rigliaco et al. (2012), leading to a difference in the LHα − Lacc conversion. section 4.1.1 using the N_Ha and the Cnt_Ha datasets and ap- plying ASDI. In addition to the 1(cid:48)(cid:48).08 × 1(cid:48)(cid:48).08 images we also ex- amined 2(cid:48)(cid:48).88 × 2(cid:48)(cid:48).88 images to search for accreting companions beyond the contrast limited region and beyond the spiral arms detected on the surface layer of the HD135344 B circumstellar disk. However, no signal was detected. We paid special attention to the separations related to the reported disk cavities (Andrews et al. 2011; Garufi et al. 2013). We chose to investigate specifi- cally the cavity seen in scattered light at 0(cid:48)(cid:48).18. The outer radius of the cavity seen in millimeter continuum is larger, but small dust grains are expected to be located inside of this radius in- creasing the opacity and making any companion detection more difficult. Neglecting the small inclination (i ∼ 11◦, Lyo et al. 2011), the disk is assumed to be face-on and the contrast value given by the curve of Figure 5 at 0(cid:48)(cid:48).18 is considered (∆N_Ha = 9.8 mag). We derived the Hα flux from the star in the N_Ha filter as presented in section 4.1.4 using the stellar flux values for the different filters given in Table 3, and calculated the upper limits for the companion flux, accretion luminosity, and mass accretion rate following Section 4.1.4 and Section 4.1.5. The accretion rate is given by Equation 4, assuming a planet mass of Mc = 10.2 MJ, the maximum mass that is nondetectable at those separations ac- cording to the analysis of Maire et al. (2017). Being consistent with their approach, we then used AMES-Cond8 evolutionary models (Allard et al. 2001; Baraffe et al. 2003) to estimate the radius of the object Rc = 1.6 RJ based on the age of the system. All values, sources, and models used are summarized in Table 3 and in Table 4 together with all the information for the other objects. The final accretion rate upper limit has been calculated to be < 2.4×10−12 M(cid:12) yr−1 at an angular separation of 0(cid:48)(cid:48).18, i.e., the outer radius of the cavity seen in scattered light. 4.3. TW Hya The TW Hya dataset does not show any point source either in the 1(cid:48)(cid:48).08× 1(cid:48)(cid:48).08 images (see Figure 6) or in the 2(cid:48)(cid:48).88× 2(cid:48)(cid:48).88 im- ages, which are large enough to probe all the previously reported disk gaps. The final contrast curves are shown in Figure 7. We 8 AMES-Cond and BT-Settl models used through the paper where downloaded on Feb. 06, 2018, from https://phoenix.ens- lyon.fr/Grids/AMES-Cond/ISOCHRONES/ and https://phoenix.ens- lyon.fr/Grids/BT-Settl/CIFIST2011_2015/ISOCHRONES/, respec- tively. 4.2. HD135344 B Visual inspection of the final PSF-subtracted ADI images of HD135344B showed a potential signal north to the star. Given the weakness of the signal and the low statistical significance, we analyze and discuss it further in Appendix D. In Figure 5 we plot the contrast curves obtained as explained in 7 They considered two different cases in which the companion may or may not be surrounded by a hot environment contributing in H+K. Because of the presence of accreting material shown in this work, we decided to consider the first case. Article number, page 10 of 20 Fig. 5. Contrast curves for HD135344 B. The vertical lines indicate the outer radii of the cavities in small and large dust grains presented in Garufi et al. (2013) and Andrews et al. (2011), respectively. 0.00.10.20.30.40.5Angular separation [as]456789101112Magnitude contrast for CL~99.99995% [mag]Scattered light cavity (Garufi+2013)Millimeter cavity (Andrews+2011)ADI N_HaASDI HaADI Cnt_HaDetection limits HD135344B G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks Fig. 6. Final PSF subtracted ADI images of TW Hya, HD100546, HD169142, and MWC 758. We applied a central mask with radius 32.4 mas and 18 PCs were removed. No companion candidates were detected. All images have a linear, but slightly different, color scale. also looked specifically at detection limits within the gaps ob- served by van Boekel et al. (2017) and focused in particular on the dark annulus at 20 AU (0(cid:48)(cid:48).39) from the central star, which has a counterpart approximately at the same position in 870 µm dust continuum observations (Andrews et al. 2016) Since the circumstellar disk has a very small inclination, we considered the disk to be face-on and assumed the gaps to be circular. At 0(cid:48)(cid:48).39, planets with contrast lower than 9.3 mag with respect to TW Hya would have been detected with the ASDI technique (cf. Figure 7). This value was then combined with the stellar flux calculated as described in section 4.1.4, to obtain the upper limit of the companion flux in the B_Ha filter. This yielded M < 1.0 × 10−11 M(cid:12) yr−1 (see Table 4) as the upper limit for the mass accretion rate based on our SPHERE/ZIMPOL dataset. 4.4. HD100546 The HD100546 dataset suffered from rather unstable and vary- ing observing conditions, which resulted in a large dispersion in the recorded flux (see Figure E.1 in Appendix E). We hence se- lected only the last 33% of the observing sequence, which had relatively stable conditions, for our analysis (see Appendix E). The Hα data did not confirm either of the two protoplanet candi- dates around HD100546 (see Figure 6) and we show the result- ing detection limits in Figure 8. In order to investigate the detection limits at the positions of the protoplanet candidates, we injected artificial planets with in- creasing contrast starting from ∆B_Ha = 8.0 mag until the signal was no longer detected with a CL of at least 99.99995%, and we repeated the process subtracting different numbers of PCs (from 10 to 30). At the position where Quanz et al. (2015) claimed the presence of a protoplanetary companion, we would have been able to detect objects with a contrast lower than 11.4 mag (us- ing PC=14 and the ADI reduction). Consequently, if existing, a 15 MJ companion (Quanz et al. 2015) located at the position of HD100546 b must be accreting at a rate < 6.4×10−12 M(cid:12) yr−1 in the framework of our analysis and assuming no dust is surround- ing the object. We note that, in comparison to the accretion lu- minosity Lacc estimated by Rameau et al. (2017), our upper limit is one order of magnitude lower (cf. Table 4). For the position of HD100546 c, we used the orbit given in Brittain et al. (2014) to infer the separation and PA of the candidate companion at the epoch of our observations, i.e., ρ (cid:39) 0(cid:48)(cid:48).14 and PA (cid:39) 133◦. At this position our data reach a contrast of 9.3 mag (using PC=14 on the continuum-subtracted dataset), implying an upper limit for the companion flux in the Hα filter of 7.9 × 10−14 erg/s/cm2 and Fig. 7. Contrast curves for TW Hya. The vertical line indicates the gap at 0(cid:48)(cid:48).39 detected in both scattered light (Akiyama et al. 2015; van Boekel et al. 2017) and submillimeter continuum (Andrews et al. 2016). a mass accretion rate < 1.1 × 10−10 M(cid:12) yr−1. This puts ∼ 2 or- ders of magnitude stronger constraints on the accretion rate of HD100546 c than the limits obtained from the polarimetric Hα images presented in Mendigutía et al. (2017) for a 15 MJ planet. We note that owing to its orbit, HD100546 c is expected to have just disappeared or to disappear quickly behind the inner edge of the disk (Brittain et al. 2014). Therefore, extinction could play a major role in future attempts to detect this source. 4.5. HD169142 We analyzed the data with ADI and ASDI reductions (see Figure 6 for the ADI image). The latter was particularly interesting in this case because the stellar flux density in the continuum and Hα filter is very similar and the continuum subtraction almost annihilated the flux from the central PSF, indicating that the central star has limited to no Hα line emission (cf. Table 3 and see Grady et al. (2007)). We calculated the detection limits as explained in section 4.1.1 for both filters for a confidence level of 99.99995%, as shown in Figure 9. We investigated with particular interest the positions of the candidates mentioned in Section 2 and derived specific detec- tion limits at their locations, independent from the azimuthally averaged contrast curve. At the position of the compact source Article number, page 11 of 20 -0.4-0.200.20.4Arcseconds-0.4-0.200.20.4ArcsecondsNETW Hya-0.4-0.200.20.4Arcseconds-0.4-0.200.20.4NEHD100546-0.4-0.200.20.4Arcseconds-0.4-0.200.20.4NEHD169142-0.4-0.200.20.4Arcseconds-0.4-0.200.20.4NEMWC 7580.00.10.20.30.40.5Angular separation [as]345678910Magnitude contrast for CL~99.99995% [mag]Scattered light cavity (Akyiama+2015, van Boekel+2016)Millimeter sized particle cavity (Andrews+2016)ADI Cnt_HaADI B_HaASDIDetection limits TW Hya A&A proofs: manuscript no. Sphere_Halpha Fig. 8. Contrast curves for HD100546. The gray dashed vertical line shows the separation of the outer gap edge cavity presented in Aven- haus et al. (2014a), while the solid blue lines indicate the separations of the forming planet candidates around HD100546 (Quanz et al. 2013a; Brittain et al. 2014). Fig. 9. Contrast curves for HD169142. The shaded region represents the annular gap observed in scattered light (Quanz et al. 2013b) and in millimeter continuum (Osorio et al. 2014). The blue vertical lines represent the separation of the companion candidates (Reggiani et al. 2014; Biller et al. 2014; Osorio et al. 2014). found by Osorio et al. (2014) (we call this potential source HD169142 c), our data are sensitive to objects 10.7 mag fainter than the central star (obtained by subtracting 16 PCs with ASDI reduction). At the position of HD169246 b (Reggiani et al. 2014; Biller et al. 2014) an object with a contrast as large as 9.9 mag could have been detected (PC=19; ASDI). For the com- pact source from Osorio et al. (2014) we found M < 4.4 × 10−11 M(cid:12) yr−1. Similarly, for the object detected by Biller et al. (2014) and Reggiani et al. (2014)9 we found an upper limit for the mass accretion rate of M < 7.6 × 10−11 M(cid:12) yr−1. 4.6. MWC 758 Our analysis of the SPHERE/ZIMPOL images did not show an Hα counterpart to the MWC 758 companion candidate detected by Reggiani et al. (2018) as shown in Figure 6. This is consistent with the recently published results from Huélamo et al. (2018). Nonetheless, we provide a detailed analysis and discussion of the same MWC 758 data to allow a comparison with the other datasets. In Figure 10 we show the detection limits obtained with ADI for the B_Ha and Cnt_Ha dataset, and the results of the ASDI approach. At separations larger than 0(cid:48)(cid:48).25, companions with a contrast smaller than 10 mag could have been detected. At the specific position of the candidate companion10 we can exclude objects with contrasts lower than 9.4 mag (obtained subtracting 15 PCs using ASDI). To explain the presence of a gap in dust-continuum emission without a counterpart in scattered light, a steady replenishment of µm-sized particle is required, which implies that a compan- ion in the inner disk should not exceed a mass of Mc = 5.5 MJ (Pinilla et al. 2015; Reggiani et al. 2018). In line with the anal- ysis of Reggiani et al. (2018), we used the BT-Settl model to estimate the radius of the companion and we derived an upper M < 5.5 × 10−11 M(cid:12) yr−1 limit for the mass accretion rate of (see Table 4). Our analysis puts slightly stronger constraints on 9 Within the uncertainties in the derived positions, these objects are indistinguishable and hence we assume it is the same candidate. 10 For our analysis we considered the position obtained from the first dataset in Reggiani et al. (2018) because the observing date was close to the epoch of the Hα observations. Article number, page 12 of 20 the mass accretion rate in comparison to that in Huélamo et al. (2018). 5. Discussion 5.1. SPHERE/ZIMPOL as hunter for accreting planets The SPHERE/ZIMPOL Hα filters allow for higher angular res- olution compared to filters in the infrared regime and can, in principle, search for companions closer to the star. For compar- ison, a resolution element is 5.8 times smaller in the Hα filter than in the L(cid:48) filter, meaning that the inner working angle (IWA) is smaller by the same amount so that closer-in objects could be observed, if bright enough11. An instrument with similar capa- bilities is MagAO (Close et al. 2014b; Morzinski et al. 2016), but as the Magellan telescope has a primary mirror of 6.5 m di- ameter, it has a slightly larger IWA than SPHERE at the 8.2 m VLT/UT3 telescope. A direct comparison of the HD142527 B detection shows that ZIMPOL reaches a factor ∼ 2.5 higher S/N in one-third of total integration time and field rotation of Ma- gAO under similar seeing conditions, even if the companion is located (cid:38) 20 mas closer to the star. The VAMPIRES instrument combined with Subaru/SCExAO will soon be a third facility able to perform Hα imaging in SDI mode (Norris et al. 2012) In terms of detection performance using different filters and reduction techniques, we re-emphasize that the N_Ha filter is more efficient in detecting Hα signals in the contrast limited regime. The smaller filter width reduces the contribution of the continuum flux, which often dominates the signal in the B_Ha filter, particularly for the central star. Hence, assuming the plane- tary companion emits only line radiation, the N_Ha filter reduces the contamination by the stellar signal in the remaining speckles. Moreover, the subtraction of the stellar continuum from Hα im- ages reduces the speckles in both B_Ha and N_Ha filters. Hence, ASDI enhances the signal of potential faint companions, in par- ticular at separations < 0(cid:48)(cid:48).3 (cf. Figures 7, 9, 10), where com- panions 0.7 mag fainter appear accessible in comparison to using simple ADI. ASDI should always be applied during the analysis of SPHERE/ZIMPOL Hα data. What remains to be quantified is 11 We note that SPHERE does not operate at similarly high Strehl ratios in the optical regime as it is able to do in the infrared. 0.00.10.20.30.40.5Angular separation [as]34567891011Magnitude contrast for CL~99.99995% [mag]Outer gap edge cavity (e.g., Avenhaus+2014)HD100546b (Quanz+2013)HD100546c (Brittain+2014)ADI Cnt_HaADI B_HaASDI HaDetection limits HD100546 G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks Fig. 10. Contrast curves for MWC 758. The gray dashed line shows the outer edge of the dust cavity observed by Andrews et al. (2011). The blue solid line indicates the separation at which Reggiani et al. (2018) found a candidate companion. how longer detector integration times (DITs) or the broad band filter could improve the detection limits in the background lim- ited regime (i.e., > 0(cid:48)(cid:48).3 where the contrast curves are typically flattening out) or for fainter natural guide stars. At these separa- tions narrow band data can be detector read noise limited and the B_Ha filter might be more suitable because of its higher through- put. However, as we show in Figure 11, it seems that at least for our HD142527 dataset this does not seem to be the case. Future studies conducted in both filters and on several objects are re- quired to derive a more comprehensive understanding. Finding the sweetspot between longer integration times and the smearing of the PSF because of field rotation is also warranted. At least for the object considered in Figure 11, at large separations (usually > 0(cid:48)(cid:48).3, in the background limited region) it is even possible to ignore completely ADI and simply apply field stabilized obser- vations. 5.2. Constraining planet accretion For our mass accretion rate estimates of HD142527 B we as- sumed that 100% of the Hα flux originates from accretion pro- cesses involving circumstellar material. We note, however, that the values may be overestimated if we consider that chromo- spheric activity of the M star (White & Basri 2003; Fang et al. 2009) can also contribute to the measured line flux. Furthermore, as mentioned in Section 4.1.5, we warn that the narrow width of the N_Ha filter might be too narrow to fully encompass all Hα line emission from fast-moving, accreting material, and there- fore the results may be underestimated. Finally, given the pres- ence of dusty material at the projected position of HD142527 B (Avenhaus et al. 2017), Hα flux might have been partially ab- sorbed. It is beyond the scope of this paper to properly estimate a value for intrinsic extinction due to disk material and consider this value in the M estimation. Nevertheless, in Figure 12 we show the fraction of Hα flux that is potentially lost because of extinction as a function of AV, converted into AHα as explained in Section 4.1.4. Only 2% of the Hα signal remains if the disk material causes an extinction of AV = 5 mag. This plot quantifies the impact of dust on the measured flux and the detectability of Hα emission from embedded objects. For the other five objects studied in this work we were not able to detect any clear accretion signature located in the disks. Fig. 11. Apparent flux detection limits as a function of the angular sep- aration from HD142527 for both B_Ha and N_Ha filters. Therefore, our data were not able to support the scenario in which protoplanets are forming in those disks. We put upper limits on the accretion luminosity and mass accretion rate. Two notes have to be made: (1) the fundamental quantities directly derived from the data are FHα and LHα; they should be used for future comparisons with other datasets or objects; (2) the pre- sented upper limits on M are only valid for an object with the mass and radius given in Table 4, while the Lacc upper limits re- fer to objects of any mass. In particular, assuming lower mass M, as shown in Figure 13: on the y-axis objects implies larger the mass accretion rate upper limits decrease as a function of the companion mass, for which the corresponding radius was cal- culated using the evolutionary models reported in Table 4 and assuming the age listed in Table 2. The plot highlights that the Macc by assumed mass of the companion may change the final more than one order of magnitude. Moreover, we overplot in vi- olet the mass accretion rates of the three objects presented in Zhou et al. (2014, see also Section 5.3) as well as LkCa15 b and PDS70 b (Sallum et al. 2015; Wagner et al. 2018), and in gray the range of mass accretion rates for HD142527. We stress that, similar to HD142527 B, we always assumed that the flux limit is completely due to Hα line emission without any contribution from continuum or chromospheric activity. Fur- thermore, for our analysis we always neglected intrinsic extinc- tion effects from disk material, which likely weaken the signal. In particular, at locations where no gap in small dust grains has been identified the extinction AHα can be significant (see Fig- ure 12). Models and precise measurements of the dust content in the individual disks would be required to properly include local extinction into our analysis. Finally, investigating the Hα lumi- nosity upper limits for the specific positions as a function of the separation from the central star, it can be noticed that the con- straints are stronger at larger separations. The only exception is HD100546, for which higher upper limits were achieved. The combination of suboptimal weather conditions, under which the dataset was taken, and the small field rotation of the subsample analyzed in this work made those limits worse. A more stable dataset with larger field rotation should provide more constrain- ing limits. 5.3. Comparison with other objects The accretion rate of HD142527 B is in good agreement with the mass accretion rates found in Rigliaco et al. (2012) for low-mass TTauri stars in the σ Ori star-forming region (5 × 10−11 M(cid:12) yr−1 < MCTTS < 10−9 M(cid:12) yr−1). A slightly broader mass accretion rate range was found by Alcalá et al. (2014), with Article number, page 13 of 20 0.00.10.20.30.40.5Angular separation [as]10141013Apparent flux [erg/s/cm2]B_HaN_Ha A&A proofs: manuscript no. Sphere_Halpha 2 × 10−12 M(cid:12) yr−1 < MCTTS < 4 × 10−8 M(cid:12) yr−1 in the Lupus star-forming region. Zhou et al. (2014) reported three very low-mass objects (GSC 06214-00210 b, GQ Lup b and DH Tau b), which ex- hibit Hα emission from accretion. Those objects have separa- tions of 100-350 AU from their parent stars and M ∼ 10−9 − 10−11 M(cid:12) yr−1 (see violet stars in Figure 13). The accretion rates measured in the paper are of the same order as the limits we found in our work. At projected distances similar to those of the three objects mentioned above, ZIMPOL would have been able to observe and detect Hα emitting companions. However, closer to the star in the contrast limited regime, our data would not have detected accretion processes occurring with M (cid:46) 10−11 M(cid:12) yr−1. The mass accretion rate of PDS70 b was estimated by Wag- ner et al. (2018) without considering any extinction effects and it is slightly lower than the limits we achieve for our sample (see violet square in Figure 13 and black star in Figure 14). The flux was calculated from the contrast in Wagner et al. (2018) assum- ing RPDS70 b = 11.7 mag and estimating the MagAO Hα filter Fig. 12. Fraction of Hα flux absorbed as a function of the disk extinction AV assuming the extinction law of Mathis (1990) as explained in Section 4.1.4. Fig. 13. Mass accretion rate upper limits as a function of the planetary mass for all the candidate forming planets investigated in this work. The violet stars represent the values reported in Zhou et al. (2014), while the violet squares indicate PDS70 b (Wagner et al. 2018) and LkCa15 b (Sallum et al. 2015). The gray shaded area represents the mass accretion rate of HD142527 B and is shown for mass accretion rate comparison purposes only. Indeed, the mass of the object is much larger than what is reported on the x-axis of the plot. Article number, page 14 of 20 widths assuming a flat SED12. In order to properly compare our limits and their Hα detection, the same confidence levels should be considered. We therefore estimated the contrast limit for a CL corresponding to a 4σ detection for HD142527 at the sepa- ration of PDS70 b, which was 0.3 mag lower than the limits cor- responding to a CL of 99.99995%. Hence, to bring all the con- trast curves from Figure 14 to a 4σ confidence level at ∼ 0(cid:48)(cid:48).19, a multiplication by a factor 0.76 is required. We note, however, that this scaling is just an approximation to provide a more direct comparison between the two studies. We also compared the Hα line luminosity upper limits ob- tained from our ZIMPOL Hα sample with that estimated by Sal- lum et al. (2015) for LkCa15 b (LHα ∼ 6 × 10−5 L(cid:12)). Our spe- cific limits for the candidates around HD169142, HD100546, and MWC 758 are slightly lower, but, except for HD100546 b and the compact source in HD169142 found by Osorio et al. (2014), of the same order of magnitude. LkCa15 itself was ob- served with SPHERE/ZIMPOL during the science verification phase in ESO period P96. We downloaded and analyzed the data, which were, however, poor in quality and also in terms of inte- gration time and field rotation. Only ∼ 1 hr of data is available with a field rotation of ∼ 16◦, a coherence time of 2.6 ± 0.8 ms, and a mean seeing of 1(cid:48)(cid:48).64±0(cid:48)(cid:48).37. As we show in Figure 14, with deeper observations including more field rotation, ZIMPOL can potentially detect the signal produced by LkCa15 b (Sallum et al. 2015) with a CL of 99.99995%. However, the higher air- mass at the Paranal Observatory and the fact that LkCa15 is a fainter guide star may complicate the redetection of the compan- ion candidate, and therefore exceptional atmospheric conditions are required. In addition to Hα also other spectral features like Paβ and Brγ lines may indicate ongoing accretion processes onto young objects. As an example, Daemgen et al. (2017) used the ab- sence of those lines in the spectrum of the low-mass compan- ion HD106906 b to infer its mass accretion rate upper limits ( M < 4.8 × 10−10 MJ/yr−1). Their constraint is stronger than the ones we were able to put with our ZIMPOL Hα data. Sev- eral other studies also detected hydrogen emission lines like Paβ from low-mass companions (e.g., Seifahrt et al. 2007; Bowler et al. 2011; Bonnefoy et al. 2014), but unfortunately they did not calculate mass accretion rates. 5.4. Comparison with existing models Two models for planetary accretion are currently used to explain the accreting phase of planet formation: magnetospheric accre- tion (Zhu 2015) and boundary layer accretion (Owen & Menou 2016). During magnetospheric accretion, the magnetic field trun- cates the CPD and hot ionized hydrogen in the closest regions of the disk falls onto the planet following the magnetic field lines. Recombination on the planet surface then produces Hα flux. For protoplanets, these models predict Hα luminosities at least three orders of magnitudes lower than in CTTS, according to equation 22 in Zhu (2015), LHα = 4.7 × 10−6L(cid:12) (cid:33)2(cid:32) (cid:32) RT RJ (cid:33) . vs 59km s−1 This is mainly owing to a one order of magnitude smaller in- fall velocity vs and a one order of magnitude smaller truncation radius RT (squared in the LHα equation) due to weaker mag- 12 https://visao.as.arizona.edu/software_files/visao/ html/group__reduction__users__guide.html#visao_filters 100101Mass [MJ]101210111010109Mass accretion rate [M/yr]HD142527 BGQ Lup bGSC 06214-00210 bDH Tau bPDS70 bLkCa15 bHD100546 cMWC 758 bHD169142 bHD100546 bHD169142 c G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks The comparison of LHα limits from Table 4 with Figure 7 from Mordasini et al. (2017) indicates that, assuming com- pletely cold accretion, the observed objects may be low-mass (0.1 − 1MJ) medium accreters ( M ∼ 10−10 − 10−9M(cid:12)/yr) or higher mass objects (1 − 15MJ) showing very little accretion ( M < 10−10.5M(cid:12)/yr). Mordasini et al. (2017) also suggested an- other possible reason for some of the nondetections in Hα. If some of the planets, such as HD100546 b, have not yet com- pletely detached from the disk, they would be cooler and would not be accreting at high accretion rates. In a later phase, they will possibly be able to open a gap and accrete a large amount of ma- terial. Another aspect that we did not consider is the effect of the circumplanetary disk inclination on the flux that is emitted. Zhu (2015) considered the disk inclination including a factor 1/ cos(i), where i is the CPD inclination. Detailed accretion mod- els should investigate the consequences of a tilted protoplanetary disk on LHα. 6. Conclusions Imaging in Hα is one of the promising techniques to detect forming planets at very small separations. In this context, the SPHERE/ZIMPOL instrument will play a major role in investi- gating local accretion signatures in circumstellar disks. An im- portant next step is to redetect the previous discoveries of Ma- gAO of Hα emission from LkCa15 b and PDS 70 b and to study potential accretion variability. None of the possible protoplanet candidates discovered in the infrared (HD169142 b, MWC758 b, and HD100546 b and c) could be confirmed in this study search- ing for accretion signatures, implying several possible scenarios. Their mass accretion rates could be lower than our limits and therefore they are currently not detectable. Other explanations are that protoplanetary accretion shows variability and some of the objects are currently going through a period of quiescence, or that extinction effects from disk material absorb a considerable fraction of the light. The study of NIR line diagnostics might reduce the effects of absorption and allow the detection of ac- cretion processes. Furthermore, it is possible that the observed candidates are disk features that have been enhanced by image post-processing (Follette et al. 2017; Ligi et al. 2018), or our un- derstanding of accretion processes during the formation of giant planets is not correct and, as an example, the use of the CTTS scaling relation is not correct. In order to investigate this, precise simulations of protoplanetary accretion, as well as of disk intrin- sic effects (via full radiative transfer), have to be developed and combined with multiwavelength observations spanning from the optical to the (sub)millimeter. The estimation of upper limits are of particular importance for the study of accretion variability of protoplanets in the fu- ture. Continuing surveys for accreting planets could possibly de- tect Hα signatures and combine these with detection limits pro- vided by this work to investigate variability in the accretion pro- cesses. Finally, we emphasize that although a lot of effort was put into the calculation of mass accretion rate upper limits, those values are model and parameter dependent. The Hα flux upper limits are, however, the fundamental quantities that were mea- sured from the data and can be directly compared with future observations. Acknowledgements. SPHERE is an instrument designed and built by a con- sortium consisting of IPAG (Grenoble, France), MPIA (Heidelberg, Ger- many), LAM (Marseille, France), LESIA (Paris, France), Laboratoire La- grange (Nice, France), INAF - Osservatorio di Padova (Italy), Observatoire de Genéve (Switzerland), ETH Zurich (Switerland), NOVA (Netherlands), ONERA Article number, page 15 of 20 Fig. 14. Detection limits in apparent flux obtained for a 99.99995% CL in this work, together with limits achieved with the available ZIMPOL dataset for LkCa15 b (red dashed line) and the result presented in Sal- lum et al. (2015) and Wagner et al. (2018). A deeper dataset is required to redetect LkCa15 b with ZIMPOL, but this detection is feasible. netic fields than in stars. We combined the magnetospheric ac- cretion models (Zhu 2015) with existing detections in the in- frared and evolutionary models. As an example, we present the case of HD100546 b. According to models (Zhu 2015), the ob- served L(cid:48) brightness could be emitted by a CPD with inner radius of 1−4 RJ and Mp M of 0.2−2.9×10−6 M2 J yr−1. The mass accre- tion constraints obtained from Hα ZIMPOL data would therefore imply that Mp (cid:38) 31 MJ. This result is in conflict with that ob- tained by Quanz et al. (2015) and the AMES-Cond evolutionary models, since the object L(cid:48) brightness excludes masses larger than ∼ 15 MJ. This is the mass expected in the case in which the L(cid:48) flux is only from photospheric emission. Moreover, a 30 MJ object would have significantly shaped the disk morphology and would have been clearly visible in other bands, such as the Ks-band, where Quanz et al. (2015) could only put upper limits to the companion brightness. Szulágyi & Mordasini (2017) found that only a minimal fraction of the hydrogen in CPDs might be thermally ionized if the planet is massive and hot enough. Consequently, the disk does not get truncated and ionized material does not get accreted through magnetospheric accretion along the field lines. Then, disk ma- terial falls directly onto the planet (boundary layer accretion). The same authors showed that material falling from the circum- stellar disk onto the CPD and the protoplanet shocks, and even- tually produces Hα line emission both from the CPD and the planet. The contribution to the Hα flux is larger from the CPD than from the planet (Szulágyi & Mordasini 2017). These au- thors also showed that the majority of the accreted gas, however, remains neutral, especially for planets < 10 MJ. Hence, the Hα flux can only estimate the ionized gas accretion rate and not the total accreted material. According to their simulations, a 10 MJ planet would be accreting at a rate of 5.7 × 10−8 MJ yr−1, pro- ducing LHα ∼ 7× 10−6 L(cid:12). This value is on the same order of the limits our data allow us to put on the Hα luminosity from known forming protoplanet candidates. Since considering lower plan- etary masses enhances the mass accretion rate (see equation 4) and higher masses should be visible in other infrared bands, we conclude that either extinction from disk material plays a major role in the nondetection of the existing candidates, or they are false positives resulting from image post-processing. 0.00.10.20.30.40.5Angular separation [as]101510141013101210111010Apparent flux [erg/s/cm2]LkCa15 b (Sallum+2015)PDS70 b (Wagner+2018) 3.9 detectionLkCa15HD100546TW HyaHD169142MWC 758HD135344BHD142527 BHa A&A proofs: manuscript no. Sphere_Halpha (France), and ASTRON (Netherlands), in collaboration with ESO. SPHERE also received funding from the European Commission Sixth and Seventh Frame- work Programmes as part of the Optical Infrared Coordination Network for As- tronomy (OPTICON) under grant number RII3-Ct-2004-001566 for FP6 (2004- 2008), grant number 226604 for FP7 (2009-2012), and grant number 312430 for FP7 (2013-2016). This work has been carried out within the frame of the National Center for Competence in Research PlanetS supported by the Swiss National Science Foundation. SPQ and HMS acknowledge the financial support of the SNSF. GC and SPQ thank the Swiss National Science Foundation for financial support under grant number 200021_169131. FMe and GvdP acknowl- edge fundings from ANR of France under contract number ANR-16-CE31-0013. This research has made use of the SIMBAD database, operated at CDS, Stras- bourg, France. This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www. cosmos.esa.int/web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions participating in the Gaia Multilateral Agreement. The authors thank Arianna Musso-Barcucci for the preliminary analysis on HD142527. 78 References Akiyama, E., Muto, T., Kusakabe, N., et al. 2015, ApJ, 802, L17 Alcalá, J. M., Natta, A., Manara, C. F., et al. 2014, A&A, 561, A2 Allard, F., Hauschildt, P. H., Alexander, D. R., Tamanai, A., & Schweitzer, A. 2001, ApJ, 556, 357 Allard, F., Homeier, D., & Freytag, B. 2012, Philosophical Transactions of the Royal Society of London Series A, 370, 2765 Amara, A. & Quanz, S. P. 2012, MNRAS, 427, 948 Amara, A., Quanz, S. P., & Akeret, J. 2015, Astronomy and Computing, 10, 107 Andrews, S. M., Wilner, D. J., Espaillat, C., et al. 2011, ApJ, 732, 42 Andrews, S. M., Wilner, D. J., Zhu, Z., et al. 2016, ApJ, 820, L40 Augereau, J. C., Lagrange, A. M., Mouillet, D., & Ménard, F. 2001, A&A, 365, Avenhaus, H., Quanz, S. P., Meyer, M. R., et al. 2014a, ApJ, 790, 56 Avenhaus, H., Quanz, S. P., Schmid, H. M., et al. 2017, AJ, 154, 33 Avenhaus, H., Quanz, S. P., Schmid, H. M., et al. 2014b, ApJ, 781, 87 Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, A&A, 402, 701 Benisty, M., Juhasz, A., Boccaletti, A., et al. 2015, A&A, 578, L6 Bertrang, G. H.-M., Avenhaus, H., Casassus, S., et al. 2018, MNRAS, 474, 5105 Beuzit, J.-L., Feldt, M., Dohlen, K., et al. 2008, in Proc. SPIE, Vol. 7014, Ground-based and Airborne Instrumentation for Astronomy II, 701418 Biller, B., Lacour, S., Juhász, A., et al. 2012, ApJ, 753, L38 Biller, B. A., Males, J., Rodigas, T., et al. 2014, ApJ, 792, L22 Boehler, Y., Ricci, L., Weaver, E., et al. 2018, ApJ, 853, 162 Bonnefoy, M., Chauvin, G., Lagrange, A.-M., et al. 2014, A&A, 562, A127 Boss, A. P. 1997, Science, 276, 1836 Bowler, B. P., Liu, M. C., Kraus, A. L., Mann, A. W., & Ireland, M. J. 2011, ApJ, Brittain, S. D., Carr, J. S., Najita, J. R., Quanz, S. P., & Meyer, M. R. 2014, ApJ, 743, 148 791, 136 alog, 2336 [arXiv:1803.09264] [arXiv:1806.11568] Grady, C. A., Muto, T., Hashimoto, J., et al. 2013, ApJ, 762, 48 Grady, C. A., Schneider, G., Hamaguchi, K., et al. 2007, ApJ, 665, 1391 Gullbring, E., Hartmann, L., Briceño, C., & Calvet, N. 1998, ApJ, 492, 323 Hartmann, L., Hewett, R., & Calvet, N. 1994, ApJ, 426, 669 Henden, A. A., Templeton, M., Terrell, D., et al. 2016, VizieR Online Data Cat- Huang, J., Andrews, S. M., Cleeves, L. I., et al. 2018, ApJ, 852, 122 Huélamo, N., Chauvin, G., Schmid, H. M., et al. 2018, ArXiv e-prints Keppler, M., Benisty, M., Müller, A., et al. 2018, ArXiv e-prints Kraus, A. L. & Ireland, M. J. 2012, ApJ, 745, 5 Lacour, S., Biller, B., Cheetham, A., et al. 2016, A&A, 590, A90 Lambrechts, M. & Johansen, A. 2012, A&A, 544, A32 Ligi, R., Vigan, A., Gratton, R., et al. 2018, MNRAS, 473, 1774 Lyo, A.-R., Ohashi, N., Qi, C., Wilner, D. J., & Su, Y.-N. 2011, AJ, 142, 151 Macintosh, B., Graham, J., Palmer, D., et al. 2006, in Proc. SPIE, Vol. 6272, So- ciety of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 62720L Maire, A.-L., Langlois, M., Dohlen, K., et al. 2016, in Proc. SPIE, Vol. 9908, Ground-based and Airborne Instrumentation for Astronomy VI, 990834 Maire, A.-L., Stolker, T., Messina, S., et al. 2017, A&A, 601, A134 Marleau, G.-D., Klahr, H., Kuiper, R., & Mordasini, C. 2017, ApJ, 836, 221 Marois, C., Lafrenière, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, ApJ, 641, 556 Marois, C., Lafrenière, D., Macintosh, B., & Doyon, R. 2008, ApJ, 673, 647 Mathis, J. S. 1990, ARA&A, 28, 37 Mawet, D., Milli, J., Wahhaj, Z., et al. 2014, ApJ, 792, 97 Meeus, G., Montesinos, B., Mendigutía, I., et al. 2012, A&A, 544, A78 Mendigutía, I., Oudmaijer, R. D., Garufi, A., et al. 2017, A&A, 608, A104 Mordasini, C., Marleau, G.-D., & Mollière, P. 2017, A&A, 608, A72 Morzinski, K. M., Close, L. M., Males, J. R., et al. 2016, in Proc. SPIE, Vol. 9909, Adaptive Optics Systems V, 990901 Müller, A., van den Ancker, M. E., Launhardt, R., et al. 2011, A&A, 530, A85 Muto, T., Grady, C. A., Hashimoto, J., et al. 2012, ApJ, 748, L22 Nelder, J. A. & Mead, R. 1965, The Computer Journal, 7, 308 Norris, B. R. M., Tuthill, P. G., Ireland, M. J., et al. 2012, in Proc. SPIE, Vol. 8445, Optical and Infrared Interferometry III, 844503 Osorio, M., Anglada, G., Carrasco-González, C., et al. 2014, ApJ, 791, L36 Owen, J. E. & Menou, K. 2016, ApJ, 819, L14 Patat, F., Moehler, S., O'Brien, K., et al. 2011, A&A, 527, A91 Perez, S., Dunhill, A., Casassus, S., et al. 2015, ApJ, 811, L5 Petit, C., Fusco, T., Charton, J., et al. 2008, in Proc. SPIE, Vol. 7015, Adaptive Optics Systems, 70151U Pineda, J. E., Quanz, S. P., Meru, F., et al. 2014, ApJ, 788, L34 Pinilla, P., de Juan Ovelar, M., Ataiee, S., et al. 2015, A&A, 573, A9 Pinte, C., Price, D. J., Ménard, F., et al. 2018, ApJ, 860, L13 Pohl, A., Benisty, M., Pinilla, P., et al. 2017, ApJ, 850, 52 Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 124, 62 Price, D. J., Cuello, N., Pinte, C., et al. 2018, MNRAS, 477, 1270 Qi, C., Ho, P. T. P., Wilner, D. J., et al. 2004, ApJ, 616, L11 Quanz, S. P., Amara, A., Meyer, M. R., et al. 2015, ApJ, 807, 64 Quanz, S. P., Amara, A., Meyer, M. R., et al. 2013a, ApJ, 766, L1 Quanz, S. P., Avenhaus, H., Buenzli, E., et al. 2013b, ApJ, 766, L2 Quanz, S. P., Schmid, H. M., Geissler, K., et al. 2011, ApJ, 738, 23 Rameau, J., Follette, K. B., Pueyo, L., et al. 2017, AJ, 153, 244 Rapson, V. A., Kastner, J. H., Millar-Blanchaer, M. A., & Dong, R. 2015, ApJ, Brittain, S. D., Najita, J. R., Carr, J. S., et al. 2013, ApJ, 767, 159 Calvet, N. & Gullbring, E. 1998, ApJ, 509, 802 Canovas, H., Ménard, F., Hales, A., et al. 2013, A&A, 556, A123 Christiaens, V., Casassus, S., Absil, O., et al. 2018, ArXiv e-prints 815, L26 [arXiv:1806.04792] Cleeves, L. I., Bergin, E. A., & Harries, T. J. 2015, ApJ, 807, 2 Close, L. M., Follette, K. B., Males, J. R., et al. 2014a, ApJ, 781, L30 Close, L. M., Males, J. R., Follette, K. B., et al. 2014b, in Proc. SPIE, Vol. 9148, Adaptive Optics Systems IV, 91481M Currie, T., Cloutier, R., Brittain, S., et al. 2015, ApJ, 814, L27 Daemgen, S., Todorov, K., Quanz, S. P., et al. 2017, A&A, 608, A71 Debes, J. H., Jang-Condell, H., Weinberger, A. J., Roberge, A., & Schneider, G. 2013, ApJ, 771, 45 Eisner, J. A. 2015, ApJ, 803, L4 Fairlamb, J. R., Oudmaijer, R. D., Mendigutía, I., Ilee, J. D., & van den Ancker, M. E. 2015, MNRAS, 453, 976 Fang, M., van Boekel, R., Wang, W., et al. 2009, A&A, 504, 461 Fedele, D., Carney, M., Hogerheijde, M. R., et al. 2017, A&A, 600, A72 Follette, K. B., Rameau, J., Dong, R., et al. 2017, AJ, 153, 264 Fukagawa, M., Tamura, M., Itoh, Y., et al. 2006, ApJ, 636, L153 Fukagawa, M., Tsukagoshi, T., Momose, M., et al. 2013, PASJ, 65, L14 Fusco, T., Sauvage, J.-F., Mouillet, D., et al. 2016, in Proc. SPIE, Vol. 9909, Adaptive Optics Systems V, 99090U Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, ArXiv e-prints [arXiv:1804.09365] Reggiani, M., Christiaens, V., Absil, O., et al. 2018, A&A, 611, A74 Reggiani, M., Quanz, S. P., Meyer, M. R., et al. 2014, ApJ, 792, L23 Rigliaco, E., Natta, A., Testi, L., et al. 2012, A&A, 548, A56 Rodigas, T. J., Follette, K. B., Weinberger, A., Close, L., & Hines, D. C. 2014, ApJ, 791, L37 Ruane, G., Mawet, D., Kastner, J., et al. 2017, AJ, 154, 73 Sallum, S., Follette, K. B., Eisner, J. A., et al. 2015, Nature, 527, 342 Schmid, H. M., Bazzon, A., Milli, J., et al. 2017, A&A, 602, A53 Schmid, H. M., Bazzon, A., Roelfsema, R., et al. 2018, ArXiv e-prints Seifahrt, A., Neuhäuser, R., & Hauschildt, P. H. 2007, A&A, 463, 309 Stolker, T., Bonse, M. J., Quanz, S. P., et al. 2018, ArXiv e-prints [arXiv:1808.05008] [arXiv:1811.03336] Stolker, T., Dominik, C., Avenhaus, H., et al. 2016, A&A, 595, A113 Szulágyi, J. & Mordasini, C. 2017, MNRAS, 465, L64 Szulágyi, J., Plas, G. v. d., Meyer, M. R., et al. 2018, MNRAS, 473, 3573 Teague, R., Bae, J., Bergin, E. A., Birnstiel, T., & Foreman-Mackey, D. 2018, ApJ, 860, L12 Uyama, T., Tanigawa, T., Hashimoto, J., et al. 2017, AJ, 154, 90 van Boekel, R., Henning, T., Menu, J., et al. 2017, ApJ, 837, 132 van den Ancker, M. E., de Winter, D., & Tjin A Djie, H. R. E. 1998, A&A, 330, Garufi, A., Quanz, S. P., Avenhaus, H., et al. 2013, A&A, 560, A105 145 Article number, page 16 of 20 G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks van der Marel, N., Cazzoletti, P., Pinilla, P., & Garufi, A. 2016, ApJ, 832, 178 Vicente, S., Merín, B., Hartung, M., et al. 2011, A&A, 533, A135 Wagner, K., Follete, K. B., Close, L. M., et al. 2018, ApJ, 863, L8 Walsh, C., Juhász, A., Pinilla, P., et al. 2014, ApJ, 791, L6 Weinberger, A. J., Anglada-Escudé, G., & Boss, A. P. 2013, ApJ, 762, 118 White, R. J. & Basri, G. 2003, ApJ, 582, 1109 Zacharias, N., Finch, C. T., Girard, T. M., et al. 2012, VizieR Online Data Cata- log, 1322 Zacharias, N., Monet, D. G., Levine, S. E., et al. 2004, in Bulletin of the Ameri- can Astronomical Society, Vol. 36, American Astronomical Society Meeting Abstracts, 1418 Zhou, Y., Herczeg, G. J., Kraus, A. L., Metchev, S., & Cruz, K. L. 2014, ApJ, 783, L17 Zhu, Z. 2015, ApJ, 799, 16 1 ETH Zurich, Institute for Particle Physics and Astrophysics, Wolfgang-Pauli-Strasse 27, CH-8093 Zurich, Switzerland 2 National Center of Competence in Research "PlanetS"(http:// 3 Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidel- nccr-planets.ch) berg, Germany 4 LESIA, CNRS, Observatoire de Paris, Université Paris Diderot, UPMC, 5 place J. Janssen, 92190 Meudon, France 5 Leiden Observatory, Leiden University, P.O. Box 9513, 2300 RA Leiden, The Netherlands 6 Univ. Grenoble Alpes, CNRS, IPAG, F-38000 Grenoble, France 7 Unidad Mixta International Franco-Chilena de Astronomia, CNRS/INSU UMI 3386 and Departemento de Astronomia, Univer- sidad de Chile, Casilla 36-D, Santiago, Chile 8 Geneva Observatory, University of Geneva, Chemin des Mailettes 51, 1290 Versoix, Switzerland 9 INAF - Osservatorio Astronomico dell'Osservatorio 5, 35122 Padova, Italy di Padova, Vicolo 10 Anton Pannekoek Astronomical Institute, University of Amsterdam, PO Box 94249, 1090 GE Amsterdam, The Netherlands 11 Space Telescope Science Institute, Baltimore 21218, MD, USA 12 Aix Marseille Université, CNRS, CNES, LAM, Marseille, France 13 Department of Astronomy, Stockholm University, AlbaNova Uni- versity Center, 106 91 Stockholm, Sweden 14 Centre de Recherche Astrophysique de Lyon, CNRS/ENSL Univer- sité Lyon 1, 9 av. Ch. André, 69561 Saint-Genis-Laval, France 15 CNRS, IPAG, 38000 Grenoble, France 16 The University of Michigan, Ann Arbor, Mi 48109, USA 17 European Southern Observatory, Alonso de Cordova 3107, Casilla 19001 Vitacura, Santiago 19, Chile 18 Physikalisches Institut, Universität Bern, Gesellschaftsstrasse 6, 3012 Bern, Switzerland 19 Monash Centre for Astrophysics (MoCa) and School of Physics and Astronomy, Monash University, Clayton Vic 3800, Australia 20 NOVA Optical Infrared Instrumentation Group at ASTRON, Oude Hoogeveensedijk 4, 7991 PD Dwingeloo, The Netherlands 21 Institute for Computational Science, University of Zurich, Win- terthurerstrasse 190, 8057 Zurich, Switzerland 22 Núcleo de Astronomía, Facultad de Ingeniería y Ciencias, Universi- dad Diego Portales, Av. Ejercito 441, Santiago, Chile 23 Escuela de Ingeniería Industrial, Facultad de Ingeniería y Ciencias, Universidad Diego Portales, Av. Ejercito 441, Santiago, Chile effect because the Nasmyth mirror of the VLT introduces an in- strument polarization of about 4 %. This is reduced by the first mirror in SPHERE to about 0 to 3 %, while the following mir- rors in the instrument add further positive or negative polariza- tion contributions of about 2 %, while polarization cross talks (linear → circular polarization) reduce the linear polarization. Thus, it is safe to adopt a maximum error of 5 % for the rela- tive difference (e.g., T(Hα) = (1 ± 0.05)T(cont)) in throughput between the two channels. We therefore tested the impact of an enhancement/decrease by at most 5% in the continuum flux, an- alyzing the consequences on the detection of HD142527B and on the contrast performances of our pipeline. The signal flux is measured in an aperture of radius 8.3 mas, and the FPF was cal- culated as explained in section 4.1.1. The results averaged over a range of PCs (PC=10,15,20,25,30) on the ASDI-processed B_Ha dataset are shown in Table A.1; a central mask of 21.6 mas was applied. As one may expect, the signal flux shows a strong variation of 20 − 50% in the ASDI images, which is mainly due to the stronger/weaker subtraction of the continuum. The rela- tive difference in this case is increased from the initial 5% by the ASDI processing, but it should be noted that together with the signal, the noise level also gets increased/decreased, caus- ing the FPF to be less subject to variations. Indeed, regarding the FPF values, we argue that depending on the arm 1 to arm 2 transmission the confidence of the detection is lower by a factor of ∼10 in both extreme cases, which corresponds approximately to a maximum variation of ∼0.1 mag in ∆mag. Therefore, we do not expect this effect to have a large impact on the detec- tion limits estimated in this work. Nevertheless, it is important to keep in mind that the calculation of the mass accretion rate of HD142527 B does not consider this effect and a more accurate description of the instrument behavior is required to correct for it. Appendix B: FPF analysis of the HD142527 B dataset order to for the identify future strategy best In SPHERE/ZIMPOL observations, we compared the FPF calculated after the subtraction of different numbers of PCs using different techniques and datasets. For the ADI technique we considered three datasets: B_Ha, N_Ha, and Cnt_Ha, of which the last set contained all the images taken with the con- tinuum filter. For SDI and ASDI, we considered the subtraction of the Cnt_Ha images from the respective Hα filter images. All the images had a size of 1(cid:48)(cid:48).08 × 1(cid:48)(cid:48).08. For each case, we applied an inner mask of varying size (10.8 mas, 21.6 mas, 32.4 mas) and chose the smallest FPF value as representative value for the detection. The FPF calculation (see Section 4.1.1) followed the prescription suggested in Mawet et al. (2014). Because of the strong negative wings of the companion in the PSF subtracted images, we decided not to consider the e-mail: [email protected] Appendix A: Influence of the beamsplitter on flux measurements We investigated the effects of the throughput uncertainties of the two ZIMPOL arms resulting from instrument polarization ef- fects. It is currently not known how the overall throughput to the individual ZIMPOL arms depends on the telescope and instru- ment configurations. However, it is easy to estimate the overall Table A.1. Resulting signal flux and FPF for different beamsplitter behaviors. Target Cnt flux 5% decreased Cnt flux not changed Cnt flux 5% enhanced Signal flux (×1000, arb. unit) 5.45 ± 0.06 3.62 ± 0.02 2.84 ± 0.03 FPF (×1012) 9.97 ± 4.72 1.69 ± 1.47 11.17 ± 2.86 Article number, page 17 of 20 A&A proofs: manuscript no. Sphere_Halpha Fig. B.1. Performance comparison for Hα imaging with SPHERE/ZIMPOL using different filters (narrow and broad Hα) and reduction techniques (ADI and ASDI) based on the HD142527 dataset. In all panels the FPF obtained for HD142527 B is shown, as a function of the number of subtracted PCs used in PynPoint. On the right side of each panel we give the scale of the S/N to improve understanding of the plot and to compare different instrumental setups. We note again that this does not correspond to the classical σ notation. The gray regions indicates a confidence level for the detection of HD142527 B of at least 99.99995%, i.e., >5σ in case of Gaussian noise. Because of the applied corrections for small sample statistics, the border of the gray area does not correspond to an S/N of 5. two background apertures closest to the signal as they are not representative of the background and speckle noise. In the four panels of Figure B.1 we analyzed the FPFs of HD142527 B, obtained using different combinations of tech- niques and datasets. In the top panels we compare the detec- tion from different filters using the same technique: ADI on the left and ASDI on the right. For the ADI analysis, the B_Ha and Cnt_Ha datasets show similar values with a stronger detection in B_Ha for fewer subtracted components, while the FPF values obtained with the N_Ha filter are, for a wide range of PCs (from 11 to 32), ∼ 5 orders of magnitude lower. The detections with the ASDI technique show a similar trend; there is a stronger de- tection in N_Ha, particularly between 10 and 27 PCs. The nor- mal SDI technique, which is not presented in the image, was not efficient enough to properly subtract the stellar PSF and did not reveal the companion. This is probably for two reasons: (1) the central star is actively accreting material and emitting strong Hα flux, which cannot be subtracted accurately with the Cnt_Ha images, impeding the detection of the companion, and (2) PSF shapes are slightly different for different filters due to nonmatch- ing bandpasses. In the lower panels, we consider the results from the B_Ha (left) and N_Ha (right) datasets for ADI and ASDI. In both cases ASDI seems to be more efficient in detecting signals. A larger gain is obtained for the B_Ha filter, while FPFs obtained with the N_Ha filter have more similar values, probably due to the minor impact of the continuum subtraction on images taken with the narrow filter with respect to the broad filter. We conclude that the best observing strategy to look for accreting compan- ions in the contrast limited regime with SPHERE/ZIMPOL is to take images in the N_Ha filter and Cnt_Ha filter simultane- Article number, page 18 of 20 ously and to perform ASDI. It is of particular interest that in the case of HD142527, ASDI also performs better than ADI. In- deed, we could expect that the presence of a clear signal in the continuum would have strongly compromised the detection with ASDI. On the contrary, the detection is even stronger, implying that the subtraction of the stellar pattern is much more impor- tant than the self-subtraction of the companion, boosting its S/N. We note, however, that observing fainter objects might cause the data to be readout noise limited. In this case, the B_Ha filter might be preferred to the N_Ha filter. This hypothesis, however, should be confirmed with a fainter source than the bright M- dwarf HD142527B. Appendix C: Impact of field rotation and total integration time In addition to the best instrumental setup for the detection of ac- creting objects in Hα imaging, we also investigated the effect of two observational parameters on the achieved upper limits: the field rotation and integration time on target. Two subsets were created from the HD142527 B_Ha data. The first was composed of every second frame of the dataset, while the second only in- cluded the first half of the dataset frames. In this way, the field ro- tation of the first subset is twice that of the second subset, while the integration time is the same for the two subsets. Figure C.1 shows the resulting contrast curves, calculated in the same way as described in Section 4.1.1. The dashed blue line represents the subset composed of the first half of the dataset, which allows us to reach ∼ 9.4 mag of contrast at 0(cid:48)(cid:48).2. The green solid line shows the contrast limits estimated from the subset composed of every G. Cugno et al.: A search for accreting young companions embedded in circumstellar disks second frame. It is clear that at all separations, this subset allows us to detect fainter objects than the other subset and at 0(cid:48)(cid:48).2 the difference reaches 1.1 mag. Finally, the entire dataset allows us to go, at the same distance, another 0.3 mag deeper. At least for this dataset, the field rotation seems to play a very important role, allowing a better modeling and subtraction of the stellar PSF. Fig. C.1. Contrast curves calculated for the "first" (blue dashed line) and the "every second frame" subsets (green solid line), and for the entire B_Ha dataset of HD142527 (orange dotted line). Appendix D: Is a companion candidate orbiting HD135344B? We visually inspected the final PSF-subtracted ADI images of HD135344 B, which showed a potential signal north of the star. The feature is persistent in the N_Ha and Cnt_Ha datasets for different mask radii (e.g., 0(cid:48)(cid:48).02, 0(cid:48)(cid:48).03, 0(cid:48)(cid:48).04, 0(cid:48)(cid:48).05, and 0(cid:48)(cid:48).06) and over a wide range of PCs (6-21). In particular, when us- ing larger mask radii, the close-in speckles are removed and the signal appears to be stronger. We then investigated smaller im- ages (101×101 pixels) with the same technique and confirmed the signal for different mask radii and PCs. Next, we examined the ASDI images and found that the signal is present once again in different reductions, but appears fainter. If the signal is from a physical source, this is expected from an accreting object emit- ting Hα line radiation. This signal is shown in Figure D.1 in the N_Ha and in the Cnt_Ha filter for the parameter setups that yield the lowest FPFs, which are 5.9 × 10−5 for N_Ha and 0.0015 for Cnt_Ha. A careful look at the bright signal in the Cnt_Ha images raises doubts on the nature of its source as it is very compact and does not have a PSF-like shape. Furthermore, the signal has high FPFs, with a minimal value of ∼ 0.0015, which is not statisti- cally significant enough to claim a detection. The signal in the N_Ha frames has a morphology resembling that of a faint physical source. Varying the number of PCs seems to influence the apparent shape and location of the signal, as ex- pected from faint close-in objects with low S/N when subtracting the stellar PSF. Even though for 9 PCs the FPF reaches a min- imal value of ∼ 5.9 × 10−5, the FPFs for 7-17 PCs are in the range 10−3 − 10−2, which does not give us sufficient confidence to claim a detection. As a final check, we used the Hessian matrix approach as described in Section 4.1.2 to perform a signal characterization. We ran the algorithm for PCs between 7 and 17 (where the final Fig. D.1. Lowest FPF images of HD135344 B (top panel: N_Ha filter; bottom panel: Cnt_Ha filter). The radius r of the inner mask and the number of subtracted PCs are given in each panel. The location of the tentative companion candidate is indicated by the arrow (see Section 4.2.1.). images showed a clear signal) with a central mask of radius 57.6 mas and a ROI of 8 × 8 pixels. The other parameters were kept identical to the analysis performed on HD142527 B. The signal appears to be located at a separation of 71.1+4.8−4.2 mas with a PA of (19.1+2.2−2.8)◦. The contrast was measured to be 8.1 ± 0.4 mag. As visible in the error bars, the positions found are spread over a range of ∼ 9 mas, which corresponds to ∼ 2.5 pixels. Normally, a physical point source should be less affected by systematics introduced by the PSF subtraction process. However, a low S/N object at a separation of 71.1 mas is more difficult to measure properly and a larger spread in the recovered positions could be the result. A similar note can be made for the contrast values, which span over a range of ∼ 0.8 mag. We conclude that to settle this issue and fully understand the origin of the signal in the Hα filter, a dataset with higher S/N would be required. Article number, page 19 of 20 0.00.10.20.30.40.5Angular separation [as]56789101112Magnitude contrast [mag]ADI first half subsetADI every second frameADI entire datasetDetection limits HD142527 B_Ha-0.20.100.10.2Arcseconds-0.20.100.10.2ArcsecondsPC = 9r = 54 masNENarrow Band Ha (656.53 nm)-0.20.100.10.2Arcseconds-0.20.100.10.2ArcsecondsPC = 14r = 64.8 masNEContinuum Band (644.9 nm)-0.4-0.200.20.40.60.81Flux (arbitrary linear scale) A&A proofs: manuscript no. Sphere_Halpha Appendix E: Frame selection for the HD100546 dataset As briefly described in Section 4.4, the large HD100546 dataset (1104 frames, cf. Table 1) was taken in unstable conditions, which made a frame selection necessary. To determine a frame selection metric, we plotted the mean count value per image (im- age dimensions 1(cid:48)(cid:48).08×1(cid:48)(cid:48).08 pixels, see Figure E.1). It turned out that three phases could be identified within the observing run: a short initial phase of stability with some outliers (120 frames), a long period of 619 frames where the mean count values spanned a range between ∼ 0 and ∼ 55 counts per pixels, and, finally, a large amount of stable frames at the end of the observations (mean pixel value ∼ 55 in B_Ha). We decided only to keep the images of the last stable period, composed of the frames 739 − 1104 to perform our analysis. This subsample has a to- tal on-target integration time of 61 minutes and its field rotation is ∼ 20.7◦. Fig. E.1. B_Ha (blue circles) and Cnt_Ha (green crosses) mean count rates as a function of the image number in the observing sequence. The shaded region at the end represents the subset of frames that was chosen for the analysis. Article number, page 20 of 20
0912.2359
2
0912
2010-05-21T00:43:47
Day-side z'-band emission and eccentricity of Wasp-12b
[ "astro-ph.EP" ]
We report the detection of the eclipse of the very-hot Jupiter WASP-12b via z'-band time-series photometry obtained with the 3.5-meter ARC telescope at Apache Point Observatory. We measure a decrease in flux of 0.082+/-0.015% during the passage of the planet behind the star. That planetary flux is equally well reproduced by atmospheric models with and without extra absorbers, and blackbody models with f > 0.585+/-0.080. It is therefore necessary to measure the planet at other wavelengths to further constrain its atmospheric properties. The eclipse appears centered at phase = 0.5100 (+0.0072,-0.0061), consistent with an orbital eccentricity of |e cos w| = 0.016 (+0.011,-0.009) (see note at end of Section 4). If the orbit of the planet is indeed eccentric, the large radius of WASP-12b can be explained by tidal heating.
astro-ph.EP
astro-ph
Day-side z' -- band emission and eccentricity of WASP-12b1 Mercedes L´opez-Morales2,4, Jeffrey L. Coughlin3,5, David K. Sing6, Adam Burrows7, D´aniel Apai8, Justin C. Rogers4,9, David S. Spiegel7 & Elisabeth R. Adams10 e-mail: [email protected] ABSTRACT We report the detection of the eclipse of the very-hot Jupiter WASP-12b via z ′-band time-series photometry obtained with the 3.5-meter ARC telescope at Apache Point Observatory. We measure a decrease in flux of 0.082±0.015% during the passage of the planet behind the star. That planetary flux is equally well reproduced by atmospheric models with and without extra absorbers, and blackbody models with f ≥ 0.585±0.080. It is therefore necessary to measure the planet at other wavelengths to further constrain its atmospheric properties. The eclipse appears centered at phase φ = 0.5100+0.0072 −0.0061, consistent with an orbital eccentricity of e cos ω = 0.016+0.011 −0.009 (see note at end of §4). If the orbit of the planet is indeed eccentric, the large radius of WASP-12b can be explained by tidal heating. 1Based on observations collected with the Apache Point Observatory 3.5-meter telescope, which is owned and operated by the Astrophysical Research Consortium (ARC). 2Hubble Fellow 3NSF Graduate Research Fellow 4Carnegie Institution of Washington, Department of Terrestrial Magnetism, 5241 Broad Branch Rd. NW, Washington D.C., 20015, USA 5Department of Astronomy, New Mexico State University, Las Cruces, NM 88003, USA 6Astrophysics group, School of Physics, University of Exeter, Stocker Road, Exeter, Ex4 4QL, United Kingdom 7Princeton University, Department of Astrophysical Sciences, Peyton Hall, Princeton, NJ, 08544, USA 8Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 9Johns Hopkins University, Department of Physics and Astronomy, 366 Bloomberg Center, 3400 N. Charles Street, Baltimore, MD 21218, USA 10Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge, MA, 02139 -- 2 -- Subject headings: planetary systems -- stars: niques: photometric individual: WASP-12 -- tech- 1. Introduction The transiting Very Hot Jupiter WASP-12b, discovered by Hebb et al. (2009), has many notable characteristics. With a mass of 1.41 ± 0.10 MJ up and a radius of 1.79 ± 0.09 RJ up, WASP-12b is the planet with the second largest radius reported to date. It is also the most heavily irradiated planet known, with an incident stellar flux at the substellar point of over 9×109 erg cm−2 s−1. In addition, model fits to its observed radial velocity and transit light curves suggest that the orbit of WASP-12b is slightly eccentric. All these attributes make WASP-12b one of the best targets to test current irradiated atmosphere and tidal heating models for extrasolar planets. In irradiated atmosphere model studies WASP-12b is an extreme case even in the cat- egory of highly irradiated gas giants. Such highly irradiated planets are expected to show thermal inversions in their upper atmospheric layers (Burrows et al. 2008), although the chemicals responsible for such inversions remain unknown. TiO and VO molecules, which can act as strong optical absorbers, have been proposed (Hubeny et al. 2003; Fortney et al. 2008), but D´esert et al. (2008) claim that the concentration of those molecules in planetary atmospheres is too low (< 10−3 − 10−2 times solar). Spiegel et al. (2009) argue that TiO needs to be at least half the solar abundance to cause thermal inversions, and very high levels of macroscopic mixing are required to keep enough TiO in the upper atmosphere of planets. S2, S3 and HS compounds have also recently been suggested and then questioned as causes of the observed thermal inversions (Zahnle et al. 2009a,b). In the case of tidal heating, detailed models are now being developed (e.g. Bodenheimer et al. 2003; Miller et al. 2009; Ibgui et al. 2009,b; Ibgui & Burrows 2009) to explain the inflated radius phenomenon observed in hot Jupiters, of which WASP-12b, with a radius over 40% larger than predicted by standard models, is also an extreme case. All models assume that the planetary orbits are slightly eccentric, and directly measuring those eccentricities is key not only to test the model hypotheses, but also to obtain information about the planets' core mass and energy dissipation mechanisms (see e.g. Ibgui et al. 2009). Here we present the detection of the eclipse of WASP-12b in the z ′-band (0.9 µm), which gives the first measurement of the atmospheric emission of this planet, and the first direct estimation of its orbital eccentricity. Section 2 summarizes the observations and analysis of -- 3 -- the data. In Section 3 we compare the emission of the planet to models. The results are discussed in Section 4. 2. Observations and Analyses We monitored WASP-12 [RA(J2000)=06:30:32.794, Dec(J2000)=+29:40:20.29, V=11.7] during two eclipses, and under photometric conditions, on February 19 and October 18 2009 UT. An additional attempt on October 30 2009 UT was lost due to weather. The data were collected with the SPICam instrument on the ARC's 3.5-meter telescope at Apache Point Observatory, using a SDSS z′ filter with an effective central wavelength of ∼0.9µm. SPICam is a backside-illuminated SITe TK2048E 2048x2048 pixel CCD with 24 micron pixels, giving an unbinned plate scale of 0.14 arc seconds per pixel and a field of view of 4.78 arc minutes square. The detector, cosmetically excellent and linear through the full A/D converter range, was binned 2x2, which gives a gain of 3.35 e−/ADU, a read noise of 1.9 DN/pixel, and a 48 second read time. On February 19 we monitored WASP-12 from 3:00 to 3:28 UT and from 3:54 to 7:10 UT, losing coverage between 3:28 and 3:54 UT when the star reached a local altitude greater than 85◦, the soft limit of the telescope at that time. These observations yielded 1.20 hours of out-of-eclipse and 2.45 hours of in-eclipse coverage, at airmasses between 1.005 -- 1.412. On October 18 we extended the altitude soft limit of the telescope to 87◦ and covered the entire eclipse from 7:05 to 12:45 UT, yielding 2.73 hours of out-of-eclipse and 2.93 hours of in-eclipse coverage, with airmasses between 1.001 -- 1.801. In both nights we defocused the telescope to a FWHM of ∼ 2′′ to reduce pixel sensitivity variation effects, and also to allow for longer integration times, which minimized scintillation noise and optimized the duty cycle of the observations. Pointing changed by less than (x,y)=(4,7) pixels in the October 18 dataset, and by less than (x,y)=(3,12) pixels on February 19, with the images for this second night suffering a small gradual drift in the y direction throughout the night. Integration times ranged from 10 to 20 seconds. Taking into account Poisson, readout, and scintillation noise, the photometric precision on WASP-12 and other bright stars in the images ranged between 0.07 -- 0.15% per exposure on February 19, and between 0.05 -- 0.09% per exposure on October 18. The field of view of SPICam was centered at RA(J2000)=06:30:25, Dec(J2000)=+29:42:05 and included WASP-12 and two other isolated stars at RA(J2000)=06:30:31.8, Dec(J2000)=+29:42:27 and RA(J2000)=06:30:22.6, Dec(J2000)=+29:44:42, with apparent brightness and B−V and J − K colors similar to the target. Each night's dataset was analyzed independently and the results combined in the end. The timing information was extracted from the headers of the -- 4 -- images and converted into Heliocentric Julian Days using the IRAF task setjd, which has been tested to provide sub-second timing accuracy. We corrected each image for bias-level and flatfield effects using standard IRAF routines. Dark current was neglegible. DAOPHOT-type aperture photometry was performed in each frame. We recorded the flux from the target and the comparison stars over a wide range of apertures and sky background annuli around each star. We used apertures between 2 and 35 pixels in one-pixel steps during a first preliminary photometry pass, and 0.05 pixel steps in the final photometric extraction. To compute the sky background around each star we used variable width annuli, with inner radii between 35 and 60 pixels sampled in one-pixel steps. The best aperture and sky annuli combinations were selected by identifying the most stable (i.e. minimum standard deviation), differential light curves between each comparison and the target at phases out-of-eclipse1. In the February 19 data, the best photometry results from an aperture radius of 14.7 pixels for both the target and the comparison stars, and sky annuli with a 52-pixel inner radius and 22-pixel wide. For the October 18 data, 17.9 pixel apertures and sky annuli with a 45-pixel inner radius and 22-pixel wide produce the best photometry. The resultant differential light curves between the target and each comparison contain systematic trends that can be attributed to either atmospheric effects, such as airmass, see- ing, or sky brightness variations, or to instrumental effects, such as small changes in the location of the stars on the detector. Systematics can also be introduced by instrumental temperature or pressure changes, but those parameters are not monitored in SPICam. We modeled systematics for each light curve by fitting linear correlations between each param- eter (airmass, seeing, sky brightness variations, and target position) and the out-of-eclipse portions of the light curves. All detected trends are linear and there are no apparent residual color difference effects. The full light curves are then de-trended using those correlation fits. In the October 18 dataset, airmass effects are the dominant systematic, introducing a linear baseline trend with an amplitude in flux of 0.07%. The February 19 dataset also shows systematics with seeing and time with a total amplitude of also 0.07%. The systematics on this night were modeled using only the after-eclipse portion of the light curve, and we consider this dataset less reliable that the October 18 one. The 18 pre-ingress images col- lected between 3:00 and 3:38 UT suffer from a ∼50 pixel position shift with respect to the rest of the images collected that night, which cannot be modeled using overall out-of-eclipse 1We had to iterate on the out-of-eclipse phase limits after finding that the eclipse was centered at φ = 0.51. Out-of-eclipse was finally defined as phases φ<0.45 and φ>0.57. -- 5 -- 0.6 0.6 0.4 0.4 0.2 0.2 0 0 -0.2 -0.2 -0.4 -0.4 -0.6 -0.6 0.4 0.4 0.2 0.2 0 0 -0.2 -0.2 0.4 0.4 0.45 0.45 0.5 0.5 phase phase 0.55 0.55 0.6 0.6 Fig. 1. -- Top: De-trended differential light curves. Open and filled dots show, respectively, the Feb 19 and Oct 18 UT 2009 data. Bottom: Combined light curves binned by a factor of 12. Blue squares correspond to WASP-12 and black dots to the differential light curve of the two comparison stars. The best fit models are shown as solid lines (for e = 0) and dashed lines (for e = 0.057). Both models produce the same depth and center phase, but the e = 0.057 model lasts 11.52 minutes longer. We attribute the flux jumps between phases 0.475 and 0.5 to unremoved systematics. Notice that, although the systematics appear in both curves, the trends in each curve are not correlated in phase. systematics. We chose not to use those points in the final analysis. Correlations with the other parameters listed above are not significant in any of the two datasets. Finally, we produce one light curve per night by combining the de-trended light curves of each comparison. The light curves are combined applying a weighted average based on the Poisson noise of the individual light curve points. The result is illustrated in Figure 1. The out-of-eclipse scatter of the combined light curves is 0.11% for the February 19 data and 0.09% for the October 18 data. De-trending significantly improves the systematics, but some unidentified residual noise sources remain, which we have not been able to fully model. -- 6 -- 2.1. Eclipse detection and error estimation The two-night combined light curve contains 421 points between phases 0.413 and 0.596, based on the Hebb et al. (2009) ephemerides. To establish the presence of the eclipse and its parameters, we fit the data to a grid of models generated using the JASMINE code, which combines the Kipping (2008) and Mandel & Agol (2002) algorithms to produce model light curves in the general case of eccentric orbits. The models do not include limb darkening and use as input parameters the orbital period, stellar and planetary radii, argument of the periastron, orbital inclination, stellar radial velocity amplitude and semi-major axis values derived by Hebb et al. (2009). The eccentricity is initially assumed to be e=0, which produces models with a total eclipse duration of 2.808 hours. The best fit model is found by χ2 minimization, with the depth, the central phase of the eclipse, and the out-of-eclipse differential flux as free parameters. First we fit the individual night light curves to ensure the eclipse signal is present in each dataset. The February 19 data give an eclipse depth of 0.100 ± 0.023%, while the derived eclipse depth for the October 18 data is 0.068 ± 0.021%. The central phases are φ=0.510 for the first eclipse and φ = 0.508 for the second. We assume the difference in depth is due to systematics we have not been able to properly model. The incomplete eclipse from February 19 might seem more prone to systematics, but our inspection of both datasets does not reveal stronger trends in that dataset. We therefore combined the data from both nights, weighting each light curve based on its out-of-eclipse scatter. The result of the combined light curve analysis is the detection of an eclipse with a depth of 0.082 ± 0.015% and centered at orbital phase φ = 0.51, as shown in Figure 1. The reduced χ2 of the fit is 0.952. The error in the eclipse depth is computed using the equation σ2 depth = σ2 r describes the red noise. The σr is estimated with the binning technique by Pont et al. (2006) to be 1.5×10−4 when binning on timescales up to the ingress and egress duration of about 20 minutes. r , where σw is the scatter per out-of-eclipse data point and σ2 w/N + σ2 We investigate to what extent the uncertainties in the system's parameters affect our eclipse depth and central phase results. Varying the impact parameter, planet-to-star ratio, and scale of the system by 1σ of the reported values in Hebb et al. (2009), the measured eclipse depth changes only by 0.004% or 0.27σdepth, while the central phase remains un- changed. Our result is therefore largely independent of the adopted system parameters. We performed three other tests to confirm the eclipse detection in a manner simi- lar to previously reported eclipse results (Deming et al. 2005; Sing & L´opez-Morales 2009; Rogers et al. 2009). From the average of the 125 out-of-eclipse light curve data points versus -- 7 -- the 228 in-eclipse points (only points where the planet is fully eclipsed, adopting φ=0.51 as the central eclipse phase), we measure an eclipse depth of 0.080 ± 0.015%. We further check the detection by producing histograms of the normalized light curve flux distribution in the in-eclipse and out-of-eclipse portions of the light curve. The result, illustrated in Figure 2, shows how the flux distribution of in-eclipse points is shifted by 0.00082 with respect to the out-of-eclipse flux distribution, centered at zero. The results of the last test, where we use a new set eclipse of models with a duration corresponding to e = 0.057, fix the out-of-eclipse baseline to zero, and leave both depth and central phase of the eclipse as a free parameters, are shown in Figure 3. The two top panels in the figure show 1-D cuts of each parameter through the z (normalized χ2), space, where z=0 gives the best fit model and z=1 defines the 1σ confidence interval of the result (see definition of z in §3 of L´opez-Morales & Bonanos 2008; Behr 2003). However, inspection of the eclipse depth in Figure 1 reveals that the 1σ errors from this method (±0.036%) are too large. To estimate more realistic errors for the central phase, we generate contours plots (bottom panel in Fig. 3), where the 1σ eclipse depth error has been fixed to the 0.015% value derived above. The resultant center phase is φ = 0.5100+0.0072 −0.0061. We also applied prayer-bead, bootstrapping, and Markov-Chain Monte Carlo (MCMC) error analysis techniques (Gillon et al. 2007; Southworth 2008) to estimate the errors in the central phase of the eclipse. The MCMC analysis contained 2.85×106 links, and we adopted the 1.5×10−4 photometric error (∼1.6 times the formal error), given by the Pont et al. (2006) binning technique. All these give central phases within φ = 0.5100+0.0030 −0.0036. The larger error in the central phase given by the normalized χ2 method suggests the presence of correlated noise in the data, which the other methods might not be correctly accounting for. We therefore adopt these larger errorbars in our final estimation of the central phase. 2.2. Eccentricity The eccentricity e of WASP-12b was calculated from the measured central phase shift value using eq. (6) from Wallenquist (1950), e cos ω = π P (t2 − t1 − P/2) 1 + csc2 i , (1) where P , i and ω are, respectively, the orbital period, inclination, and periastron angle of the system, and t2 − t1 is the time difference between transit and eclipse. In our case t2 − t1 = 0.51P . Using the values of P , i and ω from Hebb et al. (2009), we derive an e = 0.057+0.040 −0.034, which agrees with the non-zero eccentricity result reported by these authors. This eccentricity can be in principle explained if 1) the system is too young to have already circularized, 2) there are additional bodies in the system pumping the eccentricity of WASP- 12b, or 3) the tidal dissipation factor Q′ P (Goldreich 1963) of WASP-12b is several orders of -- 8 -- magnitude larger than Jupiter's, estimated to be between 6 ×104 and 2 ×106 (Yoder & Peale 1981). 3. Comparison with atmospheric models We compare the observed z ′-band flux of WASP-12b to simple blackbody models and to expectedly more realistic radiative-convective models of irradiated planetary atmospheres in chemical equilibrium, following the same procedure described in Rogers et al. (2009). The results are shown in Figures 4 and 5. In the simplistic blackbody approximation, a 0.082±0.015% deep eclipse corresponds to a z ′-band brightness temperature of Tz ′ ∼ 3028+99 −110K, slightly lower than the planet's equilibrium temperature of Tp ∼ 3129K assuming zero Bond albedo (AB = 0) and no energy reradiation (f = 2/3) (see L´opez-Morales & Seager 2007). However, when the thermal and reflected flux of the planet are included, different combinations of AB and f can yield the same eclipse depth, as illustrated in Figure 4. From that figure we can constrain the energy redistribution factor to f ≥ 0.585 ± 0.080, but the albedo is not well constrained. Assuming a maximum AB ≤ 0.4, the temperature of the day-side of WASP-12b is Tp > 2707K. The more realistic atmospheric models are derived from self-consistent coupled ra- diative transfer and chemical equilibrium calculations, based on the models described in Sudarsky et al. (2000, 2003), Hubeny et al. (2003) and Burrows et al. (2005, 2006, 2008) (see Rogers et al. 2009, for details). We generate models with and without thermal inversion layers, by adding an unidenfitied optical absorber between 0.43 and 1.0 µm, with different level of opacity κ′. The opacity of the absorber varies parabolically with frequency, with a peak value of κ′ = 0.25 cm2 gr−1. As Figure 5 shows, models with and without extra ab- sorbers produce similar fits to the observed z ′-band flux. The best model without absorber has a Pn = 0.32. The best model with an extra absorber has a Pn = 0.1 and κe = 0.1 cm2 gr−1. Observations at other wavelengths are necessary to further constrain the models. 2Pn = 0 and Pn = 0.5 correspond, respectively, to f = 2/3 and f = 1/4, however there is not well defined Pn − f relation for intermediate values since the physical models account for atmospheric parameters (e.g. pressure, opacity) in a way different than blackbody models. -- 9 -- 4. Discussion and Conclusions This first detection of the eclipse of WASP-12b agrees with the slight eccentricity of the planet's orbit found by Hebb et al. (2009), and places initial constraints to its atmospheric characteristics. The presence of other bodies in the system can be tested via radial velocity or transit timing variation observations, although the current RV curve by Hebb et al. (2009) shows no evidence of additional planets, unless they are in very long orbits. One would expect that if extra absorbers are present in the upper atmosphere of the planet in gaseous form, they might give rise to thermal inversion layers. However, as Figure 5 illustrates, the observed 0.9 µm eclipse depth can be fit equally well by a model without extra absorbers. Additional observations at longer wavelengths, specially longer than ∼ 4.0 µm, will break that model degeneracy. Observations at wavelengths below ∼ 0.6 µm will also place better constraints on AB. Note: While this paper was in the final revision stages, two new secondary eclipse observations from Spitzer were reported by Campo et al. (2010). They find secondary eclipse central phases consistent with a circular orbit. We have carefully reviewed our analysis and still arrive to a slightly eccentric orbit, although the result is only significant at the 1.4 -- 1.6σ confidence level. Possible explanations proposed by Campo et al. (2010) for the measured eccentricity difference are orbital precession or wavelength-dependent brightness variations across the surface of the planet that would shift the center of the eclipse. A similar effect has been recently suggested by Swain et al. (2010) on the surface of HD189733b. Further observations are needed to establish whether this discrepancy is a data artifact or a real effect. M.L.M. acknowledges support from NASA through Hubble Fellowship grant HF-01210.01- A/HF-51233.01 awarded by the STScI, which is operated by the AURA, Inc. for NASA, under contract NAS5-26555. J.L.C acknowledges support from a NSF Graduate Research Fellowship. A.B. is supported in part by NASA grant NNX07AG80G. D.A. and J.C.R. are grateful for support from STScI Director's Discretionary Research Fund D0101.90131. This work has been supported by NSF's grant AST-0908278. Behr, B. B. 2003, ApJS, 149, 67 REFERENCES -- 10 -- Barnes, R. et al. 2009, ArXiv e-prints Bodenheimer, P. et al. 2003, ApJ, 592, 555 Burrows, A., Budaj, J., & Hubeny, I. 2008, ApJ, 678, 1436 Burrows, A., Hubeny, I., & Sudarsky, D. 2005, ApJ, 625, L135 Burrows, A., Sudarsky, D., & Hubeny, I. 2006, ApJ, 650, 1140 Campo, C. J., Harrington, J., Hardy, R. A., Stevenson, K. B., Nymeyer, S., Ragozzine, D., Lust, N. B., Anderson, D. R., Collier-Cameron, A., Blecic, J., Britt, C. B. T., Bowman, W. C., Wheatley, P. J., Deming, D., Hebb, L., Hellier, C., Maxted, P. F. L., Pollaco, D., & West, R. G. 2010, ArXiv e-prints Deming, D. et al. 2005, Nature, 434, 740 D´esert, J. et al. 2008, A&A, 492, 585 Fortney, J. J. et al. 2008, ApJ, 678, 1419 Gillon, M. et al. 2007, A&A, 471, L51 Goldreich, R. 1963, MNRAS, 126, 257 Hebb, L. et al. 2009, ApJ, 693, 1920 Hubeny, I., Burrows, A., & Sudarsky, D. 2003, ApJ, 594, 1011 Ibgui, L. & Burrows, A. 2009, ApJ, 700, 1921 Ibgui, L., Burrows, A., & Spiegel, D. S. 2009, ArXiv e-prints Ibgui, L., Spiegel, D. S., & Burrows, A. 2009b, ArXiv e-prints Kipping, D. M. 2008, MNRAS, 389, 1383 L´opez-Morales, M. & Seager, S. 2007, ApJ, 667, L191 L´opez-Morales, M. & Bonanos, A. Z. 2008, ApJ, 685, L47 Mandel, K. & Agol, E. 2002, ApJ, 580, L171 Miller, N., Fortney, J. J., & Jackson, B. 2009, ApJ, 702, 1413 Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231 -- 11 -- Rogers, J. C. et al., 2009, ArXiv e-prints Sing, D. K. & L´opez-Morales, M. 2009, A&A, 493, L31 Southworth, J. 2008, MNRAS, 386, 1644 Spiegel, D. S., Silverio, K., & Burrows, A. 2009, ApJ, 699, 1487 Sudarsky, D., Burrows, A., & Hubeny, I. 2003, ApJ, 588, 1121 Sudarsky, D., Burrows, A., & Pinto, P. 2000, ApJ, 538, 885 Swain, M. R. et al., 2010, Nature, 463, 637 Wallenquist, A. 1950, Arkiv for Astronomi, 1, 59 Yoder, C. F. & Peale, S. J. 1981, Icarus, 47, 1 Zahnle, K. et al. 2009a, AAS Meeting Abstracts, Vol. 214, 306.01 Zahnle, K. et al. 2009b, ApJ, 701, L20 This preprint was prepared with the AAS LATEX macros v5.2. -- 12 -- Normalized Flux Histogram Normalized Flux Histogram 40 40 30 30 20 20 10 10 0 0 -0.004 -0.004 -0.002 -0.002 0 0 0.002 0.002 0.004 0.004 Differential Flux Differential Flux Fig. 2. -- Normalized flux histograms of the in-eclipse (dotted red line) and out-of-eclipse (solid line) portions of the WASP-12 light curve in Figure 1. The bin width is 0.00082 in differential flux, coincident with the detected eclipse depth. -- 13 -- ) 2 χ d e z i l a m r o N ( z 4 3 2 1 0 0.14 0.46 0.48 0.50 0.52 Center Phase 0.54 0.56 3σ 1σ 2σ 3σ ) 2 χ d e z i l a m r o N ( z 4 3 2 1 0 0.00 0.14 0.12 0.10 0.08 0.06 0.04 ) % ( h t p e d e s p i l c E 0.04 0.07 Eclipse Depth (%) 0.11 0.48 0.50 Center phase 0.52 0.54 Fig. 3. -- Top: Model eclipse 1 -- D cuts through the normalized -- χ2 parameter space for the eclipse depth and center phase. z=0 gives the best fit and z=1 shows the 1σ confidence interval. Bottom: Contour plot of the eclipse depth versus the center phase, where the 1σ depth error has been fixed to 0.015%. The best model fit is indicated by the cross symbol at φ=0.51 and depth = 0.082%, together with 1σ to 3σ confidence contours. The star symbol at φ=0.50 corresponds to a circular orbit. 0.5 0.4 0.3 B A o d e b l a d n o B 0.2 0.1 0.0 0.2 σ 1 σ 2 σ 1 σ 2 0.3 0.4 0.6 Heat redistribution factor (f) 0.5 0.7 0.8 Fig. 4. -- Values of AB and f that reproduce the observed z ′-band eclipse depth of WASP- 12b, assuming the planet emits as a blackbody. The shaded area highlights the region of allowed f values (1/4 − 2/3). The short and long dashed lines delimit, respectively, the 1σ and 2σ confidence regions of the result. -- 14 -- Fig. 5. -- Comparison of the eclipse depth (planet-to-star flux ratio) to models. The green line shows the AB = 0.4, f = 2/3 blackbody model from Figure 4. The blue and red lines show, respectively, the best fit model for an atmosphere with no extra absorber, and with an extra absorber of opacity κ′ between 0.43 and 1.0 µm. The black thin line at the bottom indicates the SPICam plus SDSS z ′-band filter response. See §3 for more details.
1605.03584
2
1605
2016-09-30T14:46:30
Robo-AO Kepler Planetary Candidate Survey III: Adaptive Optics Imaging of 1629 Kepler Exoplanet Candidate Host Stars
[ "astro-ph.EP" ]
The Robo-AO \textit{Kepler} Planetary Candidate Survey is observing every \textit{Kepler} planet candidate host star with laser adaptive optics imaging to search for blended nearby stars, which may be physically associated companions and/or responsible for transit false positives. We present in this paper the results of our search for stars nearby 1629 \textit{Kepler} planet candidate hosts. With survey sensitivity to objects as close as $\sim$0.15" and magnitude differences $\Delta$m$\le$6, we find 223 stars in the vicinity of 206 target KOIs; 209 of these nearby stars have not previously been imaged in high resolution. We measure an overall nearby-star probability for \textit{Kepler} planet candidates of 12.6\%$\pm$0.9\% at separations between 0.15" and 4.0". Particularly interesting KOI systems are discussed, including 23 stars with detected companions which host rocky, habitable zone candidates, and five new candidate planet-hosting quadruple star systems. We explore the broad correlations between planetary systems and stellar binarity using the combined dataset of Baranec et al. (2016) and this paper. Our previous 2$\sigma$ result of a low binary fraction of KOIs hosting close-in giant planets is less apparent in this larger dataset. We also find a significant correlation between binary fraction and KOI number, suggesting possible variation between early and late \textit{Kepler} data releases.
astro-ph.EP
astro-ph
Revision draft Preprint typeset using LATEX style emulateapj v. 01/23/15 ROBO-AO KEPLER PLANETARY CANDIDATE SURVEY III: ADAPTIVE OPTICS IMAGING OF 1629 KEPLER EXOPLANET CANDIDATE HOST STARS Carl Ziegler1, Nicholas M. Law1, Tim Morton2, Christoph Baranec3, Reed Riddle4, Dani Atkinson3, Anna Baker5, Sarah Roberts6, and David R. Ciardi7 Revision draft ABSTRACT The Robo-AO Kepler Planetary Candidate Survey is observing every Kepler planet candidate host star with laser adaptive optics imaging to search for blended nearby stars, which may be physically associated compan- ions and/or responsible for transit false positives. We present in this paper the results of our search for stars nearby 1629 Kepler planet candidate hosts. With survey sensitivity to objects as close as ∼0.(cid:48)(cid:48)15 and magnitude differences ∆m≤6, we find 223 stars in the vicinity of 206 target KOIs; 209 of these nearby stars have not previously been imaged in high resolution. We measure an overall nearby-star probability for Kepler planet candidates of 12.6%±0.9% at separations between 0.(cid:48)(cid:48)15 and 4.(cid:48)(cid:48)0. Particularly interesting KOI systems are discussed, including 26 stars with detected companions which host rocky, habitable zone candidates, and five new candidate planet-hosting quadruple star systems. We explore the broad correlations between planetary systems and stellar binarity using the combined dataset of Baranec et al. (2016) and this paper. Our previous 2σ result of a low detected nearby star fraction of KOIs hosting close-in giant planets is less apparent in this larger dataset. We also find a significant correlation between detected nearby star fraction and KOI number, suggesting possible variation between early and late Kepler data releases. Keywords: binaries: close - instrumentation: adaptive optics - techniques: high angular resolution - methods: data analysis - methods: observational - planets and satellites: detection - planets and satellites: fundamental parameters 1. INTRODUCTION The primary Kepler mission vastly increased the tally of known extrasolar planets, discovering over 2300 confirmed planets and approximately 4700 planet candidates (Borucki et al. 2010, 2011a,b; Batalha et al. 2013; Burke et al. 2014; Rowe et al. 2014; Coughlin et al. 2015; Morton et al. 2016). Using high-precision photometry to detect the periodic dip in stellar brightness consistent with a transiting planet, Ke- pler exoplanet candidates (Kepler Objects of Interests, or KOIs) require follow-up observations to rule out astrophys- ical false positives and for host star characterization (Brown et al. 2011). Most solar-type stars, which comprise the majority of Ke- pler targets (Batalha et al. 2013), form with at least one companion star (Duquennoy & Mayor 1991; Raghavan et al. 2010). The large effective point-spread function (6-10(cid:48)(cid:48)) and coarse resolution (pixel size of ∼4(cid:48)(cid:48)) (Haas et al. 2010) of Ke- pler allow these companion stars and background objects to be blended with the host candidate, illustrated in Figure 1. High-angular-resolution follow-up imaging is crucial to dis- tinguish these blended multiple stellar systems and identify false transit signals. Even when the candidates are bona fide [email protected] 1 Department of Physics and Astronomy, University of North Carolina at Chapel Hill, Chapel Hill, NC 27599-3255, USA 2 Department of Astrophysical Sciences, Princeton University, Prince- 3 Institute for Astronomy, University of Hawai'i at M¯anoa, Hilo, HI ton, NJ 08544, USA 96720-2700, USA 4 Division of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA 5 Durham Academy Upper School, 3601 Ridge Road, Durham, NC 27705, USA 6 Juniata College, 1700 Moore St, Huntingdon, PA 16652, USA 7 NASA Exoplanet Science Institute, California Institute of Technology, Pasadena, CA 91125, USA planets, the planet radius measurements based on the diluted transit signal are underestimated due to the presence of mul- tiple stars in the system or unbounded stars within the Kepler aperture (Ciardi et al. 2015). Before being elevated to planet candidate status, each KOI is vetted for clear signatures of being an astrophysical false positive: center-of-light shifts during transit, an identifiable secondary eclipse signal indicating the eclipsing object is self- luminous, or sharing the exact ephemeris as another KOI. While these vetting efforts on early catalogs were largely based on human inspection (Batalha et al. 2010), the most re- cent DR24 catalog has fully automated this process (Cough- lin et al. 2015). Notably, the candidate status of a KOI is not a function of its depth or shape (i.e., whether it is V- shaped), which means that a large fraction of the deeper signals (∼50%) can be expected to be false positives (San- terne et al. 2012, 2015). Shallower candidates have a much lower predicted false positive rate (∼10%) (Morton & John- son 2011; Fressin et al. 2013), a prediction that has been con- firmed by follow-up observations from the Spitzer space tele- scope (D´esert et al. 2015). Nevertheless, even if a large frac- tion of the candidate signals are real planets, many of the in- ferred properties of these planets are affected by the presence of blended sources (Dressing & Charbonneau 2013; Santerne et al. 2013). Therefore to fully characterize individual Kepler planets and to measure any possible biasing effects of stellar multiplicity on the planetary populations, every KOI needs to be searched for stellar companions8. There has been considerable effort by the community to per- form high-resolution imaging surveys of the KOIs (Howell et al. 2011; Adams et al. 2012, 2013; Lillo-Box et al. 2012, 8 For brevity we denote stars which we found within our detection radius of KOIs as "companions," in the sense that they are asterisms associated on the sky. 2 Ziegler et al. Figure 1. On the left, the full-frame Robo-AO reduced image of KOI-4418 (KIC2859893) rotated and scaled to match the Kepler view of the same field, displayed on the right, each pixel colored by the mean flux in Quarter 4. KICs in the field are marked on both images, as well as KP magnitude in the Kepler image. The 1.(cid:48)(cid:48)41 binary to KOI-4418 is not visible in the ∼4(cid:48)(cid:48) pixels of Kepler, illustrating how real companions and background stars can blend with the KOIs, resulting in astrophysical false positives or inaccurate planetary characteristics. High resolution follow-ups are a crucial step in the validation and characterization of Kepler planetary systems. 2014; Horch et al. 2012; Marcy et al. 2014; Dressing et al. 2014; Horch et al. 2014; Wang et al. 2015a,b; Torres et al. 2015; Everett et al. 2015; Kraus et al. 2016). These surveys, however, have combined covered approximately 30% of the full set of Kepler planetary candidates. This piecemeal ap- proach leads to inconsistent vetting, while limiting the com- prehensive statistics and correlations that can be derived from a large dataset of high resolution images of multiple stellar systems hosting planetary systems. In addition, target lists of past surveys are often biased towards brighter targets, possi- bly skewing any interpretations drawn from the data. Kepler planet candidates. We conclude in Section 6. 2. SURVEY TARGETS AND OBSERVATIONS 2.1. Target Selection A complete, consistent high-resolution survey of all the the KOIs with ground-based adaptive optics (AO) is limited by the typical overheads required with traditional systems. Taking advantage of the order-of-magnitude increase in time- efficiency provided by Robo-AO, the first robotic laser adap- tive optics system, we are performing high-resolution imag- ing of every KOI system. The first paper in this survey, Law et al. (2014, hereafter Paper I), observed 715 Kepler plane- tary candidates, identifying 53 companions, with 43 new dis- coveries, for a detected companion fraction of 7.4%±1.0% within separations of 0.(cid:48)(cid:48)15 to 2.(cid:48)(cid:48)5. The second paper in this survey, Baranec et al. (2016, hereafter Paper II), observed 969 Kepler planetary candidates, identifying 202 companions, with 139 new discoveries, for a detected companion frac- tion of 11.0%±1.1% within separations of 0.(cid:48)(cid:48)15 to 2.(cid:48)(cid:48)5., and 18.1%±1.3% within separations of 0.(cid:48)(cid:48)15 to 4.(cid:48)(cid:48)0. This paper presents a total of 1629 targets observed, around which we find 223 companions around 206 KOIs, 209 of which have not been previously imaged in high resolution, for a detected companion fraction of 12.6%±0.9% within 4.(cid:48)(cid:48)0 of planetary candidate hosting stars. We begin in Section 2 by describing our target selection, the Robo-AO system, and follow-up observations. In Section 3 we describe the Robo-AO data reduction and the companion detection and analysis. In Section 4 we describe the results of this survey, including discovered companions, and compare to other KOI surveys. We discuss the results in Section 5, detail- ing the effects on the planetary characteristics of the survey's discoveries and looking at the overall binarity statistics of the KOI targets were selected from the KOI catalog based on Q1-Q17 Kepler data (Borucki et al. 2010, 2011a,b; Batalha et al. 2013; Burke et al. 2014; Rowe et al. 2014; Coughlin et al. 2015). We selected targets not observed in Paper I and Paper II, with the objective of completing the Robo-AO sur- vey of all KOIs, including those with already detected com- panions. Observations in this paper are primarily from the 2014-15 observing seasons; residual observations of dim tar- gets from 2012-13 are also included, their analysis now pos- sible using our improved binary detection and characteriza- tion pipeline. KOIs flagged as false positives using Kepler data were removed. In Figure 2 the properties of the targeted KOIs in this work as well as for the full Robo-AO survey as of the end of the 2015 observing season are compared to the set of all KOIs from Q1-Q17 with CANDIDATE disposi- tions based on Kepler data. The Robo-AO target distribution closely matches the full KOI list in magnitude, planetary ra- dius, planetary orbital period, and stellar temperature. On-sky positions of all targeted KOIs in the complete survey are dis- played in Figure 3. 2.2. Observations 2.2.1. Robo-AO We obtained high-angular-resolution images of the 1629 KOIs during 55 separate nights of observations between 2012 July 16 and 2015 June 12 (UT), detailed in Table 9 in the Ap- pendix. The observations were performed using the Robo-AO laser adaptive optics system (Baranec et al. 2013, 2014b; Rid- dle et al. 2012) mounted on the Palomar 1.5-m telescope. The first robotic laser guide star adaptive optics system, the au- tomatic Robo-AO system can efficiently perform large, high angular resolution surveys. The AO system runs at a loop rate of 1.2 kHz to correct high-order wavefront aberrations, delivering a median Strehl ratio of 9% in the i(cid:48)-band. Obser- vations were taken in either a i(cid:48)-band filter or a long-pass filter cutting on at 600 nm (LP600 hereafter). The LP600 filter ap- Robo-AO Kepler Planetary Candidate Survey III 3 Figure 2. Comparison of the distribution of the Robo-AO sample in this paper as well as the combined Robo-AO survey (Paper I, Paper II, and this work) to the complete set of KOIs from Q1-Q17 (Borucki et al. 2010, 2011a,b; Batalha et al. 2013; Burke et al. 2014; Rowe et al. 2014; Coughlin et al. 2015). proximates the Kepler passband at redder wavelengths, while also suppressing blue wavelengths that reduce adaptive optics performance. Typical seeing at the Palomar Observatory is between 0.(cid:48)(cid:48)8 and 1.(cid:48)(cid:48)8, with median around 1.(cid:48)(cid:48)1 (Baranec et al. 2014b). The typical FWHM (diffraction limited) resolution of the Robo- AO system is 0.(cid:48)(cid:48)15. Images are recorded on an electron- multiplying CCD (EMCCD), allowing short frame rates for tip and tilt correction in software using a natural guide star (mV < 16) in the field of view. Specifications of the Robo-AO KOI survey are summarized in Table 1. ion detections are ongoing and will appear in future papers in this survey. 2.2.3. Gemini LGS-AO 2.2.2. Keck LGS-AO Eight candidate multiple systems were selected for re- imaging by the NIRC2 camera behind the Keck-II laser guide star adaptive optics system (Wizinowich et al. 2006; van Dam et al. 2006), on 2015 July 25 (UT) to confirm possible com- panions. The targets were selected for their low signifi- cance of detectability, either because of low contrast ratio or small angular separation. Observations were performed in the Kprime filter using the narrow mode of NIRC2 (9.952 mas pixel−1; Yelda et al. 2010), dithering the primary target at in- tervals of 30 s into the 3 lowest noise quadrants, for a total ex- posure time of 90 s. The images were corrected for geometric distortion using the NIRC2 distortion solution of Yelda et al. (2010). Targets observed with Keck are detailed in Table 2. Further follow-up observations of low-significance compan- Seven candidate multiple systems from this work and three from Paper I and Paper II, again selected for their low de- tection significance, were re-imaged with the adaptive optics assisted NIRI instrument (Hodapp et al. 2003) on the Gemini North telescope. Three targets were observed on 2015 July 31 (UT) and seven targets were observed on 2015 August 27, us- ing Band 3 allocated time. Targets observed with Gemini are detailed in Table 3. Observations were performed with the F/32 camera, providing resolution of 21.9 mas pixel−1 across a field of view of 22(cid:48)(cid:48) × 22(cid:48)(cid:48). Total integration times were 90 s in the Kprime band across three dithered images, used to increase dynamic range and allow sky subtraction. The com- mon striping pattern found in NIRI images was removed using the cleanir.py script provided by the Gemini staff. The images were flat fielded, bad pixel corrected, and sky subtracted. The distortion solution provided by the Gemini staff was used to correct the images for distortion. 3. DATA REDUCTION With the largest adaptive optics dataset yet assembled by Robo-AO, the data reduction process was automated as much as possible for efficiency and consistency. As in Paper I and Ziegler et al. (2015), after initial pipeline reductions de- scribed in Section 3.1, the target stars were identified (Section 4 Ziegler et al. Table 1 The specifications of the Robo-AO KOI survey KOI targets FWHM resolution Observation wavelengths Field size Detector format Pixel scale Exposure time Targets observed / hour Observation dates 1629 ∼0.(cid:48)(cid:48)15 (@600-750 nm) 600-950 nm 44(cid:48)(cid:48)× 44(cid:48)(cid:48) 10242 pixels 43.1 mas / pix 90 seconds 20 2012 July 16 -- 2015 June 12 Figure 3. Location on sky of targeted KOIs from Paper I (L14), Paper II (B16), and this work (TW). The median coordinates of the targeted KOIs is designated by an '×'. A projection of the Kepler field of view is provided for reference. 3.2), companions automatically identified (Section 3.5), PSF subtraction performed and companions again auto-identified (Section 3.4), and constraints of the companion sensitivity of the survey measured (Section 3.6). Finally, the properties of the detected companions are measured in Section 3.7. 3.1. Imaging Pipeline The Robo-AO imaging pipeline (Law et al. 2009, 2014) re- duced the images: the raw EMCCD output frames are dark- subtracted and flat-fielded and then stacked and aligned using the Drizzle algorithm (Fruchter & Hook 2002), which also up-samples the images by a factor of two. To avoid tip/tilt anisoplanatasism effects, the image motion was corrected by using the KOI itself as the guide star in each observation. 3.2. Target Verification To verify that the star viewed in the image is the desired KOI target, we created Digital Sky Survey cutouts of similar angular size around the target coordinates. Each image was manually checked to assure no ambiguity in the target star with images with either poor performance or incorrect fields removed. These bad images made up approximately 2% of all our images, and for all but 2 of the targets additional images were available. Figure 4. Example of PSF subtraction on KOI-5762 with companion sep- aration of 0.(cid:48)(cid:48)34. The yellow circle marks the position of the primary star's PSF peak. Both images have been scaled and smoothed for clarity. Success- ful removal of the PSF leaves residuals consistent with photon noise. The 2(cid:48)(cid:48)square field shown here is approximately equal to half the Kepler pixel size. The close companion to KOI-5762 was confirmed with NIRC2/Keck images, shown in Figure 7. To facilitate the automation of the data reduction, centered 8.(cid:48)(cid:48)5 square cutouts were created around the 1629 verified tar- get KOIs. We select a 4(cid:48)(cid:48)separation cutoff for our companion search to detect all nearby stars that would blend with the tar- get KOI in a Kepler pixel. 3.4. PSF Subtraction To identify close companions, a custom locally optimized point spread function (PSF) subtraction routine based on the Locally Optimized Combination of Images algorithm (Lafreni`ere et al. 2007) was applied to centered cutouts of all stars. Detailed in Paper I, the code uses a set of twenty KOI observations, selected from the observations within the same filter closest to the target observation in time, as refer- ence PSFs, as it is improbable that a companion would ap- pear at the same position in two different images. A locally optimized PSF is generated and subtracted from the original image, leaving residuals consistent with photon noise. This procedure was performed on all KOI images out to 2(cid:48)(cid:48), and the results visually checked for companions. Figure 4 shows an example of the PSF subtraction performance. The PSF subtracted images were subsequently run through the au- tomated companion finding routine, as described in Section 3.5. 3.3. Image Preparation 3.5. Companion Detection Robo-AO Kepler Planetary Candidate Survey III 5 An initial visual companion search was performed redun- dantly by three of the authors. This search yielded a prelimi- nary companion list, and filtered out bad images. Continuing the companion search, we ran all images through a custom automated search algorithm, based on the code described in Paper I. The algorithm slides a 5-pixel di- ameter aperture within concentric annuli centered on the tar- get star. Any aperture with >+5σ outlier to the local noise is considered a potential astrophysical source. These are sub- sequently checked manually, eliminating spurious detections with dissimilar PSFs to the target star and those having char- acteristics of a cosmic ray hit, such as a single bright pixel or bright streak. The detection significance of 'secure' compan- ions are listed in Tables 4 and 5. Many possible companions were visually identified but fell beneath the formal 5σ required for a discovery. Despite not reaching our formal significance level required for a discov- ery, previous results suggest that all but a small fraction are likely real: Keck/NIRC2 observations have confirmed all 15 'likely' detections in Paper I and all 38 re-observed 'likely' companions in Paper II. The detection significance of these 'likely' companions are listed in Tables 6 and 7. 3.6. Imaging Performance Metrics The two dominant factors that affect the image performance of the Robo-AO system are seeing and target brightness. An automated routine was used to classify the image performance for each target. Described in detail in Paper I, the code uses PSF core size as a proxy for image performance. Observa- tions were binned into three performance groups, with 31% fall in the low performance group, 41% in the medium perfor- mance group, and 28% in the high performance group. We determine the angular separation and contrast consis- tent with a 5σ detection by injecting artificial companions, a clone of the primary PSF. For concentric annuli of 0.(cid:48)(cid:48)1 width, the detection limit is calculated by steadily dimming the artifi- cial companion until the auto-companion detection algorithm (Section 3.5) fails to detect it. This process is subsequently performed at multiple random azimuths within each annulus and the limiting 5σ magnitudes are averaged. For clarity, these average magnitudes for all radii measurements are fit- ted with functions of the form a ∗ sinh(b ∗ r + c) + d (where r is the radius from the target star and a, b, c and d are fitting variables). Contrast curves for the three performance groups are shown in Section 4 in Figure 5. 3.7. Companion Characterization 3.7.1. Contrast Ratios For wide, resolved companions with little PSF overlap, the companion to primary star contrast ratio was determined us- ing aperture photometry on the original images. The aperture radius was cycled in one-pixel increments from 1-5 FWHM for each system, with background measured opposite the pri- mary from the companion (except in the few cases where an- other object falls near or within this region in the image). Pho- tometric uncertainties are estimated from the standard devia- tion of the contrast ratios measured for the various aperture sizes. For close companions, the estimated PSF was used to re- move the blended contributions of each star before aperture photometry was performed. The locally optimized PSF sub- traction algorithm can attempt to remove the flux from com- panions using other reference PSFs with excess brightness in those areas. For detection purposes, we use many PSF core sizes for optimization, and the algorithm's ability to re- move the companion light is reduced. However, the compan- ion is artificially faint as some flux has still been subtracted. To avoid this, the PSF fit was redone excluding a six-pixel- diameter region around the detected companion. The large PSF regions allow the excess light from the primary star to be removed, while not reducing the brightness of the companion. 3.7.2. Separation and Position Angles Separation and position angles were determined from the raw pixel positions. Uncertainties were found using estimated systematic errors due to blending between components. Typ- ical uncertainty in the position for each star was 1-2 pixels. Position angles and the plate scale were calculated using a distortion solution produced using Robo-AO measurements for the globular cluster M15.9 3.7.3. Companion Spectral Types The approximate spectral type of the detected companions, assuming that they are bound to the primary and all stars are main sequence dwarfs, were estimated using an SED model (Kraus & Hillenbrand 2007) and the estimated stellar effective temperatures reported on the NASA Exoplanet Archive. With the LP600 band closely matching the Kepler bandpass, the magnitude differences between the primary star and nearby stars were converted to i(cid:48)-band when necessary using the lin- ear correlation found by Lillo-Box et al. (2014): ∆mi(cid:48) = 0.947 · ∆mLP600 (1) In addition, we estimate the latest spectral type companion consistent with a 5σ detection for each observed target based on the typical contrast curve for the three image performance groups (see Section 3.6). These estimates are listed in Table 9 in the Appendix. We caution that these spectral types are approximate and do not account for factors such as giant contamination, estimated at ∼12% by Ciardi et al. (2011). In addition, the use of a lin- ear correlation in converting between passbands will result in an error with varying spectral types. We estimate this error by calculating the flux of F0V to M5V stars (Pickles 1998) in the LP600 and i(cid:48)-band inlcuding quantum efficiencies and instru- mental effects (see Figure 2 in Paper I). The maximum differ- ence between the flux of spectral types in the two passbands results in an error of ∼0.15 mags, equivalent to approximately one subspectral type. 4. DISCOVERIES Of the 1629 KOI targets observed, 206 are apparent in multiple star systems for a nearby star fraction within 4(cid:48)(cid:48)of 12.6%±0.9%10. We also found 15 triple systems for a triplet fraction of 0.92+0.30−0.18%11, and 1 quadruple system for a quadru- plet fraction of 0.06+0.14−0.02%11. Cutouts of the triple and quadru- ple systems are shown in Figure 6. One quarter (25.8%) of the companions would only be observable in a high-resolution survey (<1.(cid:48)(cid:48)0 separation), and one half (49.8%) of the com- panions are too close (<2.(cid:48)(cid:48)0) for many seeing limited surveys to accurately measure binary properties (e.g. contrast ratios). The detected companion separations and contrast ratios are 9 S. Hildebrandt (2013, private communication) 10 Error based on Poissonian statistics (Burgasser et al. 2003) 11 Error based on binomial statistics (Burgasser et al. 2003) 6 Ziegler et al. Figure 5. Separations and magnitude differences of the detected companions in Paper II and this work, with the color and shape of each star denoting the associated typical low-, medium- and high-performance 5σ contrast curve during the observation (as described in Section 3.6). plotted in Figure 5, along with the calculated 5σ detection limits as detailed in Section 3.6. Cutouts of all multiple star systems are shown in Figures 18, 19, 20, and 21. For compan- ions within 2.(cid:48)(cid:48)5, measured properties are detailed in Tables 4 and 6. For companions outside 2.(cid:48)(cid:48)5 but within 4.(cid:48)(cid:48)0, measured properties are detailed in Tables 5 and 7. Full Keck-AO Observation List Companion? Table 2 ObsID KOI 1447 1873 2117 2554 mv 13.2 15.8 16.2 15.9 2015 Jul 25 2015 Jul 25 2015 Jul 25 2015 Jul 25 We confirmed six companions to eight Robo-AO detections with NIRC2 and AO on Keck-II (Wizinowich et al. 2000). In addition, two new companions were found around KOIs 2554 and 3020. These targets were selected for followup because of their faintness and/or closely separated detected compan- ion. Low-sigma, visually detected companions to KOIs 1873 and 5257 were not detected. These non-detected companions are possibly a result of non-common path aberrations, as de- scribed in Section 5.1 of Paper II. These spurious detections all have similar separations and position angles with respect to the target star, facilitating their identification and manual removal. The PSF subtraction routine usually does not re- move these false companions, as another star exhibiting the NCP error is unlikely to be within the set of twenty reference images. The Keck-II observations are listed in Table 2, with the measured separations and position angles of the confirmed companions using Keck-II images listed in Tables 4 and 6, and the follow-up images are shown in Figure 7. We confirmed five companions to seven KOIs observed in this paper with NIRI and AO on Gemini North. We did not detect a possible companion to KOI-2198 that was visually detected, manifesting as an elongated PSF in the Robo-AO image. We observed three KOIs targeted in Paper I (KOI- 327) and Paper II (KOIs 2833 and 4301) which displayed non-common path error aberrations. No companions were ob- served to these three targets in the follow-up observations. A new companion outside our separation cutoff (ρ=4.(cid:48)(cid:48)24) was observed nearby KOI-4131. The Gemini observations are listed in Table 3, and the follow-up images are shown in Figure 8. ∆Kp 0.63±0.06 0.53±0.06 0.27±0.05a 2.96±0.10b 1.27±0.06c 5.01±0.07d 1.22±0.13 0.83±0.08 yes yes yes yes yes yes yes yes 3020 13.8 2015 Jul 25 3106 5257 5762 15.7 15.5 15.9 2015 Jul 25 2015 Jul 25 2015 Jul 25 aCompanion at ρ=0.(cid:48)(cid:48)37 bNew companion at ρ=3.(cid:48)(cid:48)55 cCompanion at ρ=0.(cid:48)(cid:48)38 dNew companion at ρ=3.(cid:48)(cid:48)86 4.1. Comparison to Other Surveys Two detected companions (KOI-326 and KOI-841) in our survey were previously found in Lillo-Box et al. (2012), who observed 98 KOIs using the AstraLux Lucky Imaging sys- tem on the 2.2m telescope at the Calar Alto Observatory. Lillo-Box et al. (2014) also previously detected companions to KOI-3263, 3649, and 3886 in a survey of 174 KOIs. Adams et al. (2012) and Adams et al. (2013) observed 87 and 13 KOIs, respectively, with the instruments ARIES and PHARO on the MMT and Palomar telescopes, respectively. They de- tect companions to KOI-126 and 266 that are fainter than our survey sensitivity. Observing 87 KOIs with ARIES at the MMT, Dressing et al. (2014) previously detected companions to KOI-2813 and KOI-3111 and also detected a companion to KOI-266 (∆mKs =6.32) that is outside our detection sensitiv- ity. Gilliland et al. (2015) found two companions to KOI-829 using the Hubble Space Telescope (HST) with ∆mKp of 2.4 Robo-AO Kepler Planetary Candidate Survey III 7 Figure 6. Normalized log-scale cutouts of 16 KOIs with multiple companions with separations <4(cid:48)(cid:48) resolved with Robo-AO. The angular scale and orientation (displayed in the first frame) is similar for each cutout, and circles are centered on the detected nearby stars. Three targets (KOIs 3214, 3463, and 6800) have a possible third companion, marked with arrows, outside our 4(cid:48)(cid:48) separation cutoff, as described in Section 5.2.1. and 6.0 and separations of 0.(cid:48)(cid:48)11 and 3.(cid:48)(cid:48)31, respectively, which were outside the detection limits of our Robo-AO image. Wang et al. (2015a) observed 84 KOIs using the PHARO and NIRC2 instruments at Palomar and Keck, respectively, with one discovered companion (KOI-3678) appearing in our sur- vey. Two of our targets (KOI-1411 and KOI-3823) have com- panions detected by Wang et al., both with ∆mK >5, which fall outside our detection sensitivity. Wang et al. (2015b) ob- served 73 multiple transiting planet KOI systems at Palomar and Keck, with the only overlapping system being a compan- ion observed near KOI-1806 which we did not detect. The companion to KOI-1806, measured by Wang et al. (2015b) as ∆K=1.45 at 3.(cid:48)(cid:48)43 separation, is well within out survey sen- sitivity, and the reason for the non-detection is unclear. The reported companion is also not visible in UKIRT images, al- Kraus et al. (2016) observed 382 KOIs with AO on the Keck-II telescope. They detected single companions to KOI- 255, 1908, 2705, and 2813, and both companions to KOI- 1201 that were detected in our survey. They also detected single companions to KOI-1298, 1681, 2179 2453, and 2862, and double companions to KOI-1361 and 2813, that all fall outside of our reported sensitivity. Kolbl et al. (2015) searched for the blended spectra of KOIs with secondary stars within ∼0.(cid:48)(cid:48)8 using Keck-HIRES opti- though it would be detectable. We detected companions to KOI-126 and 200 not detected by Howell et al. (2011); both companions are within the stated sensitivity limits for their re- spective targets, so the reason for the earlier non-detection is unclear. None of our nearby-star detections overlap with the discoveries of Everett et al. (2015). 8 Ziegler et al. Figure 7. Normalized log-scale cutouts of 8 KOIs observed with the NIRC2 instrument on Keck-II, as described in Section 2.2.2. The angular scale and orientation (displayed in the first frame) is similar for each cutout, and circles are centered on the detected nearby stars. Table 3 Full Gemini Observation List ObsID Companion? ∆Kp 2015 Aug 27 2015 Aug 27 2015 Aug 27 2015 Jul 31 2015 Aug 27 2015 Jul 31 2015 Aug 27 2015 Jul 31 2015 Aug 27 2015 Aug 27 yes yes yes yes yes yes 4.41±0.09a 4.96±0.11b 0.75±0.04 0.53±0.05c 4.11±0.09d 1.54±0.04 KOI 327 2198 2833 4131 4301 5052 5164 5243 5497 5774 mv 13.1 12.8 12.8 13.2 13.3 12.8 12.6 12.5 11.0 11.1 aCompanion at ρ=2.(cid:48)(cid:48)85 bNew companion at ρ=4.(cid:48)(cid:48)24 cCompanion at ρ=0.(cid:48)(cid:48)77 dCompanion at ρ=2.(cid:48)(cid:48)41 cal echelle spectra of 1160 California Kepler Survey KOIs. Of the 63 KOIs the authors found with evidence of a sec- ondary star, we found companions to seven (KOIs 1137, 2813, 3161, 3415, 3471, 4345, 4713) and did not detect companions to eight (KOIs 1121, 1326, 1645, 3515, 3527, 3605, 3606, 3853). The companions we did not detect likely lie at small separations inside the limits of our survey sensitivity. Two of our companions (KOIs 1137 and 3415) fall within their cal- culated flux ratio uncertainty and within their ∼0.(cid:48)(cid:48)8 separation limit. Without known separations and position angles, how- ever, it is not clear that these are the same companion stars. Nine of the widest nearby stars we detected have 2MASS (Skrutskie et al. 2006) designations. 102 of our wide (ρ >2(cid:48)(cid:48)) nearby star detections are noted on the Kepler Community Follow-up Observing Program (CFOP) using J-band, ∼1(cid:48)(cid:48) seeing-limited imaging from United Kingdom InfraRed Tele- scope (UKIRT) (Lawrence et al. 2007). However, with high- acuity imaging to resolve blended companions, providing greater precision photometry, and a filter that better simulates the Kepler bandpass, the Robo-AO survey can better evaluate the effect of the companion on the observed transit signal. 4.2. Multiplicity and Other Surveys There have been multiple past high-resolution surveys of KOIs performed, allowing our results to be put into context with the overall community follow-up program. A compar- ison of the observed nearby-star rates from various surveys with differing methodologies may also provide convergence on the intrinsic multiplicity rate of planet hosting stars. With varying sensitivities between surveys, we use a lower sepa- ration cutoff than in this paper for a uniform comparison be- tween surveys. This also has the added benefit of using only the nearest stars that have the highest probability of associa- tion. An exact comparison between surveys is still hindered, however, by the use of dissimilar instruments, passbands, and target selection criteria; in comparing results in this section we attempt to highlight major differences when comparing multiplicity rates, however we caution that in each case there are inherent biases in the coverage of the different surveys which requires detailed analysis not covered in this work. We find that 6.8% of KOIs have nearby stars within 2(cid:48)(cid:48), in agreement with other visible light surveys: 6.4% in Paper I, 8.2% in Paper II, and 6.4% (Howell et al. 2011). Horch et al. (2014) found 7.0% of KOI targets had nearby stars within 1(cid:48)(cid:48) separation, a range where we showed a 3.4% nearby star rate. Horch et al. (2014) do not report their target list, so it is not possible to identify the source of this discrepancy. It is pos- sible that this is a result of our target selection of every KOI, resulting in a dimmer overall sample than surveys which pri- oritize brighter targets. The targets in this paper have a me- dian Kp,med = 14.9, significantly fainter overall than the targets in Adams et al. (2012, Kp,med = 12.2), Dressing et al. (2014, Kp,med = 13.3), Wang et al. (2015a, Kp < 14), Paper I (Kp,med = 13.7), and in Paper II (Kp,med = 14.2). Horch et al. (2014) note that their Kepler targets mainly are between 11th and 14th magnitude. There are several reasons a brighter overall target list will inflate binarity rates: the target stars are intrin- sically more luminous, which results in more physically as- sociated companion stars as binarity correlates with luminos- ity (Duchene & Kraus 2013); the target stars are less distant, so physically associated companion of a given spectral type is brighter, thus easier to detect; brighter stars tend to have Robo-AO Kepler Planetary Candidate Survey III 9 Secure Detections of Objects within 2.(cid:48)(cid:48)5 of Kepler Planet Candidates Table 4 b ObsID ((cid:48)(cid:48)) UKIRT UKIRT UKIRT 0.02 0.01 0.01,0.02 W15 CFOP D14, K16 CFOP 0.01 0.01,0.02 21±2 85±2 UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT 9.5 2014 Aug 20 12±2 89±2 23±2 34±2 Filter Det. Significance Separation P.A. Mag. Diff. Approx. Comp. Prev. High Res. Prev. Low Res. False Positive?a NKOI (deg.) Spectral Typec Detection? Detection? σ 17 6.6 5.6 18 6.1 27 5.1 12 6.3 9.6 7.2 8.7 6.9 5.9 5.4 27 13 10 66 5.2 6.1 14 8.1 10 54 10 11 8.1 6.5 11 13 9.2 5.3 15 39 124 5.8 6.4 6.0 11 7.2 7.7 7.4 106 61 19 5.6 8.7 6.7 9.5 6.2 5.9 14 5.8 224 28 6.0 12 (mag) 1.22±0.06 214±2 -- 0.36±0.03 1.49±0.06 204±2 2.08±0.04 2.45±0.06 348±2 2.53±0.05 1.77±0.06 281±2 0.94±0.05 0.75±0.06 197±2 0.81±0.08 2.17±0.06 312±2 2.58±0.02 0.28±0.03d 212±2d 0.27±0.08 1.77±0.06 188±2 0.91±0.02 2.11±0.06 209±2 4.10±0.16 1.99±0.06 111±2 0.98±0.05 1.30±0.06 215±2 1.72±0.04 2.08±0.06 352±2 3.10±0.02 0.77±0.06 248±2 0.04±0.03 1.73±0.06 2.47±0.02 0.37±0.03d 149±2d 0.37±0.08 1.10±0.06 258±2 0.84±0.02 0.96±0.06 272±2 0.38±0.02 2.36±0.06 1.30±0.04 2.02±0.06 198±2 2.66±0.06 0.38±0.03d 272±2d 0.93±0.22 1.87±0.06 147±2 1.62±0.02 1.87±0.06 151±2 0.49±0.03 1.14±0.06 278±2 0.87±0.03 1.41±0.06 198±2 2.50±0.04 2.18±0.06 3.79±0.03 0.74±0.06 0.03±0.05 1.51±0.06 2.15±0.03 0.79±0.06 216±2 0.26±0.03 0.60±0.06 160±2 1.05±0.12 1.20±0.06 1.44±0.04 0.50±0.06 116±2 1.13±0.09 0.89±0.06 138±2 1.13±0.05 1.41±0.06 172±2 2.23±0.02 1.03±0.06 325±2 0.04±0.02 1.72±0.06 251±2 0.27±0.04 2.09±0.06 322±2 1.95±0.02 2.28±0.02 2.27±0.06 1.05±0.06 109±2 1.05±0.02 0.75±0.06 285±2 0.68±0.06 0.77±0.06 0.55±0.03 2.41±0.06 128±2 5.53±0.08 2.06±0.06 236±2 4.64±0.06 0.33±0.06 1.78±0.22 2.11±0.06 3.24±0.03 2.17±0.06 225±2 1.79±0.05 1.32±0.06 336±2 1.90±0.05 0.77±0.06 246±2 1.42±0.11 2.14±0.06 4.40±0.05 1.23±0.06 0.90±0.03 2.17±0.06 241±2 4.14±0.14 1.22±0.06 279±2 1.43±0.06 1.75±0.06 1.17±0.04 1.31±0.06 0.50±0.02 0.52±0.06 271±2 0.58±0.07 2.21±0.06 353±2 1.60±0.02 2.20±0.06 5.38±0.07 1.04±0.06 0.44±0.03 0.87±0.06 260±2 1.40±0.09 KOI mi(cid:48) (mag) 1 13.3 2012 Jul 18 LP600 163 1 14.5 2014 Jul 16 LP600 454 4 14.3 2014 Jul 14 LP600 510 1 771 15.1 2014 Aug 27 LP600 1 1137 13.8 2014 Jun 13 LP600 1 1409 15.0 2014 Jul 17 LP600 2 1447 13.0 2012 Sep 04 LP600 1 1630 14.9 2014 Jul 16 LP600 1 1687 14.9 2014 Jul 17 LP600 3 1792 11.9 2014 Sep 02 LP600 1 2091 15.5 2014 Aug 27 LP600 3 2093 15.2 2014 Aug 27 LP600 3 2163 14.4 2014 Aug 31 LP600 1 2535 14.6 2014 Aug 23 LP600 2 2554 15.0 2014 Sep 01 LP600 1 2813 13.3 2013 Aug 15 LP600 2 2896 11.9 2015 Jun 05 LP600 1 2900 15.0 2014 Sep 03 LP600 1 2976 15.6 2014 Aug 28 LP600 1 3020 13.5 2013 Aug 13 LP600 1 3042 15.8 2014 Aug 31 LP600 1 3112 15.6 2014 Sep 01 LP600 1 3120 14.6 2014 Aug 29 LP600 2 3214 11.8 2014 Aug 29 LP600 1 3413 15.0 2014 Aug 26 LP600 1 3415 13.1 2013 Jul 27 LP600 1 3483 14.7 2014 Nov 09 LP600 1 3649 15.2 2014 Aug 23 LP600 1 3660 15.3 2014 Aug 24 LP600 1 3770 13.9 2014 Jun 19 LP600 1 3886 1 4343 13.5 2014 Jun 19 LP600 1 4418 15.7 2014 Sep 03 LP600 1 4550 15.0 2014 Aug 29 LP600 1 4713 13.4 2014 Jul 16 LP600 1 4750 15.7 2014 Aug 29 LP600 2 4895 14.5 2014 Aug 31 LP600 1 5004 14.3 2014 Jul 16 LP600 1 5052 12.5 2014 Jun 17 LP600 1 5243 12.2 2014 Sep 03 LP600 1 5243 12.2 2014 Sep 03 LP600 1 5570 14.5 2014 Aug 21 LP600 1 5578 10.9 2014 Nov 09 LP600 1 5665 11.3 2014 Jul 17 LP600 1 5671 13.4 2014 Jun 16 LP600 1 5774 10.7 2014 Sep 01 LP600 1 5889 15.2 2014 Sep 01 LP600 1 6111 12.9 2015 Jun 04 LP600 3 6132 14.6 2015 Jun 12 LP600 1 6258 11.2 2015 Jun 04 LP600 1 6329 14.0 2015 Jun 04 LP600 1 6415 14.0 2015 Jun 03 LP600 1 6475 13.7 2015 Jun 07 LP600 1 6482 13.6 2015 Jun 04 LP600 1 6527 12.3 2015 Jun 07 LP600 1 6560 12.9 2015 Jun 06 LP600 1 7205 14.1 2015 Jun 04 LP600 7448 11.3 2015 Jun 12 LP600 1 Notes. -- References for previous detections are denoted using the following codes: Dressing et al. 2014 (D14), Lillo-Box et al. 2014 (LB14), Kraus et al. 2016 (K16), Wang et al. 2015a (W15), visible in United Kingdom InfraRed Telescope images (UKIRT), high angular resolution images available on Kepler Community FollowUp Observing Program (CFOP). aProbability that planetary transit signal is a false positive based on Kepler data. bNumber of planet candidates detected orbiting KOI. cEstimation method described in Section 3.7.3. dFrom Keck follow-up, described in Section 4. deeper detectable contrast ratios. K1 M0 M1 K5 K5 M0 F6 K3 M5 K4 M0 M0 G2 M2 M0 K7 F8 M0 M2 G9 K3 K3 G8 M0 M2 G9 K5 F9 K4 K1 M0 M4 M0 K4 G7 M0 M0 K3 F6 G4 M4 M4 K7 M1 K4 G8 K1 M1 G8 M2 K2 K5 M2 G7 G7 M5 G8 G0 48±2 91±2 48±2 57±2 30±2 42±2 75±2 17±2 89±2 91±2 UKIRT UKIRT UKIRT UKIRT 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 LB14 LB14 UKIRT UKIRT CFOP CFOP CFOP CFOP 0.01 0.01 The disparity in multiplicity between papers in the Robo- AO survey was explored in Section 6 of Paper II as a pos- sible result of the bias in the KOI selection process between data releases, with the median observed KOI in Paper II lo- cated nearer the Galactic plane than in Paper I. KOIs near the Galactic plane lie in denser stellar fields, increasing the likeli- hood of unassociated nearby stars with the separation cutoff. Plotting the Kepler field of view with our targeted KOIs in Figure 3, the median position of KOIs in this work is closer to the center of the field than in Paper II, and further from the Galactic plane than Paper I or Paper II. Surveys in the NIR find higher multiplicity rates within 2(cid:48)(cid:48): 13% (Dressing et al. 2014), 17% (Adams et al. 2012), 20% (Adams et al. 2013). This is likely caused by many compan- ions being cool, red dwarf stars that are faint in the optical (Ngo et al. 2015), and deeper, higher angular resolution imag- ing. i(cid:48) UKIRT 10 Ziegler et al. Secure Detections of Objects outside 2.(cid:48)(cid:48)5 and within 4.(cid:48)(cid:48)0 of Kepler Planet Candidates Table 5 i(cid:48) ObsID ((cid:48)(cid:48)) (deg.) Spectral Typec Detection? σ 5.8 6.3 6.2 7.6 5.2 7.1 5.7 7.9 21 6.1 6.6 7.9 8.8 29 31 15 19 14 9.0 Filter Det. Significance Separation P.A. Mag. Diff. Approx. Comp. Prev. High Res. KOI mi(cid:48) (mag) 14.5 255 2014 Jul 17 LP600 15.1 2014 Sep 02 LP600 734 1558 15.0 2014 Jul 11 LP600 1593 15.6 2014 Aug 24 LP600 1846 15.5 2014 Sep 02 LP600 2213 15.1 2014 Aug 24 LP600 2014 Jul 17 LP600 2744 14.9 3791 13.6 2014 Aug 22 3928 13.1 2014 Jul 14 LP600 4343 13.5 2014 Jun 19 LP600 4630 14.7 2014 Jul 17 LP600 4743 14.7 2014 Sep 03 LP600 4993 12.5 2014 Sep 01 LP600 5220 11.8 2014 Sep 03 LP600 5327 15.0 2014 Sep 01 LP600 5332 14.3 2015 Jun 12 LP600 5465 13.7 2014 Jun 19 LP600 0.01 7020 13.5 2015 Jun 12 LP600 7395 11.7 2015 Jun 12 LP600 0.01 Notes. -- References for previous detections are denoted using the following codes: Kraus et al. 2016 (K16), visible in United Kingdom InfraRed Telescope images (UKIRT), high angular resolution images available on Kepler Community FollowUp Observing Program (CFOP), companions visible in UKIRT and with 2MASS designations (J*). aProbability that planetary transit signal is a false positive based on Kepler data. bNumber of planet candidates detected orbiting KOI. cEstimation method described in Section 3.7.3. (mag) 3.41±0.06 357±2 2.14±0.04 3.51±0.06 175±2 2.05±0.04 3.61±0.06 308±2 1.09±0.04 3.24±0.06 80±2 1.60±0.03 3.77±0.06 136±2 1.07±0.03 3.94±0.06 91±2 1.67±0.02 3.50±0.06 257±2 2.12±0.03 3.50±0.06 258±2 1.89±0.04 2.96±0.06 265±2 1.21±0.03 3.68±0.06 350±2 4.81±0.15 3.94±0.06 53±2 2.17±0.05 3.06±0.06 98±2 2.29±0.04 3.49±0.06 148±2 4.13±0.02 2.89±0.06 216±2 3.27±0.05 3.96±0.06 342±2 -- 0.12±0.03 3.61±0.06 129±2 0.63±0.03 2.85±0.06 158±2 1.36±0.05 3.28±0.06 23±2 1.43±0.04 3.41±0.06 212±2 3.00±0.04 J19401085+4658310 J19192894+4643440 J19411432+4302399 M4 K7 K0 K4 K7 M0 M0 K3 G6 M3 M0 M0 M2 M3 M1 G7 K3 G9 G8 J19261347+4212546 J19405741+4219181 UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT 0.01 0.01 J19422364+4335492 K16 CFOP Prev. Low Res. Detection? False Positive?a NKOI b 2 2 1 2 1 1 2 2 1 1 1 1 1 1 1 1 1 1 1 Figure 8. Normalized log-scale cutouts of 10 KOIs observed with the NIRI instrument on Gemini North, as described in Section 2.2.3. The angular scale and orientation (displayed in the first frame) is similar for each cutout, and circles are centered on the detected nearby stars. 5. DISCUSSION In this section, we delve further into the combined datasets of Paper II and this work to further explore the implications of stellar multiplicity on the planetary candidates (Section 5.1), expand on the planetary candidates found in higher order mul- tiple systems or orbiting within the habitable zone (Section 5.2), and search for insight into the role that multiple stel- lar bodies play on planetary formation and evolution (Section 5.3). 5.1. Implications for Kepler Planet Candidates When a close companion is detected near a KOI host star, there are several potential implications. If the planet does in- deed transit the purported target star, the consequences may √ be relatively mild: the planet's radius will be slightly larger than had previously been thought -- at most by a factor of 2 in the case of an equal-brightness companion (Ciardi et al. 2015). If the eclipsed star is a faint companion, however, the radius of the eclipsing object may be many times larger, po- tentially turning a small planet into a giant planet or a planet into a false-positive eclipsing binary star. Additionally, as the properties of most of the host stars in the Kepler stellar cat- alog are based on broad-band photometry assuming that they are single, the derived stellar radii may well be incorrect if the system actually contains multiple stars. Re-fitting the stellar properties of all the companion stars, as well as for the Kepler target stars accounting for the presence of the companions, is beyond the scope of this work, but will be addressed in a future paper in this series. Finally, if a KOI system has multiple transiting planets de- tected, it might be the case that the planets are distributed around multiple stars in the system. KOI-284/Kepler-132 is a good example of such a case (Fabrycky et al. 2012; Lissauer et al. 2014): its multiple planets would be unstable if they all ObsID Robo-AO Kepler Planetary Candidate Survey III 11 Likely Detections of Objects within 2.(cid:48)(cid:48)5 of Kepler Planet Candidates Table 6 ((cid:48)(cid:48)) Spectral Typec P.A. Mag. Diff. Approx. Comp. Prev. High Res. Prev. Low Res. False Positive?a NKOI (deg.) 36±2 44±2 69±2 Detection? Detection? 0.01,0.02 LB12 0.01 b 31±2 UKIRT UKIRT i(cid:48) i(cid:48) i(cid:48) CFOP K16 UKIRT UKIRT 0.02 0.02 0.01 14±2 23±2 21±2 25±2 61±2 43±2 20±2 16±2 36±2 68±2 KOI mi(cid:48) (mag) 13.1 2015 Jun 08 LP600 14.2 2014 Sep 01 LP600 14.6 2014 Jul 16 LP600 14.5 2014 Jul 17 LP600 15.8 2012 Sep 02 LP600 σ 3.3 3.0 3.5 2.8 2.6 2.9 3.1 2.8 4.5 2.7 2.8 3.7 4.2 3.7 2.9 4.9 3.5 3.1 4.9 3.3 3.2 4.2 3.6 3.7 2.6 3.1 3.6 3.0 3.4 3.5 2.3 3.2 3.0 4.5 3.2 3.0 3.3 4.3 3.9 2.8 2.9 3.6 3.1 4.3 4.9 3.4 3.9 3.8 2.7 3.4 3.0 4.8 2.7 2.6 3.0 4.7 2.9 3.1 3.8 4.0 4.8 4.1 3.0 3.1 3.2 3.6 2.9 3.3 3.0 3.7 4.6 2.9 3.1 3.9 3.2 3.1 4.8 4.0 3.9 3.0 3.3 3.2 Filter Det. Significance Separation (mag) 0.97±0.15 0.34±0.06 0.52±0.23 0.30±0.06 0.53±0.06 338±2 0.93±0.15 0.97±0.06 232±2 3.44±0.25 2.00±0.06 3.60±0.04 1.83±0.06 340±2 1.58±0.05 0.77±0.06 107±2 1.52±0.16 1.15±0.06 3.14±0.16 1.06±0.06 189±2 1.65±0.09 1.40±0.06 2.00±0.08 0.31±0.06 215±2 0.61±0.26 0.53±0.06 284±2 1.06±0.16 1.29±0.06 260±2 4.11±0.13 0.79±0.06 1.69±0.19 1.84±0.06 353±2 3.33±0.17 0.33±0.03d 111±2d 0.71±0.17 1.05±0.06 1.46±0.10 0.40±0.06 213±2 0.46±0.12 2.10±0.06 3.21±0.03 2.36±0.06 192±2 3.41±0.02 0.31±0.06 212±2 0.59±0.28 0.60±0.06 154±2 0.86±0.13 1.09±0.06 205±2 0.86±0.04 0.45±0.06 142±2 0.84±0.16 0.35±0.06 222±2 0.72±0.25 0.39±0.06 223±2 0.45±0.08 2.31±0.06 287±2 3.44±0.03 0.68±0.06 0.17±0.05 0.33±0.06 0.27±0.09 2.65±0.04 1.39±0.06 1.15±0.06 302±2 2.47±0.14 1.14±0.06 1.94±0.07 0.30±0.03d 189±2d 0.76±0.16 1.83±0.06 238±2 2.91±0.04 0.49±0.06 320±2 0.73±0.13 0.80±0.06 276±2 2.01±0.16 2.40±0.06 2.89±0.04 2.36±0.06 127±2 1.95±0.02 1.13±0.06 1.29±0.10 0.66±0.06 113±2 1.37±0.17 0.63±0.06 224±2 3.05±0.12 0.40±0.06 210±2 0.75±0.20 2.30±0.06 267±2 2.77±0.05 1.96±0.06 310±2 3.82±0.14 1.13±0.06 272±2 3.53±0.15 1.49±0.06 3.66±0.13 1.66±0.06 194±2 3.29±0.08 1.12±0.06 2.22±0.10 2.46±0.06 303±2 3.38±0.02 2.45±0.06 322±2 4.62±0.02 0.75±0.06 149±2 1.99±0.14 0.87±0.06 340±2 0.38±0.03 0.77±0.06 324±2 2.02±0.24 0.67±0.06 2.12±0.28 2.36±0.06 146±2 3.16±0.03 0.78±0.06 123±2 1.46±0.10 1.23±0.06 242±2 3.33±0.13 1.24±0.06 3.33±0.19 1.22±0.06 222±2 3.83±0.18 1.75±0.06 200±2 4.67±0.19 1.88±0.06 211±2 3.43±0.05 2.19±0.06 2.37±0.04 1.24±0.06 217±2 2.66±0.17 0.21±0.06 0.12±0.03 2.45±0.06 345±2 3.04±0.02 0.62±0.06 270±2 1.44±0.18 0.34±0.06 333±2 0.73±0.28 0.97±0.06 346±2 2.52±0.18 0.60±0.06 163±2 1.47±0.20 0.23±0.03d 95±2d 0.65±0.25 0.60±0.06 322±2 1.30±0.17 0.77±0.06 322±2 2.49±0.27 1.75±0.06 290±2 0.83±0.10 0.75±0.06 122±2 1.72±0.15 1.41±0.06 272±2 2.78±0.14 1.58±0.06 175±2 3.89±0.17 0.77±0.06 322±2 0.54±0.05 1.73±0.06 2.68±0.04 1.41±0.06 195±2 2.88±0.08 1.04±0.06 339±2 1.44±0.10 1.94±0.06 134±2 5.04±0.11 2.45±0.06 212±2 2.37±0.02 126 200 225 532 841 1261 14.9 2014 Aug 22 1503 14.6 2014 Aug 22 1506 14.8 2014 Sep 02 LP600 1656 14.8 2014 Jun 13 LP600 1660 15.4 2014 Aug 28 LP600 1695 13.6 2014 Aug 31 LP600 1792 11.9 2014 Sep 02 LP600 1908 14.2 2014 Aug 22 1973 15.3 2014 Aug 28 LP600 2048 15.5 2014 Aug 28 LP600 2117 15.2 2014 Nov 09 LP600 2283 14.7 2014 Sep 01 LP600 2376 15.0 2014 Aug 21 LP600 2445 15.6 2014 Aug 28 LP600 2460 14.6 2014 Aug 29 LP600 2482 14.8 2014 Aug 24 LP600 2580 15.5 2014 Aug 31 LP600 2688 16.1 2014 Aug 31 LP600 2760 14.5 2014 Aug 23 LP600 2797 15.6 2014 Aug 28 LP600 2851 15.2 2014 Aug 26 LP600 2856 15.1 2014 Aug 26 LP600 2862 15.3 2014 Aug 27 LP600 2926 15.7 2014 Aug 28 LP600 2927 15.7 2014 Aug 28 LP600 2958 14.6 2014 Sep 02 LP600 3043 14.6 2014 Jul 12 LP600 3106 15.2 2014 Aug 26 LP600 3136 15.4 2014 Aug 28 LP600 3214 11.8 2014 Aug 29 LP600 3263 15.3 2014 Aug 23 LP600 3335 15.6 2014 Sep 01 LP600 3372 15.2 2014 Aug 23 LP600 3418 15.2 2014 Aug 23 LP600 2014 Jul 16 LP600 3432 14.8 3471 13.0 2014 Jul 11 LP600 3480 15.7 2014 Sep 03 LP600 3611 16.3 2014 Aug 26 LP600 3626 16.2 2014 Sep 03 LP600 3783 12.8 2014 Aug 21 LP600 4062 13.9 2014 Aug 29 LP600 4267 15.0 2014 Jun 19 LP600 4323 13.4 2014 Jun 13 LP600 4366 15.3 2014 Aug 28 LP600 4421 12.6 2014 Jul 12 LP600 4549 15.7 2014 Aug 27 LP600 4590 15.5 2014 Sep 02 LP600 2014 Jul 19 LP600 4653 13.4 4759 14.8 2014 Jul 19 LP600 4810 15.0 2014 Aug 24 LP600 4923 13.0 2014 Jul 14 LP600 4974 15.5 2014 Aug 26 LP600 5101 12.9 2014 Jul 17 LP600 5143 15.7 2014 Nov 09 LP600 5232 13.5 2014 Aug 31 LP600 5327 15.0 2014 Sep 01 LP600 5332 14.3 2015 Jun 12 LP600 5340 15.0 2014 Jun 19 LP600 5373 11.5 2015 Jun 05 LP600 5440 15.1 2014 Aug 28 LP600 5482 15.0 2014 Aug 31 LP600 5486 12.6 2015 Jun 12 LP600 5553 15.3 2014 Aug 23 LP600 5695 14.9 2015 Jun 12 LP600 5762 15.4 2014 Sep 03 LP600 6109 11.9 2015 Jun 07 LP600 6202 11.4 2014 Aug 23 6311 2015 Jun 04 LP600 6464 13.7 2015 Jun 04 LP600 6483 12.5 2015 Jun 05 LP600 6539 12.5 2015 Jun 12 LP600 6602 10.2 2015 Jun 03 LP600 6610 15.3 2015 Jun 12 LP600 6654 13.5 2015 Jun 12 LP600 6706 13.8 2015 Jun 04 LP600 6728 13.9 2015 Jun 12 LP600 7426 15.4 2015 Jun 05 LP600 Notes. -- References for previous detections are denoted using the following codes: Lillo-Box et al. 2012 (LB12), Lillo-Box et al. 2014 (LB14), Kraus et al. 2016 (K16), visible in United Kingdom InfraRed Telescope images (UKIRT). aProbability that planetary transit signal is a false positive based on Kepler data. bNumber of planet candidates detected orbiting KOI. cEstimation method described in Section 3.7.3. dFrom Keck follow-up, described in Section 4. K1 G7 G8 M1 M3 K3 M0 M1 K4 K7 G7 K4 M5 M2 M4 M0 M2 K4 M2 M4 G9 K4 M0 M0 G7 K2 M1 M2 M1 M0 K7 K5 G9 M3 G8 M5 M0 K5 K2 M0 M3 K4 M0 M2 K5 M0 M0 K5 M3 M3 K7 K3 K3 K7 M1 K2 M2 M0 M3 M2 M5 K7 M0 K3 M0 K2 F6 M1 K2 K3 G6 M2 F3 K0 M0 K7 K4 K3 K7 G6 M4 M2 28±2 96±2 0.01 0.01 0.01 0.02 0.01 0.01 0.01 0.01 0.01 0.01 7±2 81±2 UKIRT UKIRT 0.01 0.01 UKIRT UKIRT 4±2 99±2 0.01 0.01 CFOP LB14 CFOP CFOP UKIRT UKIRT 9.0 i(cid:48) 84±2 UKIRT UKIRT CFOP 0.01,0.02 0.01,0.02,0.03 2 1 1 1 5 2 1 1 1 1 1 3 2 1 2 1 1 1 1 1 1 1 1 1 1 2 1 1 4 1 1 2 1 1 2 1 1 1 1 1 1 1 1 1 1 1 1 2 1 2 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2 1 1 3 1 1 1 1 1 1 1 1 UKIRT 12 Ziegler et al. Likely Detections of Objects outside 2.(cid:48)(cid:48)5 and within 4.(cid:48)(cid:48)0 of Kepler Planet Candidates Table 7 7±2 i(cid:48) LB12 J19244737+4718244 D14 CFOP W15 K16 K16 CFOP ((cid:48)(cid:48)) (deg.) 0.02 0.02 0.01 0.01 0.01 0.01 0.01 Spectral Typec Detection? Prev. Low Res. Detection? ObsID Filter Det. Significance Separation P.A. Mag. Diff. Approx. Comp. Prev. High Res. UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT σ 2.6 4.8 3.8 4.2 4.9 4.5 4.3 3.7 2.7 3.6 3.5 2.9 2.6 4.0 3.2 3.1 4.6 4.6 4.8 2.7 4.3 3.8 2.7 3.1 3.2 4.2 4.2 4.3 3.2 4.1 2.8 4.7 3.0 2.6 3.9 2.8 3.4 4.0 4.4 3.1 2.9 4.1 3.8 3.5 4.9 3.8 3.2 3.0 3.0 3.0 3.6 3.6 3.4 3.3 2.8 3.3 4.8 4.2 3.5 2.7 2.5 2.7 2.9 3.1 KOI mi(cid:48) (mag) 13.4 2013 Jul 25 LP600 51 14.7 2014 Aug 21 LP600 193 14.2 2014 Sep 01 LP600 200 14.8 2014 Aug 21 LP600 240 13.0 2013 Aug 15 LP600 326 14.5 2014 Aug 29 LP600 541 14.5 2014 Jul 11 LP600 598 15.5 2014 Sep 03 LP600 757 15.3 2014 Aug 24 LP600 814 15.4 2014 Aug 24 LP600 816 1193 15.0 2014 Aug 26 LP600 1201 14.9 2012 Aug 04 LP600 1201 14.9 2012 Aug 04 LP600 1441 14.9 2014 Aug 31 LP600 1804 15.3 2014 Sep 02 LP600 1995 15.0 2014 Aug 24 LP600 2050 12.2 2015 Jun 07 LP600 2014 Jul 19 LP600 2206 15.0 2379 14.9 2014 Aug 29 LP600 2579 15.0 2014 Jul 12 LP600 3066 15.6 2014 Aug 24 LP600 3111 12.7 2014 Aug 20 2015 Jun 03 LP600 9.6 3161 3264 15.6 2014 Aug 28 LP600 2014 Jul 17 LP600 3341 14.7 3347 15.2 2014 Aug 28 LP600 3354 14.9 2014 Jul 16 LP600 3463 14.6 2015 Jun 07 LP600 3463 14.6 2015 Jun 07 LP600 3533 14.4 2014 Nov 09 LP600 3678 12.6 2014 Jun 17 LP600 4131 13.2 2014 Jun 19 LP600 4268 14.8 2014 Aug 31 LP600 4334 15.5 2014 Sep 01 LP600 2014 Jul 13 LP600 4345 13.2 4353 15.4 2014 Aug 24 LP600 4405 14.5 2014 Jul 17 LP600 4467 15.6 2014 Aug 26 LP600 4526 15.1 2014 Aug 24 LP600 4526 15.1 2014 Aug 24 LP600 4655 15.2 2014 Aug 23 LP600 4661 14.5 2014 Jul 18 LP600 4700 15.7 2014 Aug 31 LP600 4710 15.4 2014 Sep 01 LP600 4881 12.7 2014 Aug 21 LP600 5210 14.9 2014 Jul 14 LP600 5216 15.3 2014 Aug 31 LP600 5220 11.8 2014 Sep 03 LP600 5327 15.0 2014 Sep 01 LP600 5331 14.9 2014 Aug 31 LP600 5480 16.3 2014 Aug 29 LP600 5556 13.2 2014 Jun 13 LP600 5556 13.2 2014 Jun 13 LP600 5707 15.0 2014 Aug 23 LP600 5885 14.7 2014 Aug 21 LP600 6120 15.4 2015 Jun 08 LP600 6560 12.9 2015 Jun 06 LP600 6605 11.3 2015 Jun 08 LP600 6610 15.3 2015 Jun 12 LP600 6745 15.2 2015 Jun 12 LP600 6745 15.2 2015 Jun 12 LP600 6800 12.8 2015 Jun 12 LP600 6800 12.8 2015 Jun 12 LP600 6925 15.7 2015 Jun 03 LP600 Notes. -- References for previous detections are denoted using the following codes: Lillo-Box et al. 2012 (LB12), Dressing et al. 2014 (D14), Kraus et al. 2016 (K16), visible in United Kingdom InfraRed Telescope images (UKIRT), companions visible in UKIRT and with 2MASS designations (J*). aProbability that planetary transit signal is a false positive based on Kepler data. bNumber of planet candidates detected orbiting KOI. cEstimation method described in Section 3.7.3. (mag) 3.51±0.06 161±2 2.63±0.07 2.78±0.06 137±2 3.07±0.02 2.81±0.06 130±2 4.00±0.02 2.71±0.06 272±2 3.46±0.03 3.53±0.06 267±2 2.01±0.02 2.80±0.06 246±2 3.50±0.02 3.17±0.06 357±2 2.73±0.04 2.94±0.06 243±2 3.37±0.04 3.40±0.06 346±2 4.16±0.07 3.50±0.06 120±2 2.66±0.03 3.08±0.06 2.81±0.02 2.81±0.06 236±2 4.26±0.08 3.76±0.06 265±2 5.17±0.12 3.06±0.06 333±2 3.73±0.03 2.88±0.06 168±2 2.84±0.03 2.96±0.06 355±2 5.34±0.04 3.33±0.06 215±2 5.33±0.04 3.28±0.06 87±2 1.28±0.05 3.59±0.06 139±2 1.89±0.03 3.48±0.06 355±2 3.69±0.03 3.41±0.06 335±2 1.86±0.02 3.36±0.06 234±2 5.87±0.13 2.68±0.06 67±2 3.04±0.14 3.66±0.06 217±2 1.37±0.02 3.23±0.06 107±2 4.27±0.08 3.24±0.06 295±2 2.20±0.02 3.71±0.06 227±2 2.55±0.06 3.67±0.06 96±2 4.41±0.04 2.74±0.06 79±2 4.79±0.02 3.08±0.06 10±2 5.21±0.03 2.63±0.06 170±2 5.08±0.04 2.85±0.06 124±2 5.04±0.02 3.56±0.06 263±2 4.77±0.04 3.32±0.06 15±2 3.79±0.06 3.17±0.06 242±2 3.22±0.02 3.50±0.06 36±2 2.75±0.04 2.95±0.06 249±2 3.19±0.02 3.99±0.06 131±2 4.21±0.04 2.53±0.06 346±2 4.44±0.02 3.98±0.06 179±2 4.80±0.09 3.17±0.06 116±2 3.02±0.05 3.93±0.06 198±2 2.32±0.05 3.77±0.06 49±2 1.89±0.05 2.70±0.06 168±2 3.50±0.05 3.42±0.06 30±2 3.30±0.03 2.71±0.06 267±2 2.22±0.05 3.67±0.06 96±2 3.31±0.03 2.83±0.06 109±2 7.22±0.08 3.63±0.06 277±2 3.92±0.02 3.67±0.06 351±2 3.72±0.03 3.52±0.06 174±2 1.24±0.05 3.28±0.06 162±2 4.31±0.03 3.22±0.06 247±2 5.29±0.02 2.71±0.06 239±2 2.43±0.02 3.42±0.06 127±2 4.03±0.04 3.85±0.06 128±2 2.48±0.02 3.28±0.06 246±2 6.04±0.12 2.53±0.06 320±2 3.46±0.04 2.63±0.06 216±2 1.22±0.02 3.07±0.06 72±2 3.78±0.02 2.85±0.06 163±2 3.92±0.03 2.62±0.06 145±2 5.10±0.04 3.11±0.06 337±2 5.41±0.04 2.66±0.06 125±2 1.71±0.12 M0 M0 M2 M1 M2 M3 M2 M3 M4 M0 M1 M6 M7 M2 M3 M4 M5 K4 K3 M2 M0 M5 K5 M0 M3 M0 M0 M3 M4 M3 M5 K7 M6 M4 M2 M0 M0 M4 M3 M4 M0 M2 M0 M2 M0 M1 M1 M7 M5 M4 K1 M4 M5 K7 M3 M0 M6 M0 G3 M0 M0 M4 M5 M4 UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT UKIRT 0.01 0.01 0.01 0.01 0.01 0.01 0.01 0.01 CFOP CFOP CFOP J19295122+4117529 0.01,0.02 0.01,0.02 False Positive?a NKOI b 0.01 1 1 1 1 2 1 2 3 1 1 1 1 1 1 1 1 2 1 1 3 1 2 1 1 2 1 1 1 1 1 1 2 1 1 1 1 1 1 2 2 1 1 1 1 2 1 1 1 1 1 1 1 1 1 1 2 1 1 1 1 1 1 1 1 J19214830+3951405 Robo-AO Kepler Planetary Candidate Survey III 13 additional stars in the full 44(cid:48)(cid:48) square image, including the target and possible companions. The probability based on the background star density that all three stars are bound is ap- proximately 98%. Assuming the planets orbit the brightest star, the two close stars likely dilute the observed transit sig- nal, leading to updated planetary radii estimates of 3.3 R⊕ and 2.6 R⊕ for the planet candidates on orbits with periods of 11.5 and 25.1 days, respectively. KOI-3463 hosts a 1.3 R⊕ planetary candidate on a 32.5 day orbit. We detect two nearby stellar companions, with angular separations of 2.(cid:48)(cid:48)74 and 3.(cid:48)(cid:48)67 and magnitude differences of 4.79 and 4.41, respectively. Just outside our 4(cid:48)(cid:48) separation cut- off, another 2.44 mag dimmer star appears at 4.(cid:48)(cid:48)11. KOI-3463 lies in a relatively dense stellar field, with at least 16 stars in the full 44(cid:48)(cid:48) square image, including the target and possible companions. The probability based on the background star density that all three stars are bound is approximately 86%. If the planet candidate orbits the bright star, the additional two nearby stars in the photometric aperture only marginally di- lute the transit signal, leading to an updated planetary radius estimate of 1.3 R⊕. KOI-6800 hosts a 27.5 R⊕ planetary candidate on a 2.5 day orbit. We detect two nearby stellar companions, with angular separations of 2.(cid:48)(cid:48)62 and 3.(cid:48)(cid:48)11 and magnitude differences of 5.10 and 5.41, respectively. Outside our 4(cid:48)(cid:48) separation cutoff, another 5.27 mag dimmer object appears at 4.(cid:48)(cid:48)13, although UKIRT photometry suggests that this is highly likely (>99%) a background galaxy. In the full 44(cid:48)(cid:48) square image of KOI- 6800, 9 stars are visible, including the target and possible companions. The probability based on the background star density that all three stars are bound is approximately 97%. The two dim nearby stars within the photometric aperture only slightly increases the estimated planetary radius to 27.7 R⊕, assuming the planet orbits the brightest star. If the planet candidate orbits one of the fainter stars, the corrected plane- tary radius would be large enough to make it highly probable the transiting event is in fact a background eclipsing binary. orbited a single star, but it turns out to be a close visual bi- nary, with the only sensible interpretation being that some of the planets transit one star and some transit the other. While such "split multiple" systems are predicted to be relatively rare among the population of Kepler multiple stellar systems (Fabrycky et al. 2012), any multi-planet system with a close companion has a higher chance of being split, and thus de- serves close consideration. Barclay et al. (2015) presents a model of how such systems might be analyzed, investigating the KOI-1422/Kepler-296. 5.2. Particularly Interesting Systems Several KOIs with detected companions are of particular interest, either for displaying unusual system characteristics, rare false-positive scenarios, or planetary attributes that sat- isfy habitability requirements. 5.2.1. Possible Quadruple Systems KOI-5327 hosts a 2.24 R⊕ planetary candidate on a 5.4 day orbit. We detect two nearby stellar companions, with angu- lar separations of 1.(cid:48)(cid:48)88 and 3.(cid:48)(cid:48)63 and magnitude differences of 3.43 and 3.92, respectively. A possible fourth component of the system lies at 3.(cid:48)(cid:48)96 and is 0.12 mags brighter than the KOI target. Further multiple passband observations and radial velocity measurements are needed to understand the hierar- chy of this system. In the full 44(cid:48)(cid:48) square image, a total of 8 stars appear, including the target and possible companions. With nearly equal brightness, the third companion has a high probability of being associated. With few stars found in the full field, it is unlikely that any unassociated stars would be found within 4(cid:48)(cid:48)of the KOI. The likelihood that the other two stars are in fact bound is 97%. If physical association is con- firmed for all four components, KOI-5327 would be the third known planet residing in a quadruple star system (Schwamb et al. 2013; Roberts et al. 2015). Following the analysis of Section 5.1 in Paper I, with the planet assumed to orbit the bright target star, the updated planetary radius estimate for the planetary candidate with all three stars in the aperture is 3.3 R⊕. Without the 0.12 mags brighter star in the photomet- ric aperture, the updated radius estimate is 2.3 R⊕. The second scenario detailed in Paper I, with the planet orbiting one of the fainter companions, will be further explored in future papers in this survey for all detected companions. KOI-4495, first detected in Paper II, has three nearby stars, with angular separations of 3.(cid:48)(cid:48)04, 3.(cid:48)(cid:48)06, and 3.(cid:48)(cid:48)41 and mag- nitude differences of 4.73, 3.90, and 2.68, respectively. The system hosts a planetary candidate with period of 5.92 days and estimated radius of 1.49 R⊕. The system lies in a rela- tively dense stellar field, thus it is probable that at least one of the stars is an unassociated asterism. The Robo-AO discov- ery image is available in Figure 5 of Paper II. With all three stars in the photometric aperture diluting the transit signal, and assuming the planet does indeed orbit the bright star, the updated planetary radius estimate is 1.6 R⊕. KOI-3214 hosts planetary candidates with radii of 2.59 R⊕ and 2.02 R⊕ on 11.5 and 25.1 day orbits, respectively. We de- tect two nearby stellar companions, with angular separations of 0.(cid:48)(cid:48)49 and 1.(cid:48)(cid:48)41 and magnitude differences of 0.73 and 2.50, respectively. Outside our 4(cid:48)(cid:48) separation cutoff, another 5.33 mag dimmer star appears at and 4.(cid:48)(cid:48)34. With multiple stars in the same Kepler pixel, the probability of an eclipsing bi- nary resulting in a false planetary transit signal is increased. KOI-3214 lies in a relatively sparse stellar field, with only 6 The Robo-AO images of the four possible quadruple sys- tems from this work are displayed in Figure 6. 5.2.2. Habitable Zone Candidates The discovery of habitable exoplanets is a major goal of the Kepler mission, and an accurate knowledge of the host star's properties is required to establish unambiguously whether an exoplanet possesses the two habitability conditions -- rocky and in a location where water can be found in a liquid state on the surface (habitable zone; HZ). The exact requirements for habitability are still debated (Kasting et al. 1993; Selsis et al. 2007; Seager 2013; Zsom et al. 2013), however it has been shown that the transition between "rocky" and "non-rocky" occurs rather sharply at RP=1.6R⊕ (Rogers 2015). For this analysis, we will use a large cutoff of 4 R⊕, as the presence of a stellar companion may dramatically alter the estimated radius, and even a gaseous planet in the HZ may host a rocky exomoon (Heller 2012). Overall, the existence of an unknown stellar companion within the same photometric aperture as the KOI will increase the calculated radius of the planet, as the observed transit signal will be diluted by the constant light of the nearby star. In Paper II and this work, we detected companions to 26 KOIs which host planetary candidates with equilibrium tem- peratures, from the NEA, in the HZ range (273 K≤Teq ≤373 K) and R<4 R⊕, displayed in Table 8. All are newly de- tected in this survey. Corrected planetary radii estimates are 14 Ziegler et al. Habitable Zone Candidates with Robo-AO Detected Companions Table 8 Planet candidate 227.01c 255.01 438.02c 1503.01 1846.01 1989.01c 2174.02c 2744.01 2760.01 2862.01 2926.03 2926.04 3255.01c 3284.01c 3401.02c 3946.01c 4550.01 4810.01 5101.01 5553.01 5671.01 5707.01 5885.01 6120.02 6745.01 6745.01 Period Rp,i (R⊕) 2.45 2.51 1.76 3.79 3.81 1.84 1.88 2.46 2.19 1.79 2.43 2.09 1.37 0.98 2.20 2.36 1.73 2.07 1.64 2.59 1.73 2.88 1.87 1.67 2.78 2.78 (d) 17.7 27.5 52.7 150.2 106.0 201.1 33.1 109.6 56.6 24.6 21.0 37.6 66.7 35.2 326.7 308.5 140.3 115.2 436.2 120.9 190.9 208.8 111.1 205.4 383.9 383.9 a Rp,c b (R⊕) 2.96 2.67 1.81 4.23 4.46 1.88 2.53 2.63 2.64 2.44 3.24 2.79 1.38 1.03 2.64 2.37 2.42 2.13 1.68 2.71 1.89 3.03 1.89 1.75 2.82 2.82 Equil. Temp. (K) 350 313 271 291 322 297 343 340 317 321 357 294 294 272 283 298 257 353 331 333 356 347 364 323 314 314 Sep ((cid:48)(cid:48)) 0.33 3.36 3.28 0.76 3.7 1.12 0.92 3.44 0.44 0.67 0.33 0.33 3.15 3.94 0.65 4.27 1.03 2.32 1.22 0.95 2.13 2.67 3.36 3.78 3.02 2.81 ∆m (mag) 0.84 2.14 3.11 1.52 1.07 3.49 0.21 2.12 0.84 0.17 0.27 0.27 4.87 2.42 0.89 5.26 0.04 3.16 3.33 2.52 1.79 2.43 4.03 2.48 3.78 3.92 aInitial planetary radius estimate bCorrected planetary radius estimate cDetected in Paper II of this survey. included, as described in Section 5.1 of Paper I, with the as- sumption that the planet orbits the bright star. KOI-2926 hosts two planetary candidates within the HZ, and KOI-6745 is a possible triple system hosting a planet in the HZ. The equilibrium temperature calculation is based on an estimate of the stellar effective temperature of the host star, thus if the planet orbits the dimmer companion it is unlikely to be in the HZ. KOI-1503 hosts a planetary candidate with initial diluted radius estimate of 3.79 R⊕ in a 0.51 AU orbit. If the planet orbits the primary star, the corrected radius of the planet is 4.23±0.07 R⊕, decreasing the probability that it is rocky. KOI-5101, a sun-like star, hosts a near-Earth analog, with a calculated radius 64% larger than Earth and an orbit of 1.12 AU. With a 3.33 magnitude dimmer companion at 1.(cid:48)(cid:48)22, the KOI is likely a 1.68 R⊕ rocky planet if it orbits the primary. 5.3. Stellar Multiplicity and Kepler Planet Candidates We detect 206 planetary candidate hosts with nearby stars from 1629 targets, for an overall multiplicity fraction of 12.6%±0.9% within the detectability range of our survey (∼0.(cid:48)(cid:48)15 -- 4.(cid:48)(cid:48)0, ∆m≤6). For this analysis, we will combined the results in this work with the results from Paper II. With this large dataset we continue our search that began in Paper I for broad-scale correlations between the observed stellar multi- plicity and planetary candidate properties. Such correlations provide an avenue to constrain and test planet formation and Figure 9. The cumulative distribution of nearby stars within a given sep- aration from our observations in Paper II and this work, and the expected distribution from a set of the same number of unassociated stars. For all separations, the observed number of companions we detected is above the expected number if all stars were unassociated. evolution models. Any individual companion found may not be physically bound, however we expect a small number of unassociated asterisms within our complete set of observed targets. An ar- gument for the majority of nearby stars being associated is derived from the observed distribution of companion sepa- rations: if all companions were unassociated background or foreground stars, we would expect a quadratic distribution of companions (i.e., ∼4× the number of objects at 4(cid:48)(cid:48) as at 2(cid:48)(cid:48)). Instead we find a near linear distribution. The dissimilarity between the observed distribution and the distribution of all unassociated objects is shown in Figure 9. In addition, a re- cent follow-up study with the NIRC2 instrument on the Keck- II telescope (Atkinson et al. 2016) observed 84 KOI systems, finding that at least 14.5+3.8−3.4% of companions within ∼4(cid:48)(cid:48) are inconsistent with being physically associated based on multi- band photometric parallax. We therefore expect the overall multiplicity trends to remain relatively unchanged when the unassociated objects are removed. A summary and analysis paper in the Robo-AO survey will investigate the multiplicity properties of Kepler candidates in more detail, including quantifying the effects of association probability and incompleteness. All stellar and planetary properties for the KOIs in this sec- tion were obtained from the cumulative planet candidate list at the NASA Exoplanet Archive12 and have not been corrected for possible dilution due to the presence of nearby stars. 5.3.1. Stellar Multiplicity and KOI Number The early and late public releases of KOIs (Borucki et al. 2011c; Batalha et al. 2013; Burke et al. 2014; Coughlin et al. 2015) could conceivably have a built-in bias, either astrophys- ical in origin or as a result of the initial vetting process by the Kepler team. This bias might appear as a variation in multi- plicity with respect to KOI number. With a target list of KOIs in Paper II and this work widely dispersed in the full KOI dataset, we can search for such a trend. The fraction of KOIs with companions as a function of KOI number, as displayed in Figure 10, shows a sharp decrease at approximately KOI- 5000. We find KOI numbers less than 5000 have a nearby star fraction of 16.1%±0.9% and KOI numbers greater than 5000 have a nearby star fraction of 10.2%±1.5%, a 2.9σ disparity. The exact mechanism for this is unclear, however this may 12 http://exoplanetarchive.ipac.caltech.edu/ Robo-AO Kepler Planetary Candidate Survey III 15 Figure 10. Multiplicity fraction within 4(cid:48)(cid:48)of KOIs as a function of KOI num- ber. A 2.9σ decrease in the fraction of nearby stars between KOIs numbered less than 5000 and greater than 5000 is apparent. Figure 11. Fraction of KOIs with detected nearby (≤4(cid:48)(cid:48)) stars as a function of stellar effective temperature. be a result of better false positive detection in the later data releases due to automation of the vetting process (Mullally et al. 2015). There is no significant corresponding variation in the separations or contrasts of stellar companions between the two populations. 5.3.2. Stellar Multiplicity Rates and Host-star Temperature Revisited It has been well established that stellar multiplicity corre- lates with stellar mass and temperature (Duchene & Kraus 2013). In Paper I, it was found at low significance that this trend appears to also be true for the observed KOIs. Ngo et al. (2015) found in a sample of stars hosting close-in gi- ant planets that, with 2.9σ significance, stars hotter than 6200 K have a companion rate two times larger than their cool counterparts. We find in the combined target sample of Pa- per II and this work that 14.7%±0.9% of KOIs below 6200 K have a companion, compared to 17.2%±2.0% above 6200 K. A Fisher exact test gives an 83% probability that the two samples are indeed from two distinct populations. The trend towards higher multiplicity with higher stellar temperatures is still visually evident, as seen in Figure 11. With an emphasis on solar analogs in the input catalog, the majority of KOIs are FGK-type stars (Batalha et al. 2013), thus the small number of early type stars in our sample prevents any high significance conclusions. 5.3.3. Stellar Multiplicity and Multiple-planet Systems Revisited Figure 12. The multiplicity fraction within 4(cid:48)(cid:48)of KOIs hosting detected single- and multiple-planetary systems. Multiple star systems are thought to more commonly host single transiting planets than multiple planet systems. Per- turbations from the companion star will change the mutual inclinations of planets in the same system (Wang et al. 2014), therefore a lower number of multiple transiting planet systems are expected to have stellar companions. Multiple planet sys- tems are also subject to planet-planet effects (Rasio & Ford 1996; Wang et al. 2015a). In Paper I, we found a low-sigma disparity in multiplic- ity rates between single- and multiple-planet systems, with single-planet systems exhibiting a slightly higher nearby star fraction. With our combined sample from Paper II and this work, we revisit this result with over three times more tar- gets. We find a slightly higher nearby star fraction for multiple planetary systems, displayed in Figure 12. A Fisher exact test gives an 8.7% probability of this being a chance difference. With the expectation, given the effects of stellar perturbations and the higher false positive rate for single star systems, of a higher nearby-star fraction for single-planet candidate hosting stars, even this low-significance result is surprising. A possi- ble explanation is that the additional stellar body in the system is causing orbital migration of outer planets, moving them to shorter period orbits where Kepler has higher sensitivity to transit events. Also, multiple star systems have at least twice as many stars that could host transiting planets, resulting in a higher probability of observing multiple planetary transits. Lastly, with relatively low-significance, this result could also be a consequence of the "look-elsewhere" effect inherent to any multi-comparison study (Gross & Vitells 2010); with the parameter space explored in this section, a result of this sig- nificance is expected to arise approximately 50% of the time out of per chance. Wang et al. (2015b) studied the influence of stellar com- panions on multiple-planet systems, finding a 3.2σ deficit in multiplicity rate in multi-planet systems compared to a con- trol sample of field stars. However, they also found no signif- icant disparity in multiplicity rates between single- and multi- planet systems. 5.3.4. Stellar Multiplicity and Close-in Planets Revisited The presence of stellar companions is hypothesized to shape the formation and evolution of planetary systems. Over- all, there is evidence that planetary formation is disrupted in close binary systems (Fragner et al. 2011; Roell et al. 2012). The third body in the system can lead to Kozai oscil- lations causing orbital migration of the planets (Fabrycky & Tremaine 2007; Katz et al. 2011; Naoz et al. 2012) or tilt the circumstellar disk (Batygin 2012). Smaller planets are also 16 Ziegler et al. Figure 13. 1σ uncertainty regions for the binarity fraction as a function of KOI period for two different planetary populations. Figure 14. 1σ uncertainty regions for the binarity fraction as a function of KOI period for two different planetary populations, with only companions with separations <2.(cid:48)(cid:48)5 used to align with Paper I. more prone to the influence of a stellar companion because of weaker planet-planet dynamical coupling (Wang et al. 2015a). These dynamical interactions between small and large planets in the same system tend to differentially eject small planets more frequently than large planets (Xie et al. 2014). The pres- ence of a stellar companion increases the frequency of these interactions, leading to higher loss of small planets. Conse- quently, we would expect a correlation between binarity and planetary period for different sized planets. We previously reported a low-significance result of stellar third bodies increasing the rate of close-in giant planets, pos- sible evidence of orbital migration of the planet caused by the stellar companion. We revisit the discussion and analysis from Paper I in search of this correlation using the results of Paper II and this work. This analysis splits the "small" and "giant" planets at the arbitrary value of Neptune's radius (3.9 R⊕). The exact value does not significantly affect the results as just 11 of the detected systems have planetary radii within 20% of the cutoff value, with 1635 small and 395 giant planets in total. In Figure 13 the fraction of Kepler planet candidates with nearby stars is shown, with planets grouped into two different size ranges. We again see a small increase in the nearby star fraction for giants with periods <15 days, however the >2σ spike at periods of 2-4 days seen in Paper I is not present. If our sample is reduced to correspond to the separation range of Paper I (ρ<2.(cid:48)(cid:48)5) in Figure 14, again no binarity spike at periods <10 days is apparent. Binning our targets into four population groups in Figure 15 suggests no significant difference in the binarity rate of short period giants. We also attempt to decrease the occurrence Figure 15. Multiplicity fraction of KOIs with four planetary populations, with all contrast ratios and separations ≤4(cid:48)(cid:48). A planet is considered giant if its radius is equal to or larger to that of Neptune (3.9 R⊕). Multi-planet systems can be assigned to multiple populations. of unassociated asterisms by only using close, bright com- panions (∆m≤2, ρ≤1.(cid:48)(cid:48)5). As in Paper I, we detected an ex- cess of close-separation bright companions (Figure 5), which suggests a higher probability of association for these nearby stars. We show the binarity fraction of the four populations in Figure 16. As with the complete set of nearby stars, no sig- nificant differences between the four populations is evident. Any real disparity between the populations would also man- ifest in the physical orbital semi-major axis, which is related to the observable periods by the stellar mass. In Figure 17 we plot the two population's binarity fraction as a function of the calculated semi-major axis of the planetary candidates between 0.01 and 1.0 AU. No significant giant planet binarity spike is observed as in the periods plot. Our updated study using the targets in Paper II and this work suggests that the presence of a second stellar body in planetary systems does not appreciably affect the number of close-in giant planets. This agrees with the analysis of Wang et al. (2015a) who find a relatively uniform multiplicity rate for planets with short and long periods. They note that our previous tentative result may have been due to short-period giants with brighter stellar companions in the visible biasing our detections. Subject to the same potential biases, the larger survey in this analysis does not indicate a period-multiplicity correlation for the two planetary populations, suggesting that our previous low-sigma result may have instead been an arti- fact of small-number statistics. Kraus et al. (2016) find a 6.6σ deficit in binary stars with separation ρ<50 AU in KOIs compared to field stars, again suggesting that close-in stellar companions disrupt the forma- tion and/or evolution of planets, as had been previously hy- pothesized (Wang et al. 2014). Indeed, a quarter of all solar- type stars in the Milky Way are disallowed from hosting plan- etary systems due to the influence of binary companions. Some evidence remains, however, that stellar binarity may encourage the presence of hot Jupiters. A recent NIR survey (Ngo et al. 2015) of exoplanetary systems with known close- in giants finds that hot Jupiter hosts are twice as likely as field stars to be found in a multiple star system, with a significance of 2.8σ. However, the binarity rates of systems containing hot Jupiters remains unclear: 12% (Roell et al. 2012), 38% (Evans et al. 2016), 51% (Ngo et al. 2015). We will revisit this discussion in the last paper in this series where we will combine the full Robo-AO KOI survey dataset. 6. CONCLUSION Robo-AO Kepler Planetary Candidate Survey III 17 imaging of the hundreds of K2 (Howell et al. 2014) plane- tary candidates, ground-based transit surveys such as MEarth (Nutzman & Charbonneau 2008), KELT (Pepper et al. 2007, 2012), HATNet (Bakos et al. 2004), SuperWASP (Pollacco et al. 2006), NGTS (Wheatley et al. 2013), XO (McCullough et al. 2005), and the Evryscope (Law et al. 2015), as well as the thousands of expected exoplanet hosts discovered by the forthcoming NASA Transiting Exoplanet Survey Satellite (TESS, Ricker et al. 2015) and ESA PLAnetary Transits and Oscillations of stars 2.0 (PLATO, Rauer et al. 2014) missions. The Robo-AO survey has completed observations of over 90% of the Kepler planet candidates, with the remaining tar- gets to be observed at the Kitt Peak telescope. Future papers in this survey will present these final KOI targets, and per- form a full probability of association analysis. With the entire survey soon to be completed, providing us with an unprece- dented dataset of thousands of high angular resolution imaged planetary candidates, we can continue our search for clues to planetary formation and evolution. ACKNOWLEDGEMENTS We thank the anonymous referee for careful analysis and useful comments on the manuscript. This research is supported by the NASA Exoplanets Re- search Program, grant #NNX 15AC91G. C.B. acknowledges support from the Alfred P. Sloan Foundation. T.M is sup- ported by NASA grant #NNX 14AE11G under the Kepler Participating Scientist Program. D.A. is supported by a NASA Space Technology Research Fellowship, grant #NNX 13AL75H. The Robo-AO system is supported by collaborating part- ner institutions, the California Institute of Technology and the Inter-University Centre for Astronomy and Astrophysics, and by the National Science Foundation under Grant Nos. AST- 0906060, AST-0960343, and AST-1207891, by the Mount Cuba Astronomical Foundation, and by a gift from Samuel Oschin. We are grateful to the Palomar Observatory staff for their ongoing support of Robo-AO on the 1.5-m telescope, particularly S. Kunsman, M. Doyle, J. Henning, R. Walters, G. Van Idsinga, B. Baker, K. Dunscombe and D. Roderick. Some of the data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foun- dation. Some of the data presented herein is based on ob- servations obtained at the Gemini Observatory, operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership. We recognize and acknowledge the very significant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain. We thank Adam Kraus et al. for sharing a preprint of their This research has made use of the SIMBAD database, op- erated by Centre des Donn´ees Stellaires (Strasbourg, France), and bibliographic references from the Astrophysics Data Sys- tem maintained by SAO/NASA. This research has made use of the Kepler Community FollowUp Observing Program Web site (https://cfop.ipac.caltech.edu) and the NASA Exoplanet Archive, which is operated by the California Institute of Tech- Figure 16. Multiplicity fraction of KOIs with four planetary populations, with only companions with ∆m≤2 and separations ≤1.(cid:48)(cid:48)5, removing the faint nearby stars that are less likely to be physically associated. Figure 17. 1σ uncertainty regions for the binarity fraction as a function of KOI semi-major axis between 0.01 and 1.0 AU for two different planetary populations. We observed 1629 Kepler planetary candidates with the Robo-AO robotic laser adaptive optics system. We detected 206 planetary candidates with nearby stars, implying an over- all nearby-star probability of 12.6%±0.9% at separations be- tween ∼0.(cid:48)(cid:48)15 and 4.(cid:48)(cid:48)0 and ∆m≤6. Many of our newly found companions are of particular in- terest, including 26 habitable zone candidates found within possible multiple star systems. In addition, we found 16 KOIs with multiple nearby stars, and 5 new candidate quadruple star systems hosting planet candidates, including KOI-4495 from Paper II. We looked at broad correlations between the pres- ence of nearby stars and planetary characteristics. We find a higher detected companion rate of systems with multiple planets than in single planet systems. Our previous tentative result of a deficit of close-in giant planets when a third stellar body appears in the system is not apparent in this dataset. The Robo-AO system was installed on the 2.1-m telescope at Kitt Peak in November 2015, and a new low-noise infrared camera that will allow observations of redder companion stars will be added in the future. In addition, a second generation Robo-AO instrument on the University of Hawai'i 2.2-m tele- scope on Maunakea (Baranec et al. 2014a) is being built. The two systems will together image up to ∼500 objects per night and have access to three-quarters of the sky over the course of a year. A southern analog to Robo-AO, mounted on the Southern Astrophysical Research Telescope (SOAR) at CTIO and capable of twice HST resolution imaging, is also in de- velopment. With unmatched efficiency, Robo-AO and its lin- eage of instruments are uniquely able to perform high-acuity paper. 18 Ziegler et al. nology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Pro- gram. This work used the K2fov (Mullally 2016) Python package. Facilities: PO:1.5m (Robo-AO), Keck:II (NIRC2-LGS), Gemini:Gillett (NIRI) REFERENCES Borucki, W. J., Koch, D. G., Basri, G., Batalha, N., Brown, T. M., Bryson, S. T., Caldwell, D., Christensen-Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Gautier, III, T. N., Geary, J. C., Gilliland, R., Gould, A., Howell, S. B., Jenkins, J. M., Latham, D. W., Lissauer, J. J., Marcy, G. W., Rowe, J., Sasselov, D., Boss, A., Charbonneau, D., Ciardi, D., Doyle, L., Dupree, A. K., Ford, E. B., Fortney, J., Holman, M. J., Seager, S., Steffen, J. H., Tarter, J., Welsh, W. F., Allen, C., Buchhave, L. A., Christiansen, J. L., Clarke, B. D., Das, S., D´esert, J.-M., Endl, M., Fabrycky, D., Fressin, F., Haas, M., Horch, E., Howard, A., Isaacson, H., Kjeldsen, H., Kolodziejczak, J., Kulesa, C., Li, J., Lucas, P. W., Machalek, P., McCarthy, D., MacQueen, P., Meibom, S., Miquel, T., Prsa, A., Quinn, S. N., Quintana, E. V., Ragozzine, D., Sherry, W., Shporer, A., Tenenbaum, P., Torres, G., Twicken, J. D., Van Cleve, J., Walkowicz, L., Witteborn, F. C., & Still, M. 2011b, ApJ, 736, 19 -- . 2011c, ApJ, 736, 19 Borucki, W. J., Koch, D. G., Brown, T. M., Basri, G., Batalha, N. M., Caldwell, D. A., Cochran, W. D., Dunham, E. W., Gautier, III, T. N., Geary, J. C., Gilliland, R. L., Howell, S. B., Jenkins, J. M., Latham, D. W., Lissauer, J. J., Marcy, G. W., Monet, D., Rowe, J. F., & Sasselov, D. 2010, ApJL, 713, L126 Brown, T. M., Latham, D. W., Everett, M. E., & Esquerdo, G. A. 2011, AJ, 142, 112 Burgasser, A. J., Kirkpatrick, J. D., Reid, I. N., Brown, M. E., Miskey, C. L., Burke, C. J., Bryson, S. T., Mullally, F., Rowe, J. F., Christiansen, J. L., Thompson, S. E., Coughlin, J. L., Haas, M. R., Batalha, N. M., Caldwell, D. A., Jenkins, J. M., Still, M., Barclay, T., Borucki, W. J., Chaplin, W. J., Ciardi, D. R., Clarke, B. D., Cochran, W. D., Demory, B.-O., Esquerdo, G. A., Gautier, III, T. N., Gilliland, R. L., Girouard, F. R., Havel, M., Henze, C. E., Howell, S. B., Huber, D., Latham, D. W., Li, J., Morehead, R. C., Morton, T. D., Pepper, J., Quintana, E., Ragozzine, D., Seader, S. E., Shah, Y., Shporer, A., Tenenbaum, P., Twicken, J. D., & Wolfgang, A. 2014, ApJS, 210, 19 Ciardi, D. R., Beichman, C. A., Horch, E. P., & Howell, S. B. 2015, ApJ, 805, 16 Ciardi, D. R., von Braun, K., Bryden, G., van Eyken, J., Howell, S. B., Kane, S. R., Plavchan, P., Ram´ırez, S. V., & Stauffer, J. R. 2011, AJ, 141, 108 Coughlin, J. L., Mullally, F., Thompson, S. E., Rowe, J. F., Burke, C. J., Latham, D. W., Batalha, N. M., Ofir, A., Quarles, B. L., Henze, C. E., Wolfgang, A., Caldwell, D. A., Bryson, S. T., Shporer, A., Catanzarite, J., Akeson, R., Barclay, T., Borucki, W. J., Boyajian, T. S., Campbell, J. R., Christiansen, J. L., Girouard, F. R., Haas, M. R., Howell, S. B., Huber, D., Jenkins, J. M., Li, J., Patil-Sabale, A., Quintana, E. V., Ramirez, S., Seader, S., Smith, J. C., Tenenbaum, P., Twicken, J. D., & Zamudio, K. A. 2015, ArXiv e-prints D´esert, J.-M., Charbonneau, D., Torres, G., Fressin, F., Ballard, S., Bryson, S. T., Knutson, H. A., Batalha, N. M., Borucki, W. J., Brown, T. M., Deming, D., Ford, E. B., Fortney, J. J., Gilliland, R. L., Latham, D. W., & Seager, S. 2015, ApJ, 804, 59 Dressing, C. D., Adams, E. R., Dupree, A. K., Kulesa, C., & McCarthy, D. 2014, AJ, 148, 78 Dressing, C. D. & Charbonneau, D. 2013, ApJ, 767, 95 Duchene, G. & Kraus, A. 2013, ARA&A, 51, 269 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Evans, D. F., Southworth, J., Maxted, P. F. L., Skottfelt, J., Hundertmark, M., Jørgensen, U. G., Dominik, M., Alsubai, K. A., Andersen, M. I., Bozza, V., Bramich, D. M., Burgdorf, M. J., Ciceri, S., D'Ago, G., Figuera Jaimes, R., Gu, S. H., Haugbølle, T., Hinse, T. C., Juncher, D., Kains, N., Kerins, E., Korhonen, H., Kuffmeier, M., Peixinho, N., Popovas, A., Rabus, M., Rahvar, S., Schmidt, R. W., Snodgrass, C., Starkey, D., Surdej, J., Tronsgaard, R., von Essen, C., Wang, Y.-B., & Wertz, O. 2016, ArXiv e-prints Everett, M. E., Barclay, T., Ciardi, D. R., Horch, E. P., Howell, S. B., Crepp, J. R., & Silva, D. R. 2015, AJ, 149, 55 Fabrycky, D. & Tremaine, S. 2007, ApJ, 669, 1298 Fabrycky, D. C., Ford, E. B., Steffen, J. H., Rowe, J. F., Carter, J. A., Moorhead, A. V., Batalha, N. M., Borucki, W. J., Bryson, S., Buchhave, L. A., Christiansen, J. L., Ciardi, D. R., Cochran, W. D., Endl, M., Fanelli, M. N., Fischer, D., Fressin, F., Geary, J., Haas, M. R., Hall, J. R., Holman, M. J., Jenkins, J. M., Koch, D. G., Latham, D. W., Li, J., Lissauer, J. J., Lucas, P., Marcy, G. W., Mazeh, T., McCauliff, S., Quinn, S., Ragozzine, D., Sasselov, D., & Shporer, A. 2012, ApJ, 750, 114 Fragner, M. M., Nelson, R. P., & Kley, W. 2011, A&A, 528, A40 Fressin, F., Torres, G., Charbonneau, D., Bryson, S. T., Christiansen, J., Dressing, C. D., Jenkins, J. M., Walkowicz, L. M., & Batalha, N. M. 2013, ApJ, 766, 81 Fruchter, A. S. & Hook, R. N. 2002, PASP, 114, 144 Adams, E. R., Ciardi, D. R., Dupree, A. K., Gautier, III, T. N., Kulesa, C., & & Gizis, J. E. 2003, ApJ, 586, 512 McCarthy, D. 2012, AJ, 144, 42 Adams, E. R., Dupree, A. K., Kulesa, C., & McCarthy, D. 2013, AJ, 146, 9 Atkinson, D., Baranec, C., Ziegler, C., Law, N. M., Riddle, R., & Morton, T. 2016 Bakos, G., Noyes, R. W., Kov´acs, G., Stanek, K. Z., Sasselov, D. D., & Domsa, I. 2004, PASP, 116, 266 Baranec, C., Riddle, R., Law, N. M., Chun, M. R., Lu, J. R., Connelley, M. S., Hall, D., Atkinson, D., & Jacobson, S. 2014a, in Proc. SPIE, Vol. 9148, Adaptive Optics Systems IV, 914812 Baranec, C., Riddle, R., Law, N. M., Ramaprakash, A. N., Tendulkar, S. P., Bui, K., Burse, M. P., Chordia, P., Das, H. K., Davis, J. T. C., Dekany, R. G., Kasliwal, M. M., Kulkarni, S. R., Morton, T. D., Ofek, E. O., & Punnadi, S. 2013, Journal of Visualized Experiments, 72, e50021 Baranec, C., Riddle, R., Law, N. M., Ramaprakash, A. N., Tendulkar, S. P., Hogstrom, K., Bui, K., Burse, M., Chordia, P., Das, H., Dekany, R. G., Kulkarni, S., & Punnadi, S. 2014b, ApJL, 790, L8 Baranec, C., Ziegler, C., Law, N. M., Morton, T., Riddle, R., Atkinson, D., Schonhut, J., & Crepp, J. 2016, ArXiv e-prints Barclay, T., Quintana, E. V., Adams, F. C., Ciardi, D. R., Huber, D., Foreman-Mackey, D., Montet, B. T., & Caldwell, D. 2015, ApJ, 809, 7 Batalha, N. M., Borucki, W. J., Koch, D. G., Bryson, S. T., Haas, M. R., Brown, T. M., Caldwell, D. A., Hall, J. R., Gilliland, R. L., Latham, D. W., Meibom, S., & Monet, D. G. 2010, ApJL, 713, L109 Batalha, N. M., Rowe, J. F., Bryson, S. T., Barclay, T., Burke, C. J., Caldwell, D. A., Christiansen, J. L., Mullally, F., Thompson, S. E., Brown, T. M., Dupree, A. K., Fabrycky, D. C., Ford, E. B., Fortney, J. J., Gilliland, R. L., Isaacson, H., Latham, D. W., Marcy, G. W., Quinn, S. N., Ragozzine, D., Shporer, A., Borucki, W. J., Ciardi, D. R., Gautier, III, T. N., Haas, M. R., Jenkins, J. M., Koch, D. G., Lissauer, J. J., Rapin, W., Basri, G. S., Boss, A. P., Buchhave, L. A., Carter, J. A., Charbonneau, D., Christensen-Dalsgaard, J., Clarke, B. D., Cochran, W. D., Demory, B.-O., Desert, J.-M., Devore, E., Doyle, L. R., Esquerdo, G. A., Everett, M., Fressin, F., Geary, J. C., Girouard, F. R., Gould, A., Hall, J. R., Holman, M. J., Howard, A. W., Howell, S. B., Ibrahim, K. A., Kinemuchi, K., Kjeldsen, H., Klaus, T. C., Li, J., Lucas, P. W., Meibom, S., Morris, R. L., Prsa, A., Quintana, E., Sanderfer, D. T., Sasselov, D., Seader, S. E., Smith, J. C., Steffen, J. H., Still, M., Stumpe, M. C., Tarter, J. C., Tenenbaum, P., Torres, G., Twicken, J. D., Uddin, K., Van Cleve, J., Walkowicz, L., & Welsh, W. F. 2013, ApJS, 204, 24 Batygin, K. 2012, Nature, 491, 418 Borucki, W. J., Koch, D. G., Basri, G., Batalha, N., Boss, A., Brown, T. M., Caldwell, D., Christensen-Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Dupree, A. K., Gautier, III, T. N., Geary, J. C., Gilliland, R., Gould, A., Howell, S. B., Jenkins, J. M., Kjeldsen, H., Latham, D. W., Lissauer, J. J., Marcy, G. W., Monet, D. G., Sasselov, D., Tarter, J., Charbonneau, D., Doyle, L., Ford, E. B., Fortney, J., Holman, M. J., Seager, S., Steffen, J. H., Welsh, W. F., Allen, C., Bryson, S. T., Buchhave, L., Chandrasekaran, H., Christiansen, J. L., Ciardi, D., Clarke, B. D., Dotson, J. L., Endl, M., Fischer, D., Fressin, F., Haas, M., Horch, E., Howard, A., Isaacson, H., Kolodziejczak, J., Li, J., MacQueen, P., Meibom, S., Prsa, A., Quintana, E. V., Rowe, J., Sherry, W., Tenenbaum, P., Torres, G., Twicken, J. D., Van Cleve, J., Walkowicz, L., & Wu, H. 2011a, ApJ, 728, 117 Robo-AO Kepler Planetary Candidate Survey III 19 Gilliland, R. L., Cartier, K. M. S., Adams, E. R., Ciardi, D. R., Kalas, P., & Wright, J. T. 2015, AJ, 149, 24 Gross, E. & Vitells, O. 2010, European Physical Journal C, 70, 525 Haas, M. R., Batalha, N. M., Bryson, S. T., Caldwell, D. A., Dotson, J. L., Hall, J., Jenkins, J. M., Klaus, T. C., Koch, D. G., Kolodziejczak, J., Middour, C., Smith, M., Sobeck, C. K., Stober, J., Thompson, R. S., & Van Cleve, J. E. 2010, ApJL, 713, L115 Heller, R. 2012, A&A, 545, L8 Hodapp, K. W., Jensen, J. B., Irwin, E. M., Yamada, H., Chung, R., Fletcher, K., Robertson, L., Hora, J. L., Simons, D. A., Mays, W., Nolan, R., Bec, M., Merrill, M., & Fowler, A. M. 2003, PASP, 115, 1388 Horch, E. P., Howell, S. B., Everett, M. E., & Ciardi, D. R. 2012, AJ, 144, -- . 2014, ApJ, 795, 60 Howell, S. B., Everett, M. E., Sherry, W., Horch, E., & Ciardi, D. R. 2011, 165 AJ, 142, 19 Howell, S. B., Sobeck, C., Haas, M., Still, M., Barclay, T., Mullally, F., Troeltzsch, J., Aigrain, S., Bryson, S. T., Caldwell, D., Chaplin, W. J., Cochran, W. D., Huber, D., Marcy, G. W., Miglio, A., Najita, J. R., Smith, M., Twicken, J. D., & Fortney, J. J. 2014, PASP, 126, 398 Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108 Katz, B., Dong, S., & Malhotra, R. 2011, Physical Review Letters, 107, Kolbl, R., Marcy, G. W., Isaacson, H., & Howard, A. W. 2015, AJ, 149, 18 Kraus, A. L. & Hillenbrand, L. A. 2007, AJ, 134, 2340 Kraus, A. L., Ireland, M. J., Huber, D., Mann, A. W., & Dupuy, T. J. 2016, Lafreni`ere, D., Marois, C., Doyon, R., Nadeau, D., & Artigau, ´E. 2007, ApJ, 181101 AJ, 152, 8 660, 770 Law, N. M., Fors, O., Ratzloff, J., Wulfken, P., Kavanaugh, D., Sitar, D. J., Pruett, Z., Birchard, M. N., Barlow, B. N., Cannon, K., Cenko, S. B., Dunlap, B., Kraus, A., & Maccarone, T. J. 2015, PASP, 127, 234 Law, N. M., Mackay, C. D., Dekany, R. G., Ireland, M., Lloyd, J. P., Moore, A. M., Robertson, J. G., Tuthill, P., & Woodruff, H. C. 2009, ApJ, 692, 924 Law, N. M., Morton, T., Baranec, C., Riddle, R., Ravichandran, G., Ziegler, C., Johnson, J. A., Tendulkar, S. P., Bui, K., Burse, M. P., Das, H. K., Dekany, R. G., Kulkarni, S., Punnadi, S., & Ramaprakash, A. N. 2014, ApJ, 791, 35 Lawrence, A., Warren, S. J., Almaini, O., Edge, A. C., Hambly, N. C., Jameson, R. F., Lucas, P., Casali, M., Adamson, A., Dye, S., Emerson, J. P., Foucaud, S., Hewett, P., Hirst, P., Hodgkin, S. T., Irwin, M. J., Lodieu, N., McMahon, R. G., Simpson, C., Smail, I., Mortlock, D., & Folger, M. 2007, MNRAS, 379, 1599 Lillo-Box, J., Barrado, D., & Bouy, H. 2012, A&A, 546, A10 -- . 2014, A&A, 566, A103 Lissauer, J. J., Dawson, R. I., & Tremaine, S. 2014, Nature, 513, 336 Marcy, G. W., Isaacson, H., Howard, A. W., Rowe, J. F., Jenkins, J. M., Bryson, S. T., Latham, D. W., Howell, S. B., Gautier, III, T. N., Batalha, N. M., Rogers, L., Ciardi, D., Fischer, D. A., Gilliland, R. L., Kjeldsen, H., Christensen-Dalsgaard, J., Huber, D., Chaplin, W. J., Basu, S., Buchhave, L. A., Quinn, S. N., Borucki, W. J., Koch, D. G., Hunter, R., Caldwell, D. A., Van Cleve, J., Kolbl, R., Weiss, L. M., Petigura, E., Seager, S., Morton, T., Johnson, J. A., Ballard, S., Burke, C., Cochran, W. D., Endl, M., MacQueen, P., Everett, M. E., Lissauer, J. J., Ford, E. B., Torres, G., Fressin, F., Brown, T. M., Steffen, J. H., Charbonneau, D., Basri, G. S., Sasselov, D. D., Winn, J., Sanchis-Ojeda, R., Christiansen, J., Adams, E., Henze, C., Dupree, A., Fabrycky, D. C., Fortney, J. J., Tarter, J., Holman, M. J., Tenenbaum, P., Shporer, A., Lucas, P. W., Welsh, W. F., Orosz, J. A., Bedding, T. R., Campante, T. L., Davies, G. R., Elsworth, Y., Handberg, R., Hekker, S., Karoff, C., Kawaler, S. D., Lund, M. N., Lundkvist, M., Metcalfe, T. S., Miglio, A., Silva Aguirre, V., Stello, D., White, T. R., Boss, A., Devore, E., Gould, A., Prsa, A., Agol, E., Barclay, T., Coughlin, J., Brugamyer, E., Mullally, F., Quintana, E. V., Still, M., Thompson, S. E., Morrison, D., Twicken, J. D., D´esert, J.-M., Carter, J., Crepp, J. R., H´ebrard, G., Santerne, A., Moutou, C., Sobeck, C., Hudgins, D., Haas, M. R., Robertson, P., Lillo-Box, J., & Barrado, D. 2014, ApJS, 210, 20 McCullough, P. R., Stys, J. E., Valenti, J. A., Fleming, S. W., Janes, K. A., & Heasley, J. N. 2005, PASP, 117, 783 Morton, T. D., Bryson, S. T., Coughlin, J. L., Rowe, J. F., Ravichandran, G., Petigura, E. A., Haas, M. R., & Batalha, N. M. 2016, The Astrophysical Journal, 822, 86 Morton, T. D. & Johnson, J. A. 2011, ApJ, 738, 170 Mullally, F., Coughlin, J. L., Thompson, S. E., Rowe, J., Burke, C., Latham, D. W., Batalha, N. M., Bryson, S. T., Christiansen, J., Henze, C. E., Ofir, A., Quarles, B., Shporer, A., Van Eylen, V., Van Laerhoven, C., Shah, Y., Wolfgang, A., Chaplin, W. J., Xie, J.-W., Akeson, R., Argabright, V., Bachtell, E., Barclay, T., Borucki, W. J., Caldwell, D. A., Campbell, J. R., Catanzarite, J. H., Cochran, W. D., Duren, R. M., Fleming, S. W., Fraquelli, D., Girouard, F. R., Haas, M. R., Hełminiak, K. G., Howell, S. B., Huber, D., Larson, K., Gautier, III, T. N., Jenkins, J. M., Li, J., Lissauer, J. J., McArthur, S., Miller, C., Morris, R. L., Patil-Sabale, A., Plavchan, P., Putnam, D., Quintana, E. V., Ramirez, S., Silva Aguirre, V., Seader, S., Smith, J. C., Steffen, J. H., Stewart, C., Stober, J., Still, M., Tenenbaum, P., Troeltzsch, J., Twicken, J. D., & Zamudio, K. A. 2015, ApJS, 217, 31 Mullally, Fergal; Barclay, T. B. G. 2016, K2fov: Field of view software for NASA's K2 mission, Astrophysics Source Code Library Naoz, S., Farr, W. M., & Rasio, F. A. 2012, ApJL, 754, L36 Ngo, H., Knutson, H. A., Hinkley, S., Crepp, J. R., Bechter, E. B., Batygin, K., Howard, A. W., Johnson, J. A., Morton, T. D., & Muirhead, P. S. 2015, ApJ, 800, 138 Nutzman, P. & Charbonneau, D. 2008, PASP, 120, 317 Pepper, J., Kuhn, R. B., Siverd, R., James, D., & Stassun, K. 2012, PASP, 124, 230 Pepper, J., Pogge, R. W., DePoy, D. L., Marshall, J. L., Stanek, K. Z., Stutz, A. M., Poindexter, S., Siverd, R., O'Brien, T. P., Trueblood, M., & Trueblood, P. 2007, PASP, 119, 923 Pickles, A. J. 1998, PASP, 110, 863 Pollacco, D. L., Skillen, I., Collier Cameron, A., Christian, D. J., Hellier, C., Irwin, J., Lister, T. A., Street, R. A., West, R. G., Anderson, D. R., Clarkson, W. I., Deeg, H., Enoch, B., Evans, A., Fitzsimmons, A., Haswell, C. A., Hodgkin, S., Horne, K., Kane, S. R., Keenan, F. P., Maxted, P. F. L., Norton, A. J., Osborne, J., Parley, N. R., Ryans, R. S. I., Smalley, B., Wheatley, P. J., & Wilson, D. M. 2006, PASP, 118, 1407 Raghavan, D., McAlister, H. A., Henry, T. J., Latham, D. W., Marcy, G. W., Mason, B. D., Gies, D. R., White, R. J., & ten Brummelaar, T. A. 2010, ApJS, 190, 1 Rasio, F. A. & Ford, E. B. 1996, Science, 274, 954 Rauer, H., Catala, C., Aerts, C., Appourchaux, T., Benz, W., Brandeker, A., Christensen-Dalsgaard, J., Deleuil, M., Gizon, L., Goupil, M.-J., Gudel, M., Janot-Pacheco, E., Mas-Hesse, M., Pagano, I., Piotto, G., Pollacco, D., Santos, C., Smith, A., Su´arez, J.-C., Szab´o, R., Udry, S., Adibekyan, V., Alibert, Y., Almenara, J.-M., Amaro-Seoane, P., Eiff, M. A.-v., Asplund, M., Antonello, E., Barnes, S., Baudin, F., Belkacem, K., Bergemann, M., Bihain, G., Birch, A. C., Bonfils, X., Boisse, I., Bonomo, A. S., Borsa, F., Brandao, I. M., Brocato, E., Brun, S., Burleigh, M., Burston, R., Cabrera, J., Cassisi, S., Chaplin, W., Charpinet, S., Chiappini, C., Church, R. P., Csizmadia, S., Cunha, M., Damasso, M., Davies, M. B., Deeg, H. J., D´ıaz, R. F., Dreizler, S., Dreyer, C., Eggenberger, P., Ehrenreich, D., Eigmuller, P., Erikson, A., Farmer, R., Feltzing, S., de Oliveira Fialho, F., Figueira, P., Forveille, T., Fridlund, M., Garc´ıa, R. A., Giommi, P., Giuffrida, G., Godolt, M., Gomes da Silva, J., Granzer, T., Grenfell, J. L., Grotsch-Noels, A., Gunther, E., Haswell, C. A., Hatzes, A. P., H´ebrard, G., Hekker, S., Helled, R., Heng, K., Jenkins, J. M., Johansen, A., Khodachenko, M. L., Kislyakova, K. G., Kley, W., Kolb, U., Krivova, N., Kupka, F., Lammer, H., Lanza, A. F., Lebreton, Y., Magrin, D., Marcos-Arenal, P., Marrese, P. M., Marques, J. P., Martins, J., Mathis, S., Mathur, S., Messina, S., Miglio, A., Montalban, J., Montalto, M., Monteiro, M. J. P. F. G., Moradi, H., Moravveji, E., Mordasini, C., Morel, T., Mortier, A., Nascimbeni, V., Nelson, R. P., Nielsen, M. B., Noack, L., Norton, A. J., Ofir, A., Oshagh, M., Ouazzani, R.-M., Papics, P., Parro, V. C., Petit, P., Plez, B., Poretti, E., Quirrenbach, A., Ragazzoni, R., Raimondo, G., Rainer, M., Reese, D. R., Redmer, R., Reffert, S., Rojas-Ayala, B., Roxburgh, I. W., Salmon, S., Santerne, A., Schneider, J., Schou, J., Schuh, S., Schunker, H., Silva-Valio, A., Silvotti, R., Skillen, I., Snellen, I., Sohl, F., Sousa, S. G., Sozzetti, A., Stello, D., Strassmeier, K. G., Svanda, M., Szab´o, G. M., Tkachenko, A., Valencia, D., Van Grootel, V., Vauclair, S. D., Ventura, P., Wagner, F. W., Walton, N. A., Weingrill, J., Werner, S. C., Wheatley, P. J., & Zwintz, K. 2014, Experimental Astronomy, 38, 249 20 Ziegler et al. Ricker, G. R., Winn, J. N., Vanderspek, R., Latham, D. W., Bakos, G. ´A., Bean, J. L., Berta-Thompson, Z. K., Brown, T. M., Buchhave, L., Butler, N. R., Butler, R. P., Chaplin, W. J., Charbonneau, D., Christensen-Dalsgaard, J., Clampin, M., Deming, D., Doty, J., De Lee, N., Dressing, C., Dunham, E. W., Endl, M., Fressin, F., Ge, J., Henning, T., Holman, M. J., Howard, A. W., Ida, S., Jenkins, J. M., Jernigan, G., Johnson, J. A., Kaltenegger, L., Kawai, N., Kjeldsen, H., Laughlin, G., Levine, A. M., Lin, D., Lissauer, J. J., MacQueen, P., Marcy, G., McCullough, P. R., Morton, T. D., Narita, N., Paegert, M., Palle, E., Pepe, F., Pepper, J., Quirrenbach, A., Rinehart, S. A., Sasselov, D., Sato, B., Seager, S., Sozzetti, A., Stassun, K. G., Sullivan, P., Szentgyorgyi, A., Torres, G., Udry, S., & Villasenor, J. 2015, Journal of Astronomical Telescopes, Instruments, and Systems, 1, 014003 Riddle, R. L., Burse, M. P., Law, N. M., Tendulkar, S. P., Baranec, C., Rudy, A. R., Sitt, M., Arya, A., Papadopoulos, A., Ramaprakash, A. N., & Dekany, R. G. 2012, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8447, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 2 Roberts, Jr., L. C., Tokovinin, A., Mason, B. D., Riddle, R. L., Hartkopf, W. I., Law, N. M., & Baranec, C. 2015, AJ, 149, 118 Roell, T., Neuhauser, R., Seifahrt, A., & Mugrauer, M. 2012, A&A, 542, Rogers, L. A. 2015, ApJ, 801, 41 Rowe, J. F., Bryson, S. T., Marcy, G. W., Lissauer, J. J., Jontof-Hutter, D., Mullally, F., Gilliland, R. L., Issacson, H., Ford, E., Howell, S. B., Borucki, W. J., Haas, M., Huber, D., Steffen, J. H., Thompson, S. E., Quintana, E., Barclay, T., Still, M., Fortney, J., Gautier, III, T. N., Hunter, R., Caldwell, D. A., Ciardi, D. R., Devore, E., Cochran, W., Jenkins, J., Agol, E., Carter, J. A., & Geary, J. 2014, ApJ, 784, 45 Santerne, A., D´ıaz, R. F., Moutou, C., Bouchy, F., H´ebrard, G., Almenara, J.-M., Bonomo, A. S., Deleuil, M., & Santos, N. C. 2012, A&A, 545, A76 Santerne, A., Fressin, F., D´ıaz, R. F., Figueira, P., Almenara, J.-M., & Santos, N. C. 2013, A&A, 557, A139 Santerne, A., Moutou, C., Tsantaki, M., Bouchy, F., H´ebrard, G., Adibekyan, V., Almenara, J.-M., Amard, L., Barros, S. C. C., Boisse, I., Bonomo, A. S., Bruno, G., Courcol, B., Deleuil, M., Demangeon, O., D´ıaz, R. F., Guillot, T., Havel, M., Montagnier, G., Rajpurohit, A. S., Rey, J., & Santos, N. C. 2015, ArXiv e-prints Schwamb, M. E., Orosz, J. A., Carter, J. A., Welsh, W. F., Fischer, D. A., Torres, G., Howard, A. W., Crepp, J. R., Keel, W. C., Lintott, C. J., Kaib, N. A., Terrell, D., Gagliano, R., Jek, K. J., Parrish, M., Smith, A. M., Lynn, S., Simpson, R. J., Giguere, M. J., & Schawinski, K. 2013, ApJ, 768, 127 A92 Seager, S. 2013, Science, 340, 577 Selsis, F., Kasting, J. F., Levrard, B., Paillet, J., Ribas, I., & Delfosse, X. 2007, A&A, 476, 1373 Skrutskie, M. F., Cutri, R. M., Stiening, R., Weinberg, M. D., Schneider, S., Carpenter, J. M., Beichman, C., Capps, R., Chester, T., Elias, J., Huchra, J., Liebert, J., Lonsdale, C., Monet, D. G., Price, S., Seitzer, P., Jarrett, T., Kirkpatrick, J. D., Gizis, J. E., Howard, E., Evans, T., Fowler, J., Fullmer, L., Hurt, R., Light, R., Kopan, E. L., Marsh, K. A., McCallon, H. L., Tam, R., Van Dyk, S., & Wheelock, S. 2006, AJ, 131, 1163 Torres, G., Kipping, D. M., Fressin, F., Caldwell, D. A., Twicken, J. D., Ballard, S., Batalha, N. M., Bryson, S. T., Ciardi, D. R., Henze, C. E., Howell, S. B., Isaacson, H. T., Jenkins, J. M., Muirhead, P. S., Newton, E. R., Petigura, E. A., Barclay, T., Borucki, W. J., Crepp, J. R., Everett, M. E., Horch, E. P., Howard, A. W., Kolbl, R., Marcy, G. W., McCauliff, S., & Quintana, E. V. 2015, ApJ, 800, 99 van Dam, M. A., Bouchez, A. H., Le Mignant, D., Johansson, E. M., Wizinowich, P. L., Campbell, R. D., Chin, J. C. Y., Hartman, S. K., Lafon, R. E., Stomski, Jr., P. J., & Summers, D. M. 2006, PASP, 118, 310 Wang, J., Fischer, D. A., Horch, E. P., & Xie, J.-W. 2015a, ApJ, 806, 248 Wang, J., Fischer, D. A., Xie, J.-W., & Ciardi, D. R. 2014, ApJ, 791, 111 -- . 2015b, ArXiv e-prints Wheatley, P. J., Pollacco, D. L., Queloz, D., Rauer, H., Watson, C. A., West, R. G., Chazelas, B., Louden, T. M., Walker, S., Bannister, N., Bento, J., Burleigh, M., Cabrera, J., Eigmuller, P., Erikson, A., Genolet, L., Goad, M., Grange, A., Jord´an, A., Lawrie, K., McCormac, J., & Neveu, M. 2013, in European Physical Journal Web of Conferences, Vol. 47, European Physical Journal Web of Conferences, 13002 Wizinowich, P., Acton, D. S., Shelton, C., Stomski, P., Gathright, J., Ho, K., Lupton, W., Tsubota, K., Lai, O., Max, C., Brase, J., An, J., Avicola, K., Olivier, S., Gavel, D., Macintosh, B., Ghez, A., & Larkin, J. 2000, PASP, 112, 315 Wizinowich, P. L., Le Mignant, D., Bouchez, A. H., Campbell, R. D., Chin, J. C. Y., Contos, A. R., van Dam, M. A., Hartman, S. K., Johansson, E. M., Lafon, R. E., Lewis, H., Stomski, P. J., Summers, D. M., Brown, C. G., Danforth, P. M., Max, C. E., & Pennington, D. M. 2006, PASP, 118, 297 Xie, J.-W., Wu, Y., & Lithwick, Y. 2014, ApJ, 789, 165 Yelda, S., Lu, J. R., Ghez, A. M., Clarkson, W., Anderson, J., Do, T., & Matthews, K. 2010, ApJ, 725, 331 Ziegler, C., Law, N. M., Baranec, C., Riddle, R. L., & Fuchs, J. T. 2015, ApJ, 804, 30 Zsom, A., Seager, S., de Wit, J., & Stamenkovi´c, V. 2013, ApJ, 778, 109 Robo-AO Kepler Planetary Candidate Survey III 21 Figure 18. Color inverted, normalized log-scale cutouts of 61 multiple KOI systems [KOI-51 to KOI-2688] with separations <4(cid:48)(cid:48) resolved with Robo-AO. The angular scale and orientation is similar for each cutout. The smaller circles are centered on the detected nearby star, and the larger circle is the limit of the survey's 4(cid:48)(cid:48)separation range. 22 Ziegler et al. Figure 19. Color inverted, normalized log-scale cutouts of 61 multiple KOI systems [KOI-2744 to KOI-4405] with separations <4(cid:48)(cid:48) resolved with Robo-AO. The angular scale and orientation is similar for each cutout. The smaller circles are centered on the detected nearby star, and the larger circle is the limit of the survey's 4(cid:48)(cid:48)separation range. Robo-AO Kepler Planetary Candidate Survey III 23 Figure 20. Color inverted, normalized log-scale cutouts of 61 multiple KOI systems [KOI-4418 to KOI-6311] with separations <4(cid:48)(cid:48) resolved with Robo-AO. The angular scale and orientation is similar for each cutout. The smaller circles are centered on the detected nearby star, and the larger circle is the limit of the survey's 4(cid:48)(cid:48)separation range. 24 Ziegler et al. Figure 21. Color inverted, normalized log-scale cutouts of 23 multiple KOI systems [KOI-6329 to KOI-7448] with separations <4(cid:48)(cid:48) resolved with Robo-AO. The angular scale and orientation is similar for each cutout. The smaller circles are centered on the detected nearby star, and the larger circle is the limit of the survey's 4(cid:48)(cid:48)separation range. Robo-AO Kepler Planetary Candidate Survey III 25 7. APPENDIX In Table 9, we list our Robo-AO observed KOIs, includ- ing date the target was observed, observation quality (as de- scribed in Section 3.6), the estimated latest detectable com- panion spectral type (as described in Section 3.7.3), and the presence of detected companions. Table 9 Full Robo-AO Observation List Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) medium high high yes yes KOI K020 K051 K076 K0104 K0126 K0134 K0135 K0163 K0186 K0193 K0195 K0196 K0199 K0200 K0204 K0212 K0217 K0225 K0226 K0240 K0242 K0249 K0252 K0255 K0262 K0266 K0326 K0364 K0376 K0398 K0414 K0422 K0423 K0426 K0428 K0430 K0449 K0452 K0454 K0458 K0466 K0467 K0468 K0469 K0476 K0483 K0510 K0513 K0524 K0530 K0532 K0533 K0537 K0541 K0547 K0557 K0560 K0566 K0575 K0578 K0580 K0581 K0583 K0585 K0598 K0599 K0600 mi (mags) 13.29 14.48 10.05 12.78 13.11 15.02 13.8 13.3 14.76 14.9 14.62 14.24 14.72 14.21 14.41 14.66 14.88 14.61 14.54 14.8 14.5 14.77 14.99 14.52 10.3 11.47 12.96 9.91 13.7 15.04 14.5 14.56 14.15 14.53 14.42 14.4 14.18 14.45 14.53 14.49 14.49 14.58 14.47 14.54 14.65 14.45 14.4 14.71 14.6 14.69 14.53 14.41 14.5 14.51 14.48 14.67 14.46 14.53 14.51 14.47 14.65 14.58 14.52 14.68 14.55 14.66 14.64 ObsID low low high high medium medium low high high high i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) low high high low high low high high i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) 2014 Aug 22 medium 2013 Jul 25 LP600 medium 2014 Aug 22 2014 Aug 22 2015 Jun 08 LP600 2012 Jul 16 2012 Jul 17 2012 Jul 18 LP600 2014 Aug 31 LP600 medium 2014 Aug 21 LP600 2014 Aug 22 2014 Aug 21 2014 Aug 20 2014 Sep 01 LP600 medium 2014 Aug 20 2014 Aug 20 2014 Aug 20 2014 Jul 16 LP600 2014 Jun 19 LP600 medium 2014 Aug 21 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 23 LP600 medium 2014 Aug 22 2014 Jul 17 LP600 medium 2012 Jul 28 2014 Aug 20 2013 Aug 15 LP600 2014 Aug 20 2012 Sep 01 LP600 medium 2014 Aug 22 2012 Sep 01 LP600 medium 2014 Aug 22 2014 Jul 16 LP600 medium 2014 Aug 20 2014 Jun 19 LP600 medium 2014 Aug 22 2015 Jun 04 LP600 2014 Jul 17 LP600 medium 2014 Jul 16 LP600 medium 2014 Jul 14 LP600 medium 2014 Sep 01 LP600 2014 Aug 21 LP600 medium 2014 Aug 21 LP600 2014 Sep 02 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 22 2014 Jul 14 LP600 medium 2014 Jul 14 LP600 2014 Aug 23 LP600 2014 Sep 02 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 22 2014 Jul 19 LP600 2014 Aug 29 LP600 medium 2014 Jul 18 LP600 medium 2014 Jul 16 LP600 medium 2014 Jul 14 LP600 medium 2014 Sep 01 LP600 medium 2014 Jun 19 LP600 2014 Jun 19 LP600 medium 2014 Sep 01 LP600 medium 2014 Jul 16 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 21 LP600 medium 2014 Jul 11 LP600 2014 Sep 03 LP600 medium 2014 Jul 17 LP600 low low low low low low high high high high high high high i(cid:48) i(cid:48) M4 M4 M4 M6 M5 M4 M1 M6 M4 M4 M1 M5 M4 M4 M2 M4 M5 M0 M5 M3 M4 M7 M5 M7 M4 M3 M6 M2 M3 M4 M4 M3 M4 M2 M5 M0 M3 M5 M4 M4 M4 M5 M3 M5 M5 M4 M0 M3 M4 M3 M3 M1 M4 M5 M5 M5 M3 M4 M4 M4 M4 M4 M4 M5 M4 M4 yes yes yes yes yes yes yes yes yes yes yes yes KOI K0602 K0605 K0622 K0732 K0733 K0734 K0736 K0740 K0746 K0749 K0750 K0752 K0755 K0757 K0759 K0760 K0763 K0764 K0765 K0766 K0769 K0771 K0772 K0773 K0774 K0775 K0776 K0777 K0778 K0780 K0782 K0783 K0784 K0786 K0788 K0791 K0804 K0805 K0806 K0809 K0810 K0811 K0812 K0814 K0815 K0816 K0826 K0829 K0830 K0833 K0838 K0841 K0843 K0845 K0847 K0849 K0850 K0855 K0856 K0858 K0861 K0863 K0864 K0867 K0868 K0869 K0870 K0871 K0873 K0881 K0887 K0889 K0890 K0891 K0892 K0893 K0900 mi (mags) 14.48 14.47 14.65 15.07 15.36 15.14 15.45 15.2 14.96 15.4 15.04 15.3 15.29 15.53 14.84 15.07 15.33 15.11 15.08 15.3 15.11 15.07 15.06 14.96 15.08 14.57 15.27 15.22 14.61 15.0 15.1 14.82 14.86 15.03 14.92 14.92 15.11 15.39 15.3 15.32 14.81 15.04 15.43 15.32 15.42 15.42 14.87 15.18 14.91 15.24 15.1 15.57 15.04 15.21 14.97 14.76 15.03 14.95 15.15 14.82 14.71 15.3 15.36 14.92 15.1 15.29 14.63 15.02 14.8 15.56 14.8 14.95 15.08 14.87 14.89 15.44 15.21 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) low low high high low low low low low high low high low high low low high 2014 Aug 29 LP600 medium 2014 Aug 22 2014 Aug 31 LP600 medium 2014 Sep 02 LP600 2014 Sep 02 LP600 2014 Sep 02 LP600 medium 2014 Sep 01 LP600 2014 Aug 22 2014 Sep 02 LP600 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 2014 Sep 01 LP600 medium 2014 Sep 03 LP600 medium 2014 Sep 03 LP600 medium 2014 Jul 18 LP600 2014 Sep 01 LP600 medium 2014 Aug 28 LP600 2014 Aug 23 LP600 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Aug 28 LP600 2014 Aug 27 LP600 medium 2014 Sep 03 LP600 medium 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 medium 2014 Aug 22 2014 Sep 02 LP600 2014 Aug 31 LP600 medium 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 2014 Jul 11 LP600 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 medium 2014 Sep 02 LP600 medium 2014 Sep 02 LP600 2014 Aug 27 LP600 medium 2014 Aug 23 LP600 2014 Aug 23 LP600 2014 Aug 23 LP600 2014 Aug 31 LP600 2014 Aug 31 LP600 2014 Aug 22 2014 Aug 24 LP600 2014 Nov 09 LP600 medium 2014 Aug 24 LP600 medium 2014 Jun 19 LP600 medium 2014 Aug 22 2014 Jun 19 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 27 LP600 medium 2012 Sep 02 LP600 2014 Aug 26 LP600 medium 2014 Nov 09 LP600 medium 2014 Jul 16 LP600 2014 Jul 19 LP600 medium 2014 Aug 26 LP600 2014 Jun 19 LP600 medium 2014 Aug 26 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 27 LP600 2014 Sep 01 LP600 2014 Jun 19 LP600 2014 Jun 19 LP600 2014 Sep 01 LP600 medium 2014 Aug 22 2014 Aug 28 LP600 medium 2014 Jul 18 LP600 2014 Aug 20 2014 Aug 20 2014 Jul 16 LP600 medium 2014 Aug 23 LP600 2014 Jul 16 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 31 LP600 2014 Aug 24 LP600 medium high low high low low low high low low high low low medium high high high low low low low low i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) yes yes yes yes yes yes M3 M5 M5 M3 M3 M4 M5 M4 M6 M4 M4 M4 M4 M5 M5 M3 M1 M5 M4 M1 M5 M4 M3 M4 M4 M5 M5 M4 M7 M5 M1 M3 M7 M4 M5 M5 M5 M5 M3 M5 M3 M4 M6 M5 M4 M4 M4 M1 M5 M4 M4 M3 M4 M4 M2 M4 M5 M4 M3 M4 M5 M1 M5 M3 M5 M5 M4 M4 M5 M3 M2 M5 M4 M4 M5 M1 M4 26 KOI K0901 K0902 K0904 K0910 K0911 K0912 K0913 K0914 K0916 K0918 K0920 K0923 K0924 K0928 K0929 K0934 K0937 K0942 K0943 K0948 K0949 K0951 K0952 K0953 K0954 K0955 K0956 K0958 K0961 K0989 K01005 K01017 K01051 K01053 K01072 K01074 K01081 K01082 K01095 K01096 K01121 K01129 K01137 K01142 K01160 K01164 K01166 K01170 K01176 K01193 K01196 K01201 K01204 K01207 K01212 K01232 K01246 K01255 K01261 K01264 K01268 K01273 K01298 K01312 K01326 K01337 K01351 K01356 K01361 K01367 K01369 K01377 K01385 K01387 K01395 K01398 K01402 mi (mags) 15.27 15.7 15.35 15.35 15.21 14.59 14.97 15.15 14.88 14.77 14.83 15.32 15.04 15.02 15.44 15.63 15.15 15.08 15.45 15.35 15.28 14.85 14.65 15.71 14.5 14.9 15.1 12.24 15.34 15.4 15.39 14.75 15.19 15.12 14.55 15.26 15.04 15.39 15.37 14.48 12.9 15.12 13.77 15.48 15.72 15.1 15.22 14.45 15.31 15.05 14.82 14.96 15.11 14.89 14.76 14.14 14.72 15.68 14.92 15.52 14.64 14.63 15.8 14.56 12.8 14.53 15.35 14.98 14.46 14.74 14.67 14.6 15.64 14.46 15.71 15.67 15.69 Ziegler et al. TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) low low low high high low low low low low high low low low high high low low 2014 Aug 23 LP600 2014 Aug 31 LP600 2014 Aug 23 LP600 2014 Aug 26 LP600 2014 Aug 28 LP600 medium 2014 Jul 17 LP600 2014 Jul 17 LP600 medium 2014 Aug 28 LP600 2014 Jul 18 LP600 2014 Jul 16 LP600 medium 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 23 LP600 2014 Aug 27 LP600 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Sep 02 LP600 2014 Sep 02 LP600 medium 2014 Aug 31 LP600 medium 2012 Jul 16 2014 Sep 03 LP600 medium 2014 Sep 02 LP600 2014 Aug 29 LP600 medium 2014 Jul 16 LP600 medium 2012 Aug 02 medium 2012 Aug 04 LP600 medium 2014 Aug 24 LP600 2014 Nov 09 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 27 LP600 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 2014 Aug 26 LP600 2014 Sep 01 LP600 2014 Aug 31 LP600 2014 Aug 22 2012 Sep 14 LP600 2014 Aug 31 LP600 2014 Jun 13 LP600 medium 2014 Aug 31 LP600 2014 Sep 02 LP600 2012 Aug 04 LP600 medium 2014 Aug 26 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Aug 26 LP600 medium 2014 Jul 16 LP600 2012 Aug 04 LP600 medium 2014 Nov 08 LP600 medium 2014 Sep 01 LP600 2014 Jul 11 LP600 2014 Aug 21 2014 Jun 19 LP600 2014 Sep 01 LP600 2014 Aug 22 2014 Aug 31 LP600 2014 Aug 22 2014 Jul 17 LP600 2014 Aug 21 2014 Jul 17 LP600 medium 2014 Aug 22 medium 2014 Jul 14 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 22 2014 Aug 20 2014 Jul 16 LP600 medium 2014 Jun 19 LP600 medium 2014 Jul 17 LP600 2014 Sep 01 LP600 2014 Jul 16 LP600 medium 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Aug 24 LP600 low low low low high high low high high low low low low low low high low i(cid:48) i(cid:48) i(cid:48) i(cid:48) high low medium low low low low low low high low i(cid:48) i(cid:48) i(cid:48) M5 M5 M4 M5 M4 M4 M4 M2 M3 M4 M4 M4 M3 M3 M1 M1 M5 M5 M3 M3 M4 M5 M8 M4 M1 M3 M5 M5 M3 M5 M4 M4 M5 M3 M0 M1 M3 M2 M5 M5 M3 M5 M3 M3 M7 M4 M4 M8 M4 M1 M7 M3 M5 M1 M5 M4 M5 M1 M3 M0 M2 M3 M3 M5 M5 M4 M5 M5 M4 M0 M1 M4 M3 M5 M1 yes yes yes yes KOI K01403 K01404 K01405 K01409 K01410 K01411 K01420 K01421 K01423 K01424 K01429 K01430 K01433 K01434 K01437 K01440 K01441 K01447 K01456 K01457 K01466 K01472 K01476 K01477 K01481 K01483 K01484 K01488 K01489 K01494 K01496 K01498 K01501 K01503 K01505 K01506 K01507 K01508 K01510 K01511 K01517 K01519 K01521 K01526 K01547 K01558 K01561 K01562 K01564 K01569 K01570 K01572 K01577 K01581 K01582 K01583 K01584 K01587 K01593 K01595 K01630 K01633 K01645 K01650 K01651 K01656 K01658 K01659 K01660 K01662 K01672 K01675 K01681 K01685 K01687 K01691 K01693 mi (mags) 14.99 15.49 15.79 14.98 15.12 13.13 15.45 15.09 15.49 14.77 15.3 15.02 15.43 14.41 15.09 15.27 14.9 13.0 14.77 15.31 15.59 14.83 15.52 15.65 15.22 14.11 14.96 15.3 15.25 15.46 15.38 15.78 15.47 14.6 15.49 14.77 15.07 15.47 15.58 14.87 14.47 15.06 14.82 15.04 15.47 14.95 15.32 15.37 15.07 15.22 14.98 15.25 15.49 15.23 15.16 14.85 15.44 15.41 15.57 14.7 14.91 14.92 13.14 15.04 14.55 14.82 13.16 14.95 15.41 15.21 15.37 15.56 14.71 14.65 14.94 15.56 14.73 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) low low low i(cid:48) i(cid:48) high high low low low low high low low low high high high low low low low high 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 28 LP600 2014 Jun 15 LP600 medium 2014 Sep 02 LP600 2014 Aug 22 2014 Sep 03 LP600 medium 2014 Sep 01 LP600 medium 2014 Sep 01 LP600 2014 Aug 23 LP600 medium 2014 Sep 03 LP600 medium 2014 Jul 17 LP600 medium 2014 Sep 03 LP600 medium 2014 Sep 01 LP600 2014 Aug 31 LP600 medium 2012 Sep 04 LP600 2014 Jul 17 LP600 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Aug 20 2014 Sep 02 LP600 2014 Aug 22 2014 Aug 30 LP600 2014 Sep 02 LP600 medium 2014 Jul 13 LP600 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Aug 27 LP600 2015 Jun 04 LP600 2014 Aug 28 LP600 2014 Aug 22 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 medium 2014 Aug 28 LP600 2014 Aug 31 LP600 2014 Aug 28 LP600 2014 Jul 17 LP600 2014 Jun 16 LP600 2014 Aug 23 LP600 2015 Jun 08 LP600 medium 2014 Sep 02 LP600 2014 Aug 24 LP600 2014 Jul 11 LP600 2014 Aug 23 LP600 2014 Aug 21 LP600 2014 Aug 27 LP600 medium 2014 Aug 24 LP600 medium 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 2014 Sep 02 LP600 2014 Aug 24 LP600 2014 Aug 28 LP600 2014 Jul 13 LP600 2014 Sep 02 LP600 2014 Aug 27 LP600 2014 Aug 24 LP600 2014 Jul 13 LP600 2014 Jul 16 LP600 2014 Aug 21 LP600 medium 2014 Jun 19 LP600 2014 Sep 03 LP600 medium 2014 Jun 19 LP600 2014 Jun 13 LP600 2014 Jul 16 LP600 medium 2014 Jul 16 LP600 2014 Aug 28 LP600 2014 Aug 24 LP600 2014 Aug 28 LP600 2014 Aug 27 LP600 2014 Aug 24 LP600 2014 Jul 16 LP600 2014 Jul 17 LP600 medium 2014 Aug 26 LP600 2014 Jun 19 LP600 medium high low low low low high high low low high high low high low low high low low low high low high high low high high low high low low high high M6 M5 M3 M4 M1 M4 M4 M1 M4 M5 M2 M5 M4 M5 M4 M4 M4 M4 M5 M5 M4 M2 M3 M3 M5 M4 M6 M5 M3 M4 M5 M1 M4 M3 M4 M4 M1 M1 M6 M2 M5 M5 M5 M0 M4 M4 M1 M5 M4 M5 M4 M4 M5 M3 M3 M2 M6 M5 M1 M1 M5 M4 M5 M4 M1 M1 M3 M4 M1 M5 M4 M2 M8 M0 M5 M5 M4 yes yes yes yes yes yes yes yes yes yes yes Robo-AO Kepler Planetary Candidate Survey III 27 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? yes KOI K02076 K02078 K02080 K02091 K02092 K02093 K02094 K02101 K02103 K02104 K02106 K02107 K02108 K02113 K02116 K02117 K02121 K02125 K02126 K02128 K02129 K02134 K02140 K02146 K02150 K02152 K02154 K02163 K02164 K02166 K02167 K02171 K02177 K02179 K02180 K02182 K02183 K02188 K02193 K02195 K02198 K02199 K02200 K02205 K02206 K02210 K02212 K02213 K02216 K02217 K02221 K02223 K02227 K02229 K02236 K02237 K02242 K02245 K02248 K02250 K02255 K02257 K02271 K02274 K02283 K02285 K02291 K02294 K02307 K02308 K02309 K02313 K02316 K02323 K02327 K02328 K02329 mi (mags) 15.09 15.06 15.15 15.48 15.8 15.19 14.71 14.56 14.5 14.94 14.94 14.88 14.77 15.59 14.52 15.17 14.74 15.16 15.55 15.69 14.54 14.53 15.04 14.99 15.04 14.84 15.62 14.44 14.92 15.45 15.08 14.77 15.21 14.99 14.58 15.58 14.98 15.13 15.21 14.74 12.7 15.61 15.12 15.12 14.98 14.86 14.74 15.13 14.97 14.91 15.19 14.95 14.65 15.45 15.49 14.97 15.51 14.52 15.21 15.26 15.61 15.24 15.32 15.6 14.74 15.71 14.88 14.69 14.61 15.52 14.79 15.1 14.82 15.2 15.11 15.46 15.05 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low low high high low low low high high high high high low low high high high high 2014 Sep 02 LP600 2014 Aug 31 LP600 2014 Aug 31 LP600 medium 2014 Aug 27 LP600 2014 Aug 31 LP600 2014 Aug 27 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 21 LP600 2014 Jun 19 LP600 medium 2014 Sep 02 LP600 2014 Jul 13 LP600 2014 Jul 17 LP600 2014 Jul 16 LP600 medium 2014 Aug 31 LP600 2014 Aug 21 LP600 medium 2014 Nov 09 LP600 medium 2014 Jun 19 LP600 2014 Aug 28 LP600 medium 2014 Aug 26 LP600 2014 Aug 26 LP600 2014 Jun 19 LP600 medium 2014 Jul 14 LP600 medium 2014 Aug 28 LP600 medium 2014 Aug 31 LP600 medium 2014 Nov 08 LP600 medium 2014 Sep 02 LP600 2014 Sep 03 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 31 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 27 LP600 2014 Aug 21 LP600 2014 Sep 03 LP600 medium 2014 Jul 12 LP600 2014 Jun 19 LP600 2014 Sep 03 LP600 2014 Aug 31 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 24 LP600 2014 Sep 02 LP600 2014 Aug 29 LP600 2014 Sep 03 LP600 medium 2014 Sep 03 LP600 2014 Sep 01 LP600 medium 2014 Jul 19 LP600 2014 Jul 14 LP600 2014 Aug 29 LP600 2014 Aug 24 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 29 LP600 2014 Aug 31 LP600 medium 2014 Aug 21 LP600 2014 Aug 22 2014 Aug 24 LP600 medium 2014 Sep 02 LP600 2014 Jun 17 LP600 2014 Sep 01 LP600 2014 Jul 16 LP600 medium 2014 Sep 01 LP600 2014 Nov 09 LP600 2014 Aug 24 LP600 medium 2014 Aug 26 LP600 medium 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Sep 01 LP600 2014 Sep 02 LP600 2014 Sep 01 LP600 medium 2014 Aug 22 2014 Jul 14 LP600 medium 2014 Aug 26 LP600 2014 Jul 12 LP600 2014 Sep 02 LP600 2014 Jul 14 LP600 medium 2014 Aug 27 LP600 medium 2014 Sep 01 LP600 2014 Sep 01 LP600 2014 Sep 01 LP600 medium low high low low high high high low high low low high low high low high low high low low low low high low low i(cid:48) i(cid:48) yes yes yes yes yes yes yes M1 M6 M4 M3 M0 M2 M5 M6 M4 M1 M2 M5 M3 M3 M3 M5 M5 M3 M6 M5 M3 M4 M4 M5 M4 M2 M4 M4 M4 M4 M4 M5 M5 M8 M5 M5 M4 M4 M6 M4 M3 M5 M5 M5 M5 M4 M5 M5 M4 M3 M4 M4 M2 M4 M2 M4 M3 M4 M3 M6 M4 M4 M3 M5 M5 M1 M4 M1 M4 M2 M4 M1 M3 M3 M2 M7 KOI K01695 K01705 K01708 K01710 K01711 K01716 K01718 K01721 K01722 K01723 K01724 K01727 K01732 K01736 K01739 K01747 K01749 K01750 K01758 K01761 K01771 K01772 K01773 K01782 K01787 K01790 K01792 K01796 K01799 K01804 K01806 K01816 K01821 K01840 K01846 K01847 K01858 K01866 K01871 K01872 K01873 K01879 K01882 K01892 K01900 K01908 K01910 K01933 K01939 K01947 K01951 K01957 K01959 K01963 K01968 K01971 K01973 K01974 K01975 K01981 K01982 K01995 K01996 K01998 K02000 K02012 K02020 K02021 K02028 K02048 K02050 K02060 K02061 K02062 K02063 K02065 K02074 mi (mags) 13.58 15.26 14.55 15.21 14.72 14.55 15.02 14.77 15.31 15.33 15.39 15.42 15.5 15.69 14.92 14.56 15.55 14.57 15.72 15.44 15.76 15.63 15.4 15.32 15.49 15.04 11.92 12.74 15.29 15.29 13.34 15.13 14.68 14.71 15.52 14.53 14.53 14.78 14.48 15.59 15.47 13.42 14.48 15.26 14.29 14.25 14.41 14.82 14.9 12.17 13.42 12.52 13.91 13.93 14.92 15.38 15.32 14.69 15.77 13.93 15.49 15.05 15.27 15.19 15.34 14.67 15.07 15.16 15.9 15.49 12.22 14.86 15.43 14.79 15.38 14.09 15.52 low low low low low i(cid:48) i(cid:48) high low low low low medium high high high low low low low low low low high high 2014 Aug 31 LP600 2014 Aug 28 LP600 2014 Aug 21 LP600 medium 2014 Aug 27 LP600 2014 Aug 21 LP600 medium 2014 Jul 11 LP600 2014 Aug 27 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 28 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 24 LP600 2014 Aug 26 LP600 2014 Sep 01 LP600 2014 Aug 22 2014 Jul 18 LP600 2014 Aug 26 LP600 2014 Jul 16 LP600 2014 Aug 24 LP600 2014 Aug 24 LP600 medium 2014 Sep 03 LP600 2015 Jun 04 LP600 2014 Aug 24 LP600 medium 2014 Aug 27 LP600 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 medium 2014 Sep 02 LP600 2014 Aug 21 2014 Aug 23 LP600 2014 Sep 02 LP600 2014 Jul 13 LP600 medium 2014 Aug 28 LP600 2014 Jul 17 LP600 medium 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 2014 Jul 18 LP600 medium 2014 Aug 22 2014 Sep 02 LP600 2014 Aug 22 2014 Sep 02 LP600 2014 Aug 24 LP600 2014 Aug 28 LP600 2014 Jun 19 LP600 medium 2014 Aug 31 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 22 2014 Jul 19 LP600 medium 2014 Jul 16 LP600 medium 2014 Jun 19 LP600 2012 Aug 30 2014 Jul 16 LP600 medium 2014 Jul 16 LP600 2014 Aug 22 2014 Aug 29 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Jun 19 LP600 medium 2014 Aug 24 LP600 2014 Jun 19 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 24 LP600 2014 Aug 26 LP600 2014 Aug 28 LP600 2014 Aug 26 LP600 medium 2014 Jul 17 LP600 2014 Aug 20 2014 Aug 26 LP600 2014 Aug 31 LP600 2014 Aug 28 LP600 2015 Jun 07 LP600 2014 Sep 01 LP600 medium 2014 Sep 02 LP600 2014 Sep 01 LP600 medium 2014 Aug 31 LP600 2014 Jul 14 LP600 medium 2014 Aug 26 LP600 low low high low low high low high low high high low high high high low medium low low low low high low low low low i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) M4 M4 M4 M3 M5 M4 M3 M4 M4 M4 M1 M2 M2 M1 M2 M2 M5 M5 M3 M1 M4 M5 M3 M4 M4 M5 M4 M1 M4 M3 M3 M4 M3 M3 M5 M2 M1 M5 M3 M4 M6 M4 M4 M6 M4 M5 M4 M5 M4 M3 M4 M4 M4 M3 M4 M4 M4 M1 M5 M4 M5 M6 M0 M4 M2 M4 M6 M2 M4 M5 M5 M3 M4 M2 M4 M5 yes yes yes yes yes yes yes yes TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) medium 28 KOI K02338 K02339 K02340 K02345 K02346 K02348 K02350 K02353 K02355 K02357 K02361 K02362 K02364 K02369 K02372 K02376 K02379 K02383 K02387 K02396 K02397 K02401 K02404 K02406 K02409 K02416 K02420 K02424 K02432 K02436 K02439 K02442 K02445 K02449 K02450 K02451 K02453 K02458 K02460 K02461 K02466 K02472 K02473 K02477 K02480 K02482 K02485 K02487 K02491 K02494 K02497 K02506 K02509 K02512 K02513 K02520 K02521 K02524 K02525 K02528 K02532 K02535 K02543 K02544 K02548 K02553 K02554 K02578 K02579 K02580 K02586 K02589 K02592 K02594 K02602 K02604 K02610 mi (mags) 14.65 14.72 14.53 14.53 15.43 15.09 14.84 14.7 15.6 15.39 14.85 15.6 15.53 15.63 13.23 14.98 14.86 14.93 15.11 14.41 15.28 14.45 15.41 14.53 14.58 15.25 14.53 15.59 15.69 15.46 15.25 15.33 15.56 14.89 14.6 14.47 14.9 15.13 14.6 15.66 14.57 15.42 15.48 14.52 15.19 14.76 15.01 15.05 15.34 14.63 15.29 14.98 14.76 15.28 14.67 15.22 15.57 15.1 15.32 15.44 15.12 14.64 15.04 15.59 14.67 15.34 15.4 14.65 14.95 15.53 15.31 15.33 14.87 14.96 14.85 14.99 12.61 Ziegler et al. TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) i(cid:48) low high low low low low high low low low low low low low high low high high 2014 Sep 01 LP600 medium 2014 Aug 22 2014 Aug 21 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 medium 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 medium 2014 Aug 31 LP600 2014 Aug 31 LP600 2014 Jul 18 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Nov 09 LP600 medium 2014 Aug 22 medium 2014 Aug 21 LP600 2014 Aug 29 LP600 2014 Jul 19 LP600 medium 2014 Sep 02 LP600 medium 2014 Aug 21 LP600 2014 Aug 28 LP600 2014 Aug 22 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 2014 Jul 17 LP600 2014 Aug 31 LP600 2014 Jun 19 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 27 LP600 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Sep 02 LP600 2014 Jun 19 LP600 medium 2014 Jul 16 LP600 medium 2012 Jul 17 LP600 2014 Aug 28 LP600 2014 Aug 29 LP600 medium 2014 Aug 28 LP600 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Jul 16 LP600 medium 2014 Aug 29 LP600 2014 Aug 24 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 24 LP600 medium 2014 Jul 14 LP600 2014 Sep 03 LP600 medium 2014 Jul 17 LP600 medium 2014 Jul 18 LP600 2014 Sep 01 LP600 medium 2014 Jul 18 LP600 medium 2014 Aug 28 LP600 2014 Aug 31 LP600 2014 Aug 23 LP600 medium 2014 Aug 22 2014 Sep 01 LP600 medium 2014 Aug 26 LP600 medium 2014 Aug 23 LP600 2014 Aug 28 LP600 2014 Sep 01 LP600 medium 2014 Aug 29 LP600 2014 Sep 01 LP600 medium 2014 Sep 01 LP600 2014 Jul 14 LP600 medium 2014 Jul 12 LP600 medium 2014 Aug 31 LP600 2014 Sep 01 LP600 medium 2014 Sep 01 LP600 2014 Jul 17 LP600 2014 Nov 09 LP600 medium 2014 Aug 29 LP600 medium 2014 Aug 21 LP600 medium 2015 Jun 04 LP600 low high high low low high high low low low high high high high low low low low low M4 M6 M4 M3 M2 M4 M5 M5 M4 M1 M4 M3 M3 M4 M3 M5 M0 M4 M3 M3 M3 M4 M4 M0 M3 M1 M4 M5 M4 M4 M4 M3 M5 M5 M4 M5 M6 M5 M5 M3 M5 M1 M3 M4 M5 M4 M5 M3 M4 M5 M5 M4 M3 M4 M3 M4 M5 M3 M4 M4 M5 M6 M4 M4 M6 M5 M6 M4 M3 M3 M4 M5 M0 M4 M4 M5 M4 yes yes yes yes yes yes yes yes yes KOI K02612 K02613 K02615 K02619 K02624 K02625 K02628 K02637 K02638 K02639 K02647 K02655 K02660 K02668 K02683 K02688 K02689 K02691 K02694 K02699 K02703 K02708 K02715 K02716 K02719 K02721 K02728 K02735 K02736 K02739 K02742 K02744 K02745 K02747 K02751 K02758 K02760 K02762 K02763 K02764 K02768 K02770 K02780 K02791 K02793 K02796 K02797 K02802 K02806 K02813 K02816 K02817 K02820 K02821 K02828 K02832 K02834 K02835 K02841 K02845 K02851 K02852 K02853 K02856 K02862 K02865 K02871 K02874 K02875 K02876 K02877 K02878 K02886 K02894 K02896 K02898 K02899 mi (mags) 11.6 15.67 15.5 15.69 14.97 15.63 14.63 14.82 15.26 15.33 14.82 15.03 15.21 13.94 15.27 16.05 15.35 14.59 14.58 14.93 14.72 15.51 16.33 15.57 14.75 14.93 12.74 15.28 15.45 14.78 14.51 14.93 15.03 15.0 15.44 11.89 14.48 14.61 15.02 15.22 14.99 15.11 15.34 14.54 15.64 14.64 15.59 14.75 15.12 13.31 15.22 15.46 15.07 15.05 15.37 13.63 15.46 15.47 15.54 15.0 15.2 15.65 14.64 15.11 15.29 15.15 14.52 14.62 14.53 14.62 14.49 14.45 15.61 15.38 11.9 15.35 15.63 i(cid:48) i(cid:48) high high low high low low high low low low high low low high high low high high low 2014 Aug 20 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 24 LP600 2014 Aug 21 LP600 2014 Sep 01 LP600 2014 Sep 02 LP600 medium 2014 Aug 23 LP600 medium 2014 Aug 31 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 21 LP600 2014 Aug 20 2014 Aug 28 LP600 2014 Jul 14 LP600 medium 2014 Aug 26 LP600 medium 2014 Aug 31 LP600 2014 Sep 02 LP600 2014 Jul 14 LP600 2014 Jul 17 LP600 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 2014 Aug 26 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Jul 16 LP600 medium 2014 Jul 16 LP600 medium 2014 Jul 14 LP600 medium 2014 Aug 28 LP600 2014 Sep 01 LP600 2014 Jul 16 LP600 medium 2014 Jul 17 LP600 medium 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 medium 2014 Jul 16 LP600 medium 2014 Sep 02 LP600 2015 Jun 07 LP600 2014 Aug 23 LP600 medium 2014 Aug 21 2014 Sep 01 LP600 medium 2014 Sep 01 LP600 2014 Aug 31 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 21 LP600 2014 Sep 02 LP600 2014 Jul 17 LP600 medium 2014 Aug 28 LP600 2014 Jul 17 LP600 medium 2014 Aug 24 LP600 2013 Aug 15 LP600 medium 2014 Sep 03 LP600 medium 2014 Nov 07 LP600 2014 Sep 01 LP600 medium 2014 Aug 26 LP600 2014 Aug 26 LP600 2014 Jul 16 LP600 medium 2014 Nov 09 LP600 medium 2014 Aug 26 LP600 medium 2014 Aug 30 LP600 2014 Sep 02 LP600 2014 Aug 26 LP600 2014 Aug 24 LP600 2014 Sep 02 LP600 medium 2014 Aug 26 LP600 2014 Aug 27 LP600 2014 Aug 31 LP600 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 2014 Jul 16 LP600 2014 Jul 17 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 26 LP600 2014 Aug 27 LP600 2015 Jun 05 LP600 2014 Sep 03 LP600 medium 2014 Aug 24 LP600 low low high high low high low high low high high high low low low low high high high low low low M4 M5 M1 M5 M4 M3 M3 M4 M4 M3 M3 M2 M0 M4 M4 M6 M3 M4 M6 M5 M6 M5 M6 M2 M5 M4 M2 M5 M4 M4 M6 M4 M4 M5 M4 M5 M5 M4 M5 M7 M4 M5 M3 M1 M5 M4 M1 M3 M6 M5 M4 M3 M4 M5 M6 M5 M5 M5 M3 M5 M5 M5 M5 M1 M8 M1 M4 M4 M3 M4 M4 M5 M5 M2 M4 M3 M5 yes yes yes yes yes yes yes yes yes Robo-AO Kepler Planetary Candidate Survey III 29 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? yes high KOI K02900 K02916 K02919 K02920 K02921 K02925 K02926 K02927 K02931 K02933 K02945 K02946 K02950 K02957 K02958 K02975 K02976 K02980 K02981 K02995 K02996 K03009 K03010 K03013 K03014 K03020 K03022 K03027 K03031 K03034 K03037 K03042 K03043 K03045 K03051 K03063 K03066 K03072 K03077 K03078 K03088 K03089 K03090 K03091 K03094 K03095 K03096 K03102 K03104 K03106 K03110 K03111 K03112 K03113 K03115 K03116 K03120 K03131 K03136 K03137 K03138 K03141 K03144 K03161 K03165 K03168 K03202 K03214 K03224 K03225 K03248 K03258 K03262 K03263 K03264 K03266 K03267 mi (mags) 14.95 14.11 14.57 14.38 15.18 13.22 15.69 15.65 14.4 15.4 15.78 15.47 14.18 14.15 14.59 15.1 15.6 15.29 15.42 15.42 15.41 14.87 15.17 14.48 15.63 13.36 15.17 15.16 15.03 14.98 15.61 15.75 14.64 15.56 15.08 15.51 15.85 15.27 15.57 15.14 14.75 14.96 15.07 15.21 15.13 14.86 13.35 15.4 15.13 15.17 14.7 12.66 15.64 14.89 15.06 15.2 14.64 15.36 15.4 15.42 16.69 14.62 15.38 9.58 10.22 10.29 11.53 11.81 11.96 11.97 12.24 15.45 14.87 15.27 15.6 15.04 14.53 i(cid:48) i(cid:48) low low high high high high low low low low low low high low low low low high low high high low low 2014 Sep 03 LP600 2014 Jul 17 LP600 medium 2014 Sep 02 LP600 medium 2014 Jul 14 LP600 medium 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 22 2014 Aug 28 LP600 2014 Aug 29 LP600 2014 Sep 02 LP600 2014 Jul 17 LP600 medium 2014 Jul 16 LP600 medium 2014 Sep 02 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Aug 31 LP600 2014 Aug 21 LP600 2014 Aug 22 2014 Jul 13 LP600 2014 Aug 24 LP600 2013 Aug 13 LP600 2014 Aug 27 LP600 2014 Sep 01 LP600 2014 Aug 27 LP600 2014 Jul 16 LP600 medium 2014 Nov 09 LP600 medium 2014 Aug 31 LP600 2014 Jul 12 LP600 2014 Nov 08 LP600 medium 2014 Aug 31 LP600 2014 Aug 28 LP600 medium 2014 Aug 24 LP600 2014 Aug 23 LP600 2014 Aug 31 LP600 2014 Aug 28 LP600 2014 Jul 12 LP600 2014 Jul 17 LP600 medium 2014 Aug 27 LP600 2014 Sep 02 LP600 2014 Sep 02 LP600 medium 2014 Aug 21 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 23 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Jul 18 LP600 2014 Aug 20 2014 Sep 01 LP600 2014 Nov 07 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 29 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Aug 31 LP600 2014 Jul 17 LP600 medium 2014 Nov 09 LP600 medium 2015 Jun 03 LP600 2014 Aug 22 2014 Aug 22 2014 Aug 22 2014 Aug 29 LP600 2014 Aug 23 2014 Aug 22 2014 Aug 22 2014 Aug 26 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 23 LP600 2014 Aug 28 LP600 2014 Aug 26 LP600 2014 Aug 21 LP600 high low low high low low high low low low low low high high high high high high high high low low high high low low low low high i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) high low i(cid:48) M5 M5 M4 M4 M5 M5 M7 M4 M3 M4 M2 M1 M4 M5 M3 M4 M3 M1 M5 M0 M3 M3 M5 M5 M1 M4 M5 M3 M2 M6 M4 M1 M5 M3 M1 M2 M3 M3 M6 M4 M1 M5 M7 M0 M7 M4 M4 M7 M4 M5 M0 M1 M5 M4 M1 M4 M1 M2 M4 M3 M4 M8 M3 M4 M5 M5 M4 M5 M5 M5 M4 M3 M6 M4 M4 M4 KOI K03268 K03271 K03274 K03279 K03280 K03281 K03282 K03283 K03287 K03298 K03303 K03305 K03306 K03307 K03312 K03313 K03316 K03319 K03320 K03323 K03329 K03330 K03331 K03335 K03337 K03341 K03342 K03344 K03347 K03348 K03349 K03352 K03354 K03355 K03356 K03357 K03358 K03361 K03367 K03370 K03372 K03375 K03383 K03386 K03391 K03393 K03394 K03395 K03396 K03397 K03404 K03407 K03410 K03411 K03412 K03413 K03415 K03416 K03417 K03418 K03420 K03421 K03426 K03428 K03432 K03433 K03434 K03436 K03443 K03445 K03449 K03455 K03458 K03463 K03465 K03467 K03470 mi (mags) 15.33 15.46 15.11 15.56 15.37 14.9 15.27 15.32 13.88 15.49 15.35 15.58 15.54 15.35 15.54 14.41 15.26 14.8 15.62 14.84 15.62 14.72 15.69 15.63 15.66 14.69 15.53 15.04 15.23 15.02 15.11 15.19 14.86 14.93 15.24 14.96 13.79 15.28 14.87 14.43 15.24 15.45 15.59 14.94 15.45 15.13 15.3 14.42 15.33 15.2 14.64 15.2 15.37 14.43 14.13 15.02 13.08 15.11 14.99 15.2 15.38 13.79 14.13 14.52 14.76 15.11 14.76 13.2 15.52 15.42 14.55 15.57 15.44 14.56 12.89 15.0 15.2 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) low low high low low high high high low high low low high high high high low high high low high low 2014 Sep 03 LP600 medium 2014 Aug 31 LP600 2014 Aug 24 LP600 medium 2014 Aug 26 LP600 2014 Sep 03 LP600 medium 2015 Jun 04 LP600 2014 Aug 31 LP600 medium 2015 Jun 08 LP600 2014 Aug 20 2014 Aug 26 LP600 2014 Aug 26 LP600 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Aug 26 LP600 2014 Sep 03 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 26 LP600 2014 Aug 31 LP600 medium 2014 Sep 01 LP600 2014 Aug 31 LP600 medium 2014 Aug 31 LP600 2014 Jul 16 LP600 medium 2014 Aug 31 LP600 2014 Sep 01 LP600 2014 Sep 03 LP600 2014 Jul 17 LP600 medium 2014 Aug 29 LP600 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Aug 31 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 24 LP600 2014 Jul 16 LP600 2014 Jul 18 LP600 2014 Aug 26 LP600 medium 2014 Sep 01 LP600 2014 Jul 16 LP600 medium 2014 Aug 24 LP600 medium 2014 Jun 19 LP600 medium 2014 Jul 13 LP600 medium 2014 Aug 23 LP600 2014 Sep 01 LP600 medium 2014 Aug 24 LP600 2014 Jul 17 LP600 medium 2014 Aug 28 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 medium 2014 Jul 14 LP600 medium 2014 Aug 26 LP600 2014 Aug 28 LP600 medium 2014 Jul 14 LP600 2014 Aug 24 LP600 medium 2014 Aug 28 LP600 2015 Jun 03 LP600 2014 Jul 17 LP600 medium 2014 Aug 26 LP600 medium 2013 Jul 27 LP600 2014 Nov 07 LP600 2014 Jul 14 LP600 2014 Aug 23 LP600 2014 Aug 26 LP600 2014 Aug 22 2014 Jul 14 LP600 2014 Aug 20 2014 Jul 16 LP600 2014 Aug 21 LP600 2014 Jul 18 LP600 2014 Jul 14 LP600 2014 Aug 31 LP600 2014 Sep 02 LP600 2014 Aug 21 LP600 medium 2014 Aug 27 LP600 2014 Aug 28 LP600 2015 Jun 07 LP600 medium 2014 Aug 29 LP600 2014 Jun 19 LP600 medium 2014 Sep 01 LP600 medium high low high low high high low low high low high low low low high high high medium low low high high high low low i(cid:48) i(cid:48) M4 M3 M4 M5 M4 M3 M6 M1 M6 M4 M4 M5 M3 M4 M5 M5 M4 M3 M4 M0 M5 M5 M0 M3 M3 M4 M1 M5 M4 M3 M5 M5 M4 M4 M5 M4 M4 M3 M5 M4 M4 M5 M3 M4 M4 M3 M4 M2 M2 M1 M4 M1 M1 M5 M3 M5 M3 M5 M1 M1 M2 M4 M1 M5 M3 M4 M6 M3 M2 M5 M5 M4 M4 M5 M3 M4 yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes yes Ziegler et al. TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? yes high 30 KOI K03471 K03472 K03473 K03478 K03480 K03482 K03483 K03487 K03493 K03495 K03504 K03515 K03520 K03522 K03527 K03531 K03533 K03545 K03554 K03560 K03565 K03581 K03605 K03606 K03611 K03617 K03620 K03626 K03641 K03647 K03649 K03652 K03660 K03674 K03675 K03678 K03685 K03690 K03692 K03696 K03709 K03716 K03717 K03720 K03726 K03738 K03741 K03749 K03765 K03767 K03770 K03771 K03783 K03784 K03791 K03794 K03801 K03815 K03818 K03823 K03830 K03853 K03855 K03867 K03871 K03878 K03880 K03886 K03890 K03901 K03909 K03913 K03923 K03928 K03933 K03936 K03939 mi (mags) 12.96 14.25 13.53 13.35 15.7 14.9 14.7 14.09 15.02 14.92 15.72 12.13 13.67 10.66 13.85 14.42 14.44 15.47 14.99 11.69 15.64 16.35 13.87 14.05 16.3 14.61 15.05 16.21 13.76 16.44 15.21 15.58 15.31 15.4 16.28 12.59 14.42 15.33 14.92 15.47 15.07 11.61 14.79 13.26 15.4 17.17 13.15 15.64 16.02 17.01 13.93 16.34 12.84 14.02 13.61 15.26 15.75 15.23 11.91 13.86 16.77 10.32 18.23 14.64 11.94 12.72 10.76 9.46 12.87 15.09 14.78 14.8 15.11 13.06 14.45 13.03 15.02 i(cid:48) low high high low high low high high high low low high high high high low high high low high high high 2014 Jul 11 LP600 2014 Jun 19 LP600 medium 2014 Jun 15 LP600 medium 2015 Jun 10 LP600 medium 2014 Sep 03 LP600 2014 Jun 19 LP600 medium 2014 Nov 09 LP600 medium 2014 Jul 17 LP600 2014 Sep 01 LP600 medium 2014 Jun 19 LP600 medium 2015 Jun 06 LP600 2014 Jun 13 LP600 2015 Jun 04 LP600 medium 2015 Jun 12 LP600 2014 Jun 15 LP600 medium 2014 Jul 18 LP600 2014 Nov 09 LP600 2014 Aug 24 LP600 2014 Aug 22 2014 Jul 16 LP600 2014 Aug 28 LP600 2014 Jul 18 LP600 2014 Jul 16 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 26 LP600 2014 Aug 21 LP600 2014 Sep 03 LP600 2014 Sep 03 LP600 2014 Jul 14 LP600 2014 Aug 23 LP600 medium 2014 Aug 23 LP600 2014 Aug 29 LP600 2014 Aug 24 LP600 2014 Sep 03 LP600 medium 2014 Aug 31 LP600 2014 Jun 17 LP600 2014 Aug 21 LP600 medium 2014 Aug 31 LP600 medium 2014 Jul 14 LP600 2014 Aug 26 LP600 medium 2014 Aug 26 LP600 2015 Jun 05 LP600 2014 Sep 02 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 29 LP600 2014 Aug 23 LP600 2014 Jun 13 LP600 medium 2014 Aug 28 LP600 2014 Sep 02 LP600 2014 Aug 21 LP600 2014 Jun 19 LP600 medium 2014 Aug 23 LP600 2014 Aug 21 LP600 2014 Jul 16 LP600 medium 2014 Aug 22 2014 Sep 02 LP600 medium 2014 Aug 20 2014 Aug 26 LP600 medium 2014 Aug 22 2014 Jul 17 LP600 medium 2014 Aug 26 LP600 2014 Aug 22 2014 Nov 09 LP600 2014 Jul 17 LP600 medium 2014 Aug 31 LP600 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2014 Aug 20 2014 Aug 20 2014 Sep 03 LP600 medium 2014 Aug 31 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 26 LP600 medium 2014 Jul 14 LP600 2014 Jul 12 LP600 2014 Jun 13 LP600 medium 2014 Aug 28 LP600 high high high high high low low low high high high high high high high low high high low i(cid:48) i(cid:48) i(cid:48) high high low low i(cid:48) i(cid:48) i(cid:48) KOI K03975 K04002 K04007 K04009 K04011 K04015 K04022 K04024 K04036 K04062 K04076 K04084 K04087 K04091 K04099 K04103 K04117 K04120 K04121 K04126 K04127 K04131 K04135 K04139 K04178 K04185 K04189 K04192 K04202 K04204 K04207 K04222 K04232 K04234 K04246 K04255 K04257 K04259 K04261 K04267 K04268 K04290 K04294 K04298 K04307 K04323 K04333 K04334 K04343 K04345 K04346 K04351 K04353 K04357 K04360 K04361 K04363 K04364 K04366 K04371 K04372 K04377 K04384 K04385 K04386 K04391 K04398 K04405 K04410 K04418 K04421 K04424 K04426 K04442 K04444 K04447 K04450 mi (mags) 14.71 14.75 15.28 15.06 12.37 15.02 13.1 13.53 13.7 13.91 15.05 14.9 14.54 14.71 15.05 14.84 14.69 14.93 15.46 15.17 15.0 13.17 14.49 15.15 14.4 15.29 15.37 14.77 15.45 13.8 14.88 13.52 14.98 14.97 13.16 14.68 13.1 14.92 15.17 15.0 14.77 16.58 14.66 13.37 14.04 13.4 14.01 15.49 13.49 13.19 14.67 14.71 15.36 14.75 14.62 14.93 15.71 15.17 15.29 15.19 15.14 13.34 14.74 15.51 15.51 14.75 15.24 14.49 15.54 15.71 12.63 15.14 13.49 14.11 15.08 14.44 14.89 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) high low low low low high high high high low high high low high low high high low high low low low low 2014 Aug 21 LP600 medium 2014 Jul 14 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 26 LP600 medium 2014 Jul 13 LP600 2014 Aug 24 LP600 medium 2014 Jul 14 LP600 2014 Jul 17 LP600 2014 Aug 20 2014 Aug 29 LP600 medium 2014 Sep 02 LP600 2014 Jun 15 LP600 2014 Aug 22 2014 Jul 14 LP600 2014 Aug 31 LP600 2015 Jun 08 LP600 2014 Aug 31 LP600 medium 2014 Jul 19 LP600 2014 Aug 31 LP600 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Jun 19 LP600 2014 Aug 20 2014 Aug 28 LP600 medium 2015 Jun 12 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Aug 29 LP600 2014 Aug 26 LP600 medium 2014 Aug 29 LP600 2014 Jun 19 LP600 2014 Jun 13 LP600 medium 2014 Aug 21 LP600 2014 Aug 21 LP600 2014 Jul 11 LP600 medium 2014 Aug 21 LP600 medium 2014 Jul 11 LP600 2014 Jul 17 LP600 2014 Sep 02 LP600 2014 Jun 19 LP600 medium 2014 Aug 31 LP600 2015 Jun 10 LP600 2014 Sep 01 LP600 medium 2014 Sep 03 LP600 2014 Aug 20 2014 Jun 13 LP600 2014 Aug 20 2014 Sep 01 LP600 2014 Jun 19 LP600 medium 2014 Jul 13 LP600 medium 2014 Jul 16 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 24 LP600 2014 Jul 16 LP600 2014 Jul 18 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 28 LP600 2014 Aug 26 LP600 2014 Aug 28 LP600 2014 Sep 03 LP600 medium 2014 Sep 02 LP600 2014 Jul 16 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 28 LP600 2014 Aug 26 LP600 2014 Jul 17 LP600 medium 2014 Aug 31 LP600 2014 Jul 17 LP600 medium 2014 Aug 31 LP600 2014 Sep 03 LP600 2014 Jul 12 LP600 2014 Aug 24 LP600 medium 2014 Aug 20 2014 Jul 17 LP600 medium 2014 Aug 28 LP600 2014 Aug 21 2014 Aug 31 LP600 medium high low high low low high low high low low high high low low high low high low low low low low high i(cid:48) i(cid:48) low low i(cid:48) i(cid:48) M5 M5 M3 M3 M4 M5 M4 M5 M4 M3 M2 M5 M5 M1 M3 M3 M5 M0 M3 M3 M5 M2 M4 M3 M4 M3 M4 M5 M4 M5 M6 M4 M1 M1 M4 M5 M5 M3 M1 M3 M6 M9 M3 M3 M1 M4 M1 M4 M4 M4 M4 M5 M1 M3 M3 M4 M5 M3 M5 M3 M4 M3 M4 M5 M3 M4 M5 M3 M1 M2 M4 M3 M0 M4 M4 M4 M4 yes yes yes yes yes yes yes yes yes yes yes yes yes M6 M3 M5 M5 M3 M4 M3 M5 M4 M4 M2 M4 M4 M0 M4 M1 M4 M6 M2 M4 M5 M6 M5 M2 M1 M2 M5 M5 M5 M3 M4 M5 M2 M4 M3 M5 M4 M3 M4 M4 M4 M3 M1 M6 M5 M3 M8 M4 M5 M3 M5 M3 M4 M0 M5 M2 M4 K7 M4 M5 M5 M2 M4 M5 M4 M4 M6 M6 M4 M3 M3 M6 M4 M3 M5 M0 yes yes yes yes yes yes yes yes yes yes yes yes yes Robo-AO Kepler Planetary Candidate Survey III 31 KOI K04451 K04452 K04462 K04464 K04466 K04467 K04473 K04477 K04480 K04484 K04487 K04491 K04496 K04497 K04498 K04499 K04500 K04503 K04504 K04506 K04508 K04509 K04513 K04519 K04524 K04526 K04546 K04548 K04549 K04550 K04551 K04557 K04571 K04577 K04583 K04585 K04587 K04590 K04595 K04597 K04602 K04622 K04626 K04630 K04632 K04636 K04640 K04643 K04647 K04649 K04650 K04653 K04655 K04659 K04661 K04666 K04667 K04676 K04686 K04694 K04698 K04700 K04705 K04710 K04711 K04713 K04714 K04715 K04716 K04717 K04725 K04730 K04733 K04735 K04742 K04743 K04745 mi (mags) 15.63 13.62 11.1 14.7 14.84 15.63 13.66 15.23 13.9 14.91 14.1 14.82 15.38 14.97 14.64 14.46 15.55 15.47 15.69 15.42 14.96 15.2 14.52 14.6 14.74 15.06 13.31 14.5 15.73 15.05 14.46 13.51 14.35 14.7 15.53 13.37 15.36 15.53 15.2 14.46 14.41 14.61 14.66 14.66 15.34 15.23 13.48 15.09 13.43 15.2 15.46 13.38 15.18 13.47 14.49 15.11 15.24 13.68 12.09 14.68 14.61 15.7 14.83 15.38 15.41 13.6 15.01 13.77 12.17 15.33 13.01 14.04 15.63 14.4 14.56 14.71 15.54 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) high high high high low low low high high high low high high high low high high 2014 Sep 01 LP600 2014 Jul 16 LP600 2014 Aug 20 2014 Aug 21 LP600 medium 2014 Jun 19 LP600 medium 2014 Aug 26 LP600 2014 Jul 17 LP600 medium 2014 Aug 27 LP600 medium 2014 Aug 29 LP600 medium 2014 Jul 11 LP600 2014 Aug 21 LP600 2014 Jun 17 LP600 2014 Nov 10 LP600 medium 2014 Sep 03 LP600 2014 Jul 16 LP600 2014 Jul 16 LP600 medium 2014 Aug 24 LP600 2014 Aug 28 LP600 2014 Aug 28 LP600 2014 Nov 09 LP600 medium 2014 Aug 31 LP600 medium 2014 Sep 01 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 29 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 24 LP600 2014 Jul 14 LP600 medium 2014 Jul 14 LP600 2014 Aug 27 LP600 2014 Aug 29 LP600 2014 Jul 18 LP600 medium 2014 Jul 16 LP600 2014 Jul 16 LP600 medium 2014 Aug 31 LP600 medium 2014 Sep 01 LP600 medium 2014 Jul 19 LP600 medium 2014 Nov 09 LP600 medium 2014 Sep 02 LP600 2014 Aug 28 LP600 2014 Jul 17 LP600 medium 2014 Aug 29 LP600 medium 2014 Aug 23 LP600 2014 Aug 31 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 21 LP600 2014 Aug 27 LP600 2014 Jun 19 LP600 2014 Sep 02 LP600 medium 2014 Jul 13 LP600 2014 Aug 26 LP600 medium 2014 Sep 02 LP600 2014 Jul 19 LP600 2014 Aug 23 LP600 2014 Aug 20 2014 Jul 18 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 28 LP600 2014 Jul 14 LP600 medium 2014 Aug 20 2014 Aug 21 LP600 medium 2014 Sep 02 LP600 medium 2014 Aug 31 LP600 2014 Jul 14 LP600 2014 Sep 01 LP600 medium 2014 Aug 31 LP600 2014 Jul 16 LP600 medium 2014 Sep 02 LP600 medium medium 2014 Aug 21 2014 Aug 22 medium 2014 Sep 03 LP600 medium 2014 Jul 17 LP600 2014 Aug 21 LP600 medium 2014 Sep 01 LP600 2014 Jul 17 LP600 medium 2014 Aug 22 2014 Sep 03 LP600 medium 2014 Aug 22 high high low low high high high high high high low low low high high high high high i(cid:48) i(cid:48) low low i(cid:48) i(cid:48) i(cid:48) i(cid:48) M4 M1 M2 M5 M3 M5 M4 M4 M4 M4 M5 M0 M4 M5 M1 M4 M5 M5 M5 M5 M5 M6 M4 M3 M4 M4 M3 M3 M5 M6 M4 M4 M4 M4 M4 M2 M4 M3 M4 M4 M4 M6 M4 M4 M5 M5 M4 M4 M5 M4 M6 M4 M1 M1 M5 M4 M4 M5 M4 M3 M2 M5 M2 M4 M3 M4 M5 M4 M3 M4 M6 M5 M6 M5 M4 M4 M6 yes yes yes yes yes yes yes yes yes yes yes yes yes KOI K04749 K04750 K04755 K04756 K04758 K04759 K04762 K04765 K04766 K04771 K04774 K04780 K04790 K04794 K04799 K04803 K04804 K04810 K04815 K04820 K04827 K04837 K04838 K04840 K04848 K04857 K04859 K04862 K04864 K04878 K04881 K04885 K04886 K04890 K04893 K04895 K04896 K04900 K04902 K04912 K04913 K04922 K04923 K04926 K04927 K04928 K04933 K04936 K04938 K04939 K04946 K04950 K04958 K04959 K04960 K04961 K04962 K04967 K04968 K04971 K04972 K04974 K04975 K04976 K04977 K04978 K04980 K04982 K04985 K04986 K04988 K04991 K04992 K04993 K04997 K04999 K05002 mi (mags) 15.37 15.67 14.67 15.23 15.69 14.77 15.25 13.24 14.93 13.85 13.5 15.16 14.25 14.09 14.01 14.71 15.0 15.04 14.76 15.27 15.15 15.1 14.59 14.44 15.14 15.17 15.19 15.23 13.02 12.12 12.69 14.77 15.6 15.61 15.45 14.52 15.25 14.94 15.15 14.69 13.23 13.4 13.02 15.57 14.56 14.97 14.92 12.54 15.37 15.97 16.84 15.92 12.1 13.71 14.22 14.94 14.05 15.44 11.1 15.17 13.83 15.51 15.47 13.41 15.12 15.01 11.15 14.19 12.53 14.59 13.66 15.06 14.9 12.47 15.32 15.28 12.64 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) low low low low low high high high low low high low high high low 2014 Aug 28 LP600 medium 2014 Aug 29 LP600 2014 Jul 16 LP600 medium 2014 Aug 27 LP600 2014 Sep 01 LP600 2014 Jul 19 LP600 medium 2014 Sep 01 LP600 medium 2014 Jul 17 LP600 2014 Nov 08 LP600 medium 2014 Aug 22 medium 2014 Aug 20 2014 Aug 28 LP600 medium 2014 Jun 13 LP600 medium 2014 Jul 19 LP600 medium 2014 Jul 14 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 26 LP600 2014 Aug 24 LP600 medium 2014 Jun 19 LP600 medium 2014 Aug 23 LP600 medium 2014 Aug 21 LP600 2014 Aug 23 LP600 2014 Jul 16 LP600 medium 2014 Jul 14 LP600 medium 2014 Aug 23 LP600 2014 Aug 24 LP600 medium 2014 Aug 23 LP600 2014 Sep 03 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 22 2014 Aug 21 LP600 medium 2014 Jul 16 LP600 medium 2014 Aug 24 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Aug 31 LP600 medium 2014 Sep 02 LP600 2014 Jul 14 LP600 medium 2015 Jun 12 LP600 2014 Sep 03 LP600 medium 2014 Jul 14 LP600 2014 Jul 14 LP600 medium 2014 Jul 14 LP600 medium 2015 Jun 04 LP600 2014 Jul 14 LP600 2014 Sep 03 LP600 medium 2014 Aug 30 LP600 2015 Jun 04 LP600 medium 2014 Aug 24 LP600 2014 Aug 23 LP600 2014 Nov 08 LP600 2014 Nov 08 LP600 medium 2014 Aug 20 2014 Jul 17 LP600 medium 2014 Jul 16 LP600 medium 2014 Sep 03 LP600 medium 2014 Jul 16 LP600 2014 Aug 26 LP600 medium 2014 Aug 30 LP600 2014 Aug 29 LP600 medium 2014 Jul 17 LP600 2014 Aug 26 LP600 2014 Aug 24 LP600 medium 2014 Jul 16 LP600 medium 2014 Nov 09 LP600 medium 2014 Aug 23 LP600 medium 2014 Jul 14 LP600 2015 Jun 04 LP600 medium 2014 Jul 16 LP600 2014 Aug 31 LP600 2014 Jun 19 LP600 medium 2014 Aug 26 LP600 2014 Jun 13 LP600 2014 Sep 01 LP600 2014 Aug 28 LP600 2014 Sep 03 LP600 medium 2014 Aug 22 medium high high high low high high low low high high high low low low high high high low low low i(cid:48) i(cid:48) yes yes yes yes yes yes yes yes M4 M5 M4 M4 M3 M4 M5 M4 M5 M2 M3 M4 M3 M5 M4 M4 M1 M4 M4 M4 M2 M3 M6 M4 M4 M4 M4 M5 M4 M4 M2 M4 M5 M4 M1 M4 M1 M4 M4 M3 M4 M3 M3 M4 M0 L0 M0 M4 M4 M6 M3 M4 M4 M2 M5 M1 M3 M6 M8 M4 M5 M4 M1 M5 M4 M4 M3 M3 M2 M4 M5 M5 M0 M4 M3 Ziegler et al. TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? yes low 32 KOI K05004 K05007 K05018 K05021 K05023 K05027 K05030 K05031 K05033 K05034 K05039 K05040 K05043 K05046 K05047 K05048 K05052 K05053 K05057 K05058 K05059 K05067 K05068 K05070 K05071 K05079 K05080 K05081 K05083 K05084 K05085 K05086 K05087 K05088 K05092 K05093 K05098 K05099 K05101 K05102 K05104 K05107 K05109 K05110 K05115 K05117 K05119 K05121 K05122 K05123 K05124 K05126 K05129 K05131 K05132 K05135 K05136 K05138 K05142 K05143 K05148 K05149 K05156 K05158 K05159 K05164 K05169 K05176 K05178 K05184 K05186 K05191 K05192 K05194 K05196 K05198 K05201 mi (mags) 14.26 13.47 14.92 14.97 13.46 13.17 14.45 11.97 13.83 13.25 15.58 14.38 15.08 12.86 14.7 13.42 12.53 13.81 12.19 13.31 13.03 14.25 12.95 9.9 15.66 13.05 14.61 14.08 13.81 15.76 15.94 10.95 12.3 15.53 13.89 12.69 14.86 12.71 12.89 14.85 15.53 13.74 14.09 11.65 15.77 15.71 12.64 12.41 16.16 14.03 15.77 13.64 11.6 15.22 12.78 15.28 12.2 14.08 15.51 15.67 15.1 15.65 12.07 11.87 14.73 12.37 15.17 13.17 14.28 15.48 13.29 14.34 14.87 13.56 12.25 15.43 15.0 i(cid:48) low high medium high high high low low high low low high low high low high high low low low high low low 2014 Jul 16 LP600 2014 Jul 16 LP600 medium 2014 Jun 15 LP600 2014 Sep 03 LP600 medium 2014 Jul 14 LP600 medium 2015 Jun 06 LP600 2015 Jun 03 LP600 2014 Aug 20 2015 Jun 03 LP600 2015 Jun 05 LP600 medium 2014 Sep 03 LP600 medium 2015 Jun 03 LP600 2014 Aug 31 LP600 2014 Jun 13 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2014 Jun 17 LP600 medium 2014 Jul 16 LP600 2014 Jul 14 LP600 2015 Jun 08 LP600 medium 2015 Jun 07 LP600 2015 Jun 03 LP600 2014 Jul 14 LP600 2014 Sep 01 LP600 2015 Jun 10 LP600 2014 Jun 17 LP600 medium 2014 Jul 17 LP600 2014 Jul 17 LP600 2015 Jun 12 LP600 medium 2015 Jun 08 LP600 2015 Jun 08 LP600 2015 Jun 03 LP600 2014 Aug 20 2015 Jun 03 LP600 2015 Jun 08 LP600 medium 2015 Jun 10 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 2014 Jul 17 LP600 2014 Aug 30 LP600 medium 2015 Jun 08 LP600 2015 Jun 07 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 03 LP600 medium 2015 Jun 10 LP600 2014 Aug 24 LP600 2015 Jun 06 LP600 2014 Jul 14 LP600 2014 Nov 09 LP600 2014 Jul 17 LP600 medium 2015 Jun 08 LP600 2015 Jun 12 LP600 medium 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 2014 Aug 27 LP600 medium 2014 Aug 20 2015 Jun 12 LP600 2014 Sep 03 LP600 medium 2014 Nov 09 LP600 2014 Aug 24 LP600 medium 2015 Jun 12 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 medium 2014 Jun 13 LP600 2014 Aug 23 2014 Aug 28 LP600 2014 Jul 17 LP600 2015 Jun 12 LP600 medium 2014 Aug 31 LP600 2014 Jul 17 LP600 2014 Jul 14 LP600 medium 2015 Jun 10 LP600 2014 Jul 17 LP600 medium 2014 Aug 20 2014 Nov 09 LP600 medium 2014 Aug 27 LP600 medium low low high high low high low high high high low high low high low high low high medium low low high high low low low low i(cid:48) i(cid:48) i(cid:48) i(cid:48) KOI K05202 K05205 K05206 K05207 K05210 K05211 K05212 K05216 K05219 K05220 K05223 K05224 K05225 K05228 K05230 K05232 K05235 K05236 K05237 K05238 K05241 K05243 K05245 K05247 K05248 K05249 K05254 K05255 K05257 K05261 K05267 K05269 K05279 K05281 K05283 K05284 K05287 K05288 K05290 K05297 K05298 K05300 K05308 K05309 K05310 K05321 K05322 K05323 K05324 K05325 K05326 K05327 K05329 K05331 K05332 K05335 K05336 K05339 K05340 K05341 K05342 K05344 K05347 K05356 K05357 K05372 K05373 K05376 K05379 K05380 K05381 K05387 K05388 K05390 K05403 K05405 K05406 mi (mags) 14.48 12.84 14.74 13.05 14.86 12.7 12.12 15.27 14.94 11.83 13.68 14.69 15.07 14.61 13.11 13.55 12.74 13.09 15.11 10.86 15.23 12.21 13.03 14.81 14.91 14.35 10.73 13.16 14.98 14.92 14.33 15.39 14.78 13.48 13.84 14.57 13.56 15.43 15.67 14.44 13.54 12.9 14.1 14.68 13.99 13.64 13.91 13.67 13.71 15.46 12.41 14.99 15.39 14.89 14.3 11.19 13.36 15.35 14.97 12.49 15.08 13.73 14.38 14.8 13.86 13.79 11.51 13.56 12.22 15.31 15.4 14.07 13.4 14.46 14.78 13.87 11.19 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? i(cid:48) i(cid:48) i(cid:48) low low low high high high low high high low low low high high low high 2014 Jun 19 LP600 medium 2015 Jun 03 LP600 medium 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2014 Jul 14 LP600 medium 2014 Jun 17 LP600 2014 Aug 22 medium 2014 Aug 31 LP600 medium 2014 Sep 03 LP600 medium 2014 Sep 03 LP600 2015 Jun 03 LP600 medium 2014 Jul 14 LP600 medium 2014 Nov 08 LP600 medium 2014 Jul 17 LP600 medium 2014 Jun 19 LP600 2014 Aug 31 LP600 2014 Aug 20 2015 Jun 06 LP600 2014 Aug 29 LP600 medium 2014 Jul 19 LP600 medium 2015 Jun 12 LP600 2014 Sep 03 LP600 2015 Jun 03 LP600 medium 2015 Jun 10 LP600 2014 Jul 19 LP600 medium 2014 Sep 03 LP600 medium 2014 Aug 22 2015 Jun 10 LP600 medium 2014 Jul 16 LP600 medium 2014 Jul 17 LP600 2015 Jun 04 LP600 2015 Jun 10 LP600 2015 Jun 08 LP600 2015 Jun 04 LP600 medium 2014 Jun 19 LP600 medium 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2014 Aug 29 LP600 medium 2015 Jun 10 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 medium 2014 Jul 17 LP600 2015 Jun 08 LP600 medium 2014 Jul 16 LP600 2015 Jun 04 LP600 2014 Jul 14 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2015 Jun 08 LP600 medium 2014 Aug 27 LP600 medium 2014 Aug 20 2014 Sep 01 LP600 2015 Jun 08 LP600 2014 Aug 31 LP600 medium 2015 Jun 12 LP600 medium 2014 Jun 13 LP600 2014 Jul 16 LP600 2015 Jun 03 LP600 2014 Jun 19 LP600 2015 Jun 03 LP600 medium 2014 Aug 23 LP600 medium 2015 Jun 04 LP600 2014 Jun 19 LP600 medium 2014 Jun 19 LP600 medium 2014 Jun 19 LP600 medium 2014 Jul 18 LP600 medium 2015 Jun 05 LP600 2014 Aug 31 LP600 medium 2014 Jul 18 LP600 2014 Aug 31 LP600 2014 Aug 26 LP600 2014 Jul 16 LP600 2014 Jul 14 LP600 medium 2015 Jun 10 LP600 medium 2015 Jun 10 LP600 2014 Jul 17 LP600 medium 2014 Aug 20 medium high high low low high low high high low low low low low high low low low high high low low i(cid:48) i(cid:48) yes yes yes yes yes yes yes yes yes yes M4 M3 M1 M4 M5 M3 M4 M4 M6 M5 M4 M3 M8 M5 M4 M0 M4 M4 M4 M1 M4 M5 M0 M4 M3 M5 M3 M4 M4 M0 M2 M2 M5 M4 M2 M5 M4 M2 M3 M4 M4 M5 M2 M1 M4 M5 M5 M4 M1 M8 M1 M5 M3 M5 M4 K2 M0 M3 K4 M5 M3 M4 M6 M3 M5 M1 M5 M6 M4 M5 M0 M3 M2 M4 M5 M4 M0 M6 M0 yes yes yes M4 M3 M5 M0 M2 M4 M1 M5 M1 M5 M6 M3 M4 M2 M4 M0 M1 M3 M5 M5 M2 M2 M5 M2 M2 M2 M4 M3 M5 M4 M4 M1 M5 M4 M5 M0 M3 M5 M4 M4 M4 M3 M5 M0 M6 M5 M1 M5 M3 M2 M4 M2 M3 M3 M5 M2 M1 M4 M5 M4 M0 M4 M3 Robo-AO Kepler Planetary Candidate Survey III 33 KOI K05408 K05409 K05410 K05411 K05412 K05413 K05416 K05417 K05418 K05433 K05435 K05440 K05447 K05451 K05458 K05459 K05461 K05465 K05471 K05472 K05476 K05480 K05478 K05482 K05485 K05486 K05487 K05490 K05497 K05498 K05499 K05500 K05506 K05509 K05511 K05514 K05517 K05520 K05521 K05528 K05529 K05530 K05533 K05541 K05545 K05546 K05553 K05554 K05556 K05564 K05565 K05566 K05567 K05568 K05570 K05572 K05577 K05578 K05579 K05583 K05585 K05587 K05597 K05604 K05605 K05608 K05612 K05622 K05623 K05625 K05628 K05632 K05633 K05638 K05649 K05652 K05653 mi (mags) 13.74 14.82 13.66 13.34 15.91 13.86 15.91 14.93 14.92 14.38 15.58 15.15 12.55 13.16 15.6 11.32 13.08 13.67 15.67 12.35 15.35 16.27 15.48 14.98 16.64 12.57 15.67 15.26 10.84 12.65 15.12 13.67 14.0 15.21 12.39 13.68 13.6 14.37 13.69 14.95 11.7 12.79 13.35 14.7 13.41 14.69 15.26 11.04 13.16 10.16 11.83 12.55 14.86 13.44 14.55 13.6 13.29 10.89 15.24 15.04 12.1 14.27 13.72 12.79 14.32 15.01 15.6 15.44 14.34 15.71 15.34 11.04 10.99 14.59 14.93 14.64 15.17 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low low i(cid:48) i(cid:48) high low low low high high high high high high low high high high low low low low high high low 2014 Jul 17 LP600 medium 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2015 Jun 10 LP600 medium 2015 Jun 10 LP600 2014 Jul 18 LP600 medium 2014 Sep 01 LP600 2015 Jun 10 LP600 2014 Aug 21 LP600 medium 2014 Jun 19 LP600 medium 2014 Aug 24 LP600 medium 2014 Aug 28 LP600 2014 Aug 21 2015 Jun 10 LP600 medium 2015 Jun 03 LP600 2014 Aug 23 2014 Jul 17 LP600 medium 2014 Jun 19 LP600 2015 Jun 03 LP600 2014 Jul 14 LP600 2014 Aug 26 LP600 medium 2014 Aug 29 LP600 2014 Aug 24 LP600 2014 Aug 31 LP600 2014 Jul 18 LP600 2015 Jun 12 LP600 2014 Aug 31 LP600 2014 Aug 28 LP600 medium 2014 Aug 21 2015 Jun 07 LP600 2014 Aug 27 LP600 2015 Jun 12 LP600 medium 2014 Jul 17 LP600 2014 Aug 24 LP600 medium 2015 Jun 12 LP600 2014 Jul 17 LP600 2014 Aug 21 LP600 medium 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2014 Jul 17 LP600 2015 Jun 07 LP600 2014 Aug 20 2014 Jun 15 LP600 medium 2015 Jun 07 LP600 medium 2014 Aug 20 2014 Sep 03 LP600 medium 2014 Aug 23 LP600 2014 Aug 20 2014 Jun 13 LP600 2014 Jun 13 LP600 2014 Aug 23 2015 Jun 04 LP600 medium 2015 Jun 07 LP600 medium 2015 Jun 12 LP600 medium 2014 Aug 21 LP600 2015 Jun 08 LP600 medium 2014 Jul 16 LP600 medium 2014 Nov 09 LP600 2014 Aug 28 LP600 2014 Jul 18 LP600 2014 Jul 17 LP600 2015 Jun 04 LP600 medium 2015 Jun 12 LP600 medium 2014 Jul 16 LP600 medium 2015 Jun 12 LP600 2014 Aug 27 LP600 medium 2014 Sep 01 LP600 2014 Aug 27 LP600 2014 Jul 17 LP600 medium 2014 Aug 31 LP600 2014 Aug 28 LP600 medium 2014 Jul 17 LP600 2014 Aug 20 2014 Jun 19 LP600 2014 Jul 17 LP600 2014 Jul 17 LP600 medium 2014 Aug 21 LP600 low high high high high high low high high low high low high low medium low low high high low low low low low i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) yes yes yes yes yes yes yes yes yes M5 M2 M3 M0 M2 M4 M5 M1 M4 M3 M5 M4 M5 M4 M4 M5 M5 M3 M5 M4 M1 M1 M1 M5 M3 M0 M4 M3 M4 M4 M3 M4 M4 M5 M4 M5 M1 M5 M1 M4 M0 M4 M2 M3 M3 M4 M5 M6 M4 M2 M4 M3 M5 M5 M4 M5 M4 M5 M5 M3 M5 M3 M3 M4 M3 M3 M3 M3 M5 M4 M3 M4 M3 M5 M4 KOI K05656 K05657 K05660 K05663 K05665 K05671 K05680 K05684 K05688 K05692 K05694 K05695 K05702 K05704 K05706 K05707 K05712 K05718 K05727 K05736 K05737 K05740 K05747 K05748 K05749 K05758 K05762 K05772 K05774 K05782 K05785 K05786 K05788 K05795 K05796 K05798 K05799 K05800 K05801 K05802 K05805 K05806 K05807 K05810 K05815 K05816 K05819 K05825 K05826 K05827 K05829 K05834 K05835 K05840 K05842 K05849 K05852 K05855 K05856 K05869 K05875 K05877 K05885 K05888 K05889 K05899 K05902 K05904 K05909 K05911 K05913 K05918 K05919 K05920 K05924 K05925 K05927 mi (mags) 12.51 14.21 13.28 13.06 11.31 13.39 13.03 10.78 16.58 14.14 14.29 14.92 15.26 13.14 15.58 15.03 16.55 16.68 15.26 13.76 13.77 14.64 13.63 14.49 15.4 14.84 15.39 13.82 10.71 11.09 14.76 14.48 14.37 15.72 14.52 15.39 13.61 15.16 15.38 13.95 14.25 12.17 14.27 13.5 13.43 12.24 13.95 13.74 15.34 13.33 14.61 15.56 14.91 13.29 14.58 14.24 15.37 14.93 14.55 15.35 14.5 15.34 14.66 15.23 15.24 13.9 14.18 11.47 15.58 15.2 15.2 13.99 13.42 13.59 15.46 12.98 13.12 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low low low high high low low low high high high high low high high low low high low high low high 2014 Jul 17 LP600 2014 Aug 21 LP600 medium 2014 Jul 16 LP600 medium 2015 Jun 07 LP600 medium 2014 Jul 17 LP600 2014 Jun 16 LP600 medium 2014 Jul 12 LP600 2014 Jun 19 LP600 2014 Aug 28 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 medium 2014 Aug 24 LP600 medium 2014 Jun 13 LP600 2014 Aug 28 LP600 2014 Aug 23 LP600 2014 Aug 24 LP600 2014 Aug 31 LP600 2014 Sep 02 LP600 2015 Jun 04 LP600 2015 Jun 12 LP600 2014 Aug 21 LP600 medium 2014 Jul 16 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2014 Sep 03 LP600 medium 2014 Sep 03 LP600 medium 2015 Jun 04 LP600 2014 Sep 01 LP600 2014 Aug 21 LP600 2014 Jul 16 LP600 medium 2015 Jun 04 LP600 2014 Jul 16 LP600 medium 2014 Sep 03 LP600 2014 Aug 21 LP600 medium 2014 Aug 23 LP600 2015 Jun 04 LP600 medium 2014 Sep 02 LP600 2014 Sep 01 LP600 2015 Jun 06 LP600 2014 Jul 16 LP600 medium 2014 Aug 20 medium 2014 Jul 16 LP600 medium 2014 Jul 16 LP600 2014 Jul 16 LP600 medium 2014 Sep 02 LP600 2014 Sep 02 LP600 medium 2014 Aug 21 LP600 medium 2014 Aug 23 LP600 2015 Jun 04 LP600 medium 2014 Jul 18 LP600 medium 2014 Aug 23 LP600 2014 Sep 02 LP600 2014 Jun 19 LP600 2014 Jul 17 LP600 medium 2014 Jun 19 LP600 medium 2014 Sep 03 LP600 medium 2014 Jul 18 LP600 medium 2014 Jul 18 LP600 2014 Aug 23 LP600 2015 Jun 04 LP600 2014 Sep 02 LP600 2014 Aug 21 LP600 medium 2014 Sep 02 LP600 2014 Sep 01 LP600 2015 Jun 04 LP600 medium 2014 Jul 17 LP600 medium 2014 Aug 20 medium 2014 Aug 23 LP600 2014 Sep 01 LP600 2014 Sep 03 LP600 medium 2014 Aug 29 LP600 medium 2014 Aug 29 LP600 medium 2014 Jul 13 LP600 medium 2014 Aug 28 LP600 2015 Jun 12 LP600 medium 2014 Aug 29 LP600 low low low high high low high low low high high high high low high high high high low i(cid:48) i(cid:48) yes yes yes yes yes yes yes yes M4 M5 M3 M5 M4 M5 M6 M3 M7 M4 M3 M3 M5 M1 M0 M6 M2 M6 M2 M5 M3 M2 M3 M4 M5 M4 M1 M3 M5 M0 M4 M4 M3 M4 M1 M2 M5 M5 M4 M5 M5 M6 M4 M4 M6 M4 M4 M4 M0 M4 M3 M4 M4 M4 M1 M4 M0 M5 M4 M5 M4 M4 M4 M5 M6 M0 M4 M5 M3 M2 M4 M4 M5 Ziegler et al. TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low 34 KOI K05932 K05935 K05938 K05948 K05949 K05950 K05952 K05959 K05960 K05965 K05968 K05971 K06047 K06093 K06101 K06102 K06103 K06108 K06109 K06111 K06112 K06115 K06118 K06120 K06130 K06132 K06134 K06137 K06141 K06145 K06150 K06151 K06165 K06175 K06176 K06178 K06183 K06194 K06202 K06209 K06223 K06228 K06245 K06253 K06254 K06258 K06259 K06261 K06262 K06263 K06266 K06267 K06291 K06293 K06295 K06299 K06311 K06318 K06320 K06326 K06329 K06338 K06341 K06342 K06343 K06352 K06353 K06355 K06357 K06361 K06368 K06372 K06375 K06378 K06385 K06391 K06392 mi (mags) 15.89 15.34 12.67 13.65 13.14 13.47 14.79 14.29 13.1 15.59 11.37 13.4 11.96 13.33 12.55 10.82 13.44 11.9 11.86 12.91 12.79 13.5 12.77 15.42 15.12 14.62 11.94 14.19 8.72 14.26 12.18 14.4 11.93 10.66 11.46 15.86 13.25 12.06 11.38 16.15 15.0 14.22 12.43 15.98 13.24 11.21 14.6 14.14 15.77 12.77 14.52 15.37 13.3 14.2 15.62 13.85 9.0 15.73 13.12 13.39 14.02 12.16 15.2 12.42 14.54 15.22 13.47 14.7 15.42 11.27 15.81 14.6 14.56 13.43 13.77 11.3 14.48 i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) i(cid:48) high high high high high medium high high high high high high low high high low high high high high high low high high high high low 2015 Jun 12 LP600 2014 Aug 31 LP600 medium 2014 Aug 22 medium 2014 Sep 03 LP600 2014 Aug 29 LP600 2014 Jul 11 LP600 medium 2014 Aug 31 LP600 2014 Jul 18 LP600 medium 2014 Jul 11 LP600 2014 Sep 03 LP600 medium 2014 Aug 21 2014 Jul 11 LP600 medium 2015 Jun 06 LP600 2015 Jun 04 LP600 medium 2015 Jun 07 LP600 2014 Aug 22 2015 Jun 03 LP600 2014 Aug 22 2015 Jun 07 LP600 2015 Jun 04 LP600 medium 2015 Jun 03 LP600 medium 2015 Jun 08 LP600 medium 2015 Jun 07 LP600 2015 Jun 08 LP600 2015 Jun 03 LP600 2015 Jun 12 LP600 2015 Jun 03 LP600 medium 2015 Jun 06 LP600 2014 Aug 22 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 2015 Jun 07 LP600 2015 Jun 05 LP600 2014 Aug 22 2015 Jun 07 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 2014 Aug 23 2015 Jun 05 LP600 2015 Jun 04 LP600 2015 Jun 10 LP600 medium 2015 Jun 06 LP600 medium 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2015 Jun 06 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 07 LP600 2015 Jun 05 LP600 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 low low high low low low low low high low low high low high low high low low low high high high high low low low KOI K06398 K06399 K06401 K06408 K06409 K06410 K06414 K06415 K06420 K06425 K06438 K06447 K06449 K06450 K06454 K06455 K06460 K06464 K06469 K06475 K06482 K06483 K06489 K06497 K06499 K06501 K06503 K06505 K06507 K06511 K06512 K06513 K06514 K06516 K06518 K06519 K06521 K06522 K06526 K06527 K06532 K06533 K06534 K06538 K06539 K06542 K06546 K06557 K06559 K06560 K06563 K06567 K06568 K06570 K06574 K06578 K06579 K06582 K06598 K06601 K06602 K06604 K06605 K06610 K06611 K06617 K06618 K06630 K06631 K06635 K06646 K06648 K06649 K06653 K06654 K06657 K06668 mi (mags) 10.59 14.72 13.75 11.46 13.37 13.92 13.86 14.03 14.23 14.16 12.22 12.96 15.67 12.36 15.0 14.13 11.11 13.72 12.87 13.73 13.6 12.53 13.7 13.34 13.39 13.57 13.91 12.1 11.48 15.75 15.04 15.63 11.86 15.93 13.53 11.4 9.81 14.13 13.89 12.33 15.57 12.63 12.05 14.52 12.51 11.27 15.42 13.8 14.77 12.94 11.78 13.09 15.52 10.9 14.68 11.38 15.67 15.22 11.97 14.6 10.2 12.96 11.32 15.34 13.33 15.03 15.17 12.95 13.27 14.35 15.13 11.87 13.57 14.93 13.48 13.7 13.18 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? high high high high high low high high high high low low high low high high low high low high low 2015 Jun 05 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 medium 2015 Jun 06 LP600 2015 Jun 08 LP600 medium 2015 Jun 07 LP600 medium 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 08 LP600 medium 2015 Jun 03 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 08 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 08 LP600 medium 2015 Jun 05 LP600 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 07 LP600 2015 Jun 04 LP600 medium 2015 Jun 05 LP600 2015 Jun 12 LP600 medium 2015 Jun 10 LP600 medium 2015 Jun 08 LP600 medium 2015 Jun 05 LP600 medium 2015 Jun 06 LP600 medium 2015 Jun 05 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 2015 Jun 07 LP600 2015 Jun 08 LP600 medium 2015 Jun 08 LP600 2015 Jun 06 LP600 2015 Jun 12 LP600 medium 2015 Jun 08 LP600 medium 2015 Jun 07 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 10 LP600 medium 2015 Jun 12 LP600 medium 2015 Jun 05 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 2015 Jun 06 LP600 2015 Jun 06 LP600 2015 Jun 03 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 06 LP600 2015 Jun 07 LP600 2015 Jun 12 LP600 2015 Jun 08 LP600 2015 Jun 05 LP600 medium 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 10 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 2015 Jun 05 LP600 medium 2015 Jun 10 LP600 2015 Jun 10 LP600 2015 Jun 12 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 07 LP600 2015 Jun 03 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 medium 2015 Jun 08 LP600 low high low high low low low high high high low high low high low high high low high high high low high low high low low high M6 M2 M2 M2 M4 M3 M0 M5 M3 M4 M3 M2 M5 M4 M0 M3 M3 M3 M4 M8 M4 M4 M2 M4 M2 M1 M5 M2 M4 M2 M2 M4 M3 M5 M4 M2 M3 M1 M3 M1 M1 M4 M4 M5 M4 M5 M3 M3 M5 M2 M0 M1 M5 M6 M3 M3 M2 M0 M2 M4 M8 M4 M3 M0 M3 M4 M1 M5 yes yes yes yes yes yes yes yes yes yes yes yes M2 M3 M5 M4 M5 M5 M4 M5 M5 M5 M5 M3 M4 M5 M4 M1 M4 M4 M2 M3 M3 M4 M4 M4 M2 M4 M2 M2 M5 M6 M3 M4 M1 M4 M6 M5 M1 M4 M2 M3 M3 M4 M2 K4 M3 M4 M1 M3 M4 M0 M6 M3 M2 M4 M5 M5 M3 M5 M2 M4 M5 M4 M2 M1 M4 M0 M2 M1 M0 M1 yes yes yes yes yes yes yes yes Robo-AO Kepler Planetary Candidate Survey III KOI K06670 K06672 K06680 K06687 K06689 K06705 K06706 K06707 K06728 K06731 K06733 K06734 K06744 K06745 K06746 K06747 K06750 K06751 K06752 K06753 K06762 K06763 K06774 K06781 K06788 K06791 K06800 K06819 K06839 K06860 K06862 K06863 K06877 K06883 K06888 K06889 K06890 K06892 K06897 K06898 K06913 K06919 K06921 K06925 K06927 K06934 K06936 K06949 K06952 K06954 mi (mags) 12.67 15.18 13.69 15.79 15.03 15.67 13.85 15.86 13.91 15.26 13.9 15.48 10.21 15.23 14.12 14.67 14.43 15.18 15.54 9.28 14.45 12.16 15.57 14.95 15.6 14.36 12.85 11.43 16.67 13.29 14.8 15.74 10.97 14.93 12.85 11.65 15.38 14.99 15.38 13.64 13.75 15.86 15.1 15.66 14.79 14.16 15.18 12.05 13.62 13.11 TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low low high low low low high low low high low low high low low low high high 2015 Jun 05 LP600 medium 2015 Jun 04 LP600 2015 Jun 03 LP600 medium 2015 Jun 07 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 10 LP600 2015 Jun 12 LP600 2015 Jun 06 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 medium 2015 Jun 08 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 medium 2015 Jun 04 LP600 medium 2015 Jun 04 LP600 2015 Jun 10 LP600 medium 2015 Jun 03 LP600 2015 Jun 07 LP600 2015 Jun 12 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 03 LP600 medium 2015 Jun 04 LP600 2015 Jun 07 LP600 2015 Jun 08 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 12 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 08 LP600 medium 2015 Jun 10 LP600 2015 Jun 12 LP600 2015 Jun 03 LP600 2015 Jun 03 LP600 2015 Jun 05 LP600 medium 2015 Jun 04 LP600 2015 Jun 03 LP600 2015 Jun 04 LP600 medium 2015 Jun 06 LP600 medium high low low low low low low high low low low low low low high low high high low M2 M3 M4 M4 M2 M0 M3 M5 M1 M0 M3 M2 M4 M4 M1 M4 M2 M4 M4 M1 M5 M2 M5 M3 M4 M8 M6 M3 M8 M6 M2 M3 M4 M3 M1 M4 M3 K1 M4 M0 M5 M5 M4 M3 M4 M4 yes yes yes yes yes K06960 K06962 K06978 K06980 K06993 K06994 K07007 K07014 K07020 K07033 K07044 K07051 K07054 K07069 K07084 K07092 K07093 K07109 K07133 K07144 K07182 K07205 K07208 K07209 K07219 K07224 K07232 KOI K07235 K07255 K07266 K07279 K07296 K07298 K07305 K07319 K07327 K07345 K07350 K07352 K07378 K07395 K07400 K07426 K07430 K07432 K07448 K07449 K07468 13.87 15.43 13.36 15.19 15.57 12.83 15.7 11.54 13.48 11.51 11.52 15.86 10.29 15.57 11.96 15.45 13.65 13.68 14.91 15.7 16.54 14.11 15.56 13.82 11.25 15.67 15.61 mi (mags) 14.71 14.25 11.36 13.72 10.63 15.48 9.98 15.93 15.86 15.04 12.73 13.25 16.32 11.7 15.52 15.35 11.47 15.01 11.27 14.34 14.76 35 yes yes M3 M5 M4 M0 M3 M4 M4 M4 M4 M4 M6 M6 M5 M1 M0 M2 M2 M2 M3 M4 M5 M3 M3 M4 M1 M3 high low low low high high high high high high low high low 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 2015 Jun 03 LP600 medium 2015 Jun 12 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 medium 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 2015 Jun 05 LP600 2015 Jun 07 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 2015 Jun 08 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 2015 Jun 08 LP600 medium 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 low low low low high low high low low TABLE 9 -- Continued ObsID Filter Obs. Qual. Latest Det. Comp. Comp. SpT Det.? low low high high low high low low high 2015 Jun 12 LP600 2015 Jun 04 LP600 2015 Jun 08 LP600 2015 Jun 04 LP600 medium 2015 Jun 12 LP600 2015 Jun 04 LP600 2015 Jun 07 LP600 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 10 LP600 2015 Jun 06 LP600 medium 2015 Jun 04 LP600 2015 Jun 04 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 2015 Jun 05 LP600 2015 Jun 04 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 2015 Jun 12 LP600 medium 2015 Jun 12 LP600 low low high low low high low high low M3 M0 M4 M3 M4 M5 M4 M4 M5 M0 M4 M1 M3 M4 M5 M3 M3 M5 M2 yes yes yes
1610.09073
1
1610
2016-10-28T03:59:34
Latitudinal variability in Jupiter's tropospheric disequilibrium species: GeH$_4$, AsH$_3$ and PH$_3$
[ "astro-ph.EP" ]
Jupiter's tropospheric composition is studied using high resolution spatially-resolved 5-micron observation from the CRIRES instrument at the Very Large Telescope. The high resolving power (R=96,000) allows us to spectrally resolve the line shapes of individual molecular species in Jupiter's troposphere and, by aligning the slit north-south along Jupiter's central meridian, we are able to search for any latitudinal variability. Despite the high spectral resolution, we find that there are significant degeneracies between the cloud structure and aerosol scattering properties that complicate the retrievals of tropospheric gaseous abundances and limit conclusions on any belt-zone variability. However, we do find evidence for variability between the equatorial regions of the planet and the polar regions. Arsine (AsH$_3$) and phosphine (PH$_3$) both show an enhancement at high latitudes, while the abundance of germane (GeH$_4$) remains approximately constant. These observations contrast with the theoretical predictions from Wang et al. (2016) and we discuss the possible explanations for this difference.
astro-ph.EP
astro-ph
Latitudinal variability in Jupiter's tropospheric disequilibrium species: GeH4, AsH3 and PH3 R. S. Gilesa,∗, L. N. Fletcherb, P. G. J. Irwina aAtmospheric, Oceanic & Planetary Physics, Department of Physics, University of Oxford, Clarendon Laboratory, Parks Road, Oxford, bDepartment of Physics and Astronomy, University of Leicester, University Road, Leicester, LE1 7RH, UK OX1 3PU, UK 6 1 0 2 t c O 8 2 . ] P E h p - o r t s a [ 1 v 3 7 0 9 0 . 0 1 6 1 : v i X r a Abstract Jupiter's tropospheric composition is studied using high resolution spatially-resolved 5-m observation from the CRIRES instrument at the Very Large Telescope. The high resolving power (R=96,000) allows us to spectrally re- solve the line shapes of individual molecular species in Jupiter's troposphere and, by aligning the slit north-south along Jupiter's central meridian, we are able to search for any latitudinal variability. Despite the high spectral resolution, we find that there are significant degeneracies between the cloud structure and aerosol scattering properties that complicate the retrievals of tropospheric gaseous abundances and limit conclusions on any belt-zone variability. However, we do find evidence for variability between the equatorial regions of the planet and the polar regions. Arsine (AsH3) and phosphine (PH3) both show an enhancement at high latitudes, while the abundance of germane (GeH4) remains approximately constant. These observations contrast with the theoretical predictions from Wang et al. (2016) and we discuss the possible explanations for this difference. Keywords: Jupiter, Atmospheres, composition, Atmospheres, structure 1. Introduction The 5-m atmospheric window is a unique region of Jupiter's spectrum where a dearth of opacity from the planet's principal infrared absorbers gives us access to parts of the atmosphere that are otherwise hidden from view. In the absence of clouds, the sensitivity of the 4.5- 5.2 m spectrum peaks in the 4-8 bar region of Jupiter's atmosphere (Giles et al., 2015). These pressure levels correspond to Jupiter's middle troposphere, a region of the planet that is home to many interesting phenomena, including multiple cloud layers, disequilibrium chemical species and dynamics ranging from small-scale meteoro- logical features to planetary-scale overturning. The atmospheric window was first described by Gillett et al. (1969) and the high degree of spatial variability was noted by Westphal (1969). The brightness at 5-m is very sensitive to the opacity of the tropospheric clouds; in the relatively cloud-free belts, the bright radiation from depth can be observed, but this is partially shielded by the thicker clouds in the zones. Since then, the 5-m region has been the focus of many studies, from ground-based telescopes, airborne observatories and spacecraft. Ground- based and airborne telescopes generally offer a higher spec- tral resolution, so these observations are well suited to the study of individual gaseous species in the troposphere. ∗Corresponding author Email address: [email protected] (R. S. Giles) Bjoraker et al. (1986) used data from the Kuiper Airborne Observatory to determine the tropospheric composition, and a similar analysis was carried out by Encrenaz et al. (1996) using the Infrared Space Observatory. The United Kingdom Infrared Telescope, Canada-France-Hawaii Tele- scope and the Keck Observatory have been used by several groups in order to target individual gaseous species in dis- crete regions of the planet (B´ezard et al., 2002; Noll et al., 1989; Bjoraker et al., 2015). Germane (GeH4), arsine (AsH3) and phosphine (PH3) are all disequilibrium species in Jupiter's troposphere. Their abundances are determined by temperature- dependent equilibrium reactions and deep in the atmo- sphere, the temperature is high enough that they are chem- ically stable. The temperature in the observable tropo- sphere is much lower so we would expect the abundances to rapidly drop off (e.g. Fegley and Lodders, 1994). However, GeH4, AsH3 and PH3 have all been detected in Jupiter's 5-m window with higher abundances than would be ex- pected from equilibrium chemistry (Fink et al., 1978; Noll et al., 1989; Ridgway et al., 1976). Further background about GeH4, AsH3 and PH3 are given in Sections 6, 7 and 8 respectively. The apparently enhanced abundances are indicative of rapid vertical mixing, which brings the disequilibrium species upwards from deeper pressures where they are more abundant. Instead of dropping off rapidly with alti- tude, the abundance is "frozen in" at a roughly constant value. This observed abundance corresponds to the equi- Preprint submitted to Icarus June 7, 2021 librium abundance at a much deeper level, known as the "quench level", which depends on both the strength of ver- tical mixing and the chemical kinetic rates (Lewis and Feg- ley, 1984). In the case of PH3, the quench level is ∼900 K, which corresponds to >100 bar (Wang et al., 2016). This means that disequilibrium species such as GeH4, AsH3 and PH3 act as tracers of atmospheric dynamics in Jupiter's troposphere (Taylor et al., 2004). Because vertical mixing strengths vary as a function of latitude (Flasar and Gierasch, 1978; Wang et al., 2015), the tropospheric abundances of these species are also ex- pected to vary. Wang et al. (2016) recently carried out a theoretical study where they combined diffusion-kinetics calculations with estimates for how Jupiter's eddy diffu- sion coefficient varies with latitude, in order to predict the latitudinal distribution of disequilibrium species, in- cluding GeH4, AsH3 and PH3. Because of the different kinetics rates for the different gases, they predict that the GeH4 abundance should be a maximum in the equatorial latitudes, and decrease towards the poles, while AsH3 and PH3 should be latitudinally uniform. In this paper, we use high-resolution ground-based ob- servations from the CRIRES instrument at the Very Large Telescope in Chile to search for evidence of latitudinal vari- ability in Jupiter's tropospheric composition. In Sections 2 and 3 we describe our observations and our atmospheric retrieval algorithm. In Sections 4 and 5 we discuss the effect of cloud structure and scattering properties on the atmospheric retrievals. In Sections 6, 7 and 8 we measure the abundances of GeH4, AsH3 and PH3 respectively. The results of these sections are then discussed in Section 9. 2. Observations 2.1. CRIRES observations Observations of Jupiter 5-m window were made us- ing CRIRES, a cryogenic high-resolution infrared echelle spectrograph (Kaufl et al., 2004), located at the Euro- pean Southern Observatory's Very Large Telescope (VLT). The instrument provides long-slit (0.2×40") spectroscopy across the wavelength range 0.95-5.38 m, with a resolving power of 96,000. The observations were made on 12 November 2012 (05:00-05:40 UT) and 1 January 2013 (02:50-03:30 UT). The atmospheric conditions were better on 12 November 2012, so the majority of this work focuses on those obser- vations. However, the observations from 1 January 2013 are used in Sections 7 and 8 to show that the same con- clusions can be drawn from both datasets. On both dates, the slit was aligned north-south on Jupiter, along the planet's central meridian, allowing us to measure spatial variability with latitude, but not longi- tude. Figure 1 shows the acquisition image of Jupiter made on 12 November, using a Ks-band filter (2.2 m). The ver- tical black line shows how the narrow slit was aligned. The slit length (40") was smaller than the angular diameter of Figure 1: Acquisition image of Jupiter made at 05:02 UT on 12 November 2012, immediately before the 5-m obser- vations. This image was made using a Ks-band filter (2.2 m). The vertical black line shows the alignment of the narrow slit. Figure 2: Measured radiance at 5.0 m as a function of latitude. Jupiter on the observation dates (47-48"), so the observa- tions cut off the southern polar region. Figure 2 shows the observed radiance as a function of latitude at 5.0 m. The high degree of spatial variability between the belts and the zones is due to the varying opacity of the tropospheric clouds. On each date, observations were made in 14 different wavelength settings, which together cover the entire 4.55- 5.26 m wavelength range. The time lag between the first and last observation was ∼40 minutes; during this time period, Jupiter rotated ∼25◦, so the 14 different obser- vations of different spectral regions do not correspond to precisely the same locations on the planet. Observations were made of a standard star in order to provide radiometric calibration. Pi-2 Orionis (HIP 22509) and Zeta Tauri (HIP 26451) were used for the November and January observations respectively. Dark images and flatfields were also taken. 2 −90−60−300306090Latitude020406080Radiance (µWcm−2sr−1µm−1) 2.2. Data reduction Initial data reduction was performed using EsoRex, the ESO Recipe Execution Tool (Ballester et al., 2006). The flatfields, darks, and nodded pairs of observations were used to produce a single combined observation of Jupiter for each of the 14 wavelength segments. The same process was applied to the observations of the standard star. In both cases, wavelength calibration was achieved by com- paring telluric absorption lines with the HITRAN line database (Rothman et al., 2009). EsoRex was also used to calculate the background noise on both the stellar and the planetary observations. Subsequent data reduction was performed indepen- dently of the EsoRex software. This included straight- ening the observations, to correct observed shears in both the spectral and spatial directions. The spectral shear was corrected using the wavelength calibration map gen- erated by EsoRex. The spatial shear was cross-correlating by cross-correlating the north-south latitudinal profiles at different wavelengths i.e. requiring the belts and zones to be located at the same pixel values for all wavelengths. The jovian spectra were divided through by the standard star observations in order to remove the telluric lines, and were multiplied by the standard star reference spectrum to convert Jupiter's photon count into spectral radiance. For Pi-2 Orionis, the standard star reference spectrum was provided by EsoRex. For Zeta Tauri, a black-body curve was scaled to match the observed magnitudes from the 2MASS all-sky survey (Cutri et al., 2003). Because the slit is very narrow, some stellar flux will be lost; this was accounted for in the radiometric calibration by assuming that the star is centrally placed in the slit, using a mea- surement of the star's point spread function made in the perpendicular direction. Once this calibration has been applied, we find that the 5-m brightness temperatures vary between ∼190 K in the coolest parts of the planet and ∼260 K in the brightest regions (see Figure 2), which is roughly consistent with the 180-240 K range found in the Cassini VIMS 5-m observations (Giles et al., 2015). The noise on the spectra is roughly 5-10%, but this does not include the systematic error that could be introduced by the assumptions in the radiometric calibration. These systematic errors could increase or decrease the average radiance, but will not alter the shape of the spectra. The average radiance is primarily determined by the opacity of the upper cloud deck (Giles et al., 2015). Uncertainties in the radiometric calibration can therefore alter the appar- ent thickness of the clouds, but will not affect the analysis of the molecular species. Geometric data was calculated using information from JPL HORIZONS, and by assuming that the slit was aligned with the central meridian. Planetocentric lati- tudes were assigned to each pixel using information about Jupiter's angular size at the observation time, the angu- lar size of each pixel, the centre-of-target latitude, and the location of the planet's limb. In addition, the emission an- gle, solar zenith angle and azimuthal angle were calculated for each pixel. Due to Jupiter's rotation, each of the 14 wavelength segments observed a slightly different longitude, with a ∼ 25◦ difference between the first and last segment. Be- cause Jupiter is longitudinally inhomogeneous, this means that each wavelength segment is observing a different cloud opacity; this prevents the different segments from being joined together into a single continuous observation. These latitudinally resolved spectra will be used in the following sections in order to search for variability in the deep cloud structure (Section 5) and in the abundances of germane, arsine and phosphine (Sections 6-8). 3. Spectral modelling The CRIRES spectra were analysed with the NEME- SIS retrieval algorithm developed by Irwin et al. (2008). NEMESIS uses a radiative transfer code to compute the top-of-atmosphere spectral radiance for a given atmo- spheric profile, and then iteratively adjusts the atmo- spheric parameters in order to produce an optimal fit to the observed spectrum. For this work, we use line- by-line calculations in the radiative transfer code; pre- vious work has made use of the correlated-k approxi- mation (e.g. Giles et al., 2015) but the high spectral resolution of the CRIRES observations means that this approximation would not save significant computational time. We therefore use the more accurate line-by-line method. The majority of the spectral line data were ob- tained from the HITRAN 2008 (Rothman et al., 2009) or GEISA 2009 (Jacquinet-Husson et al., 2011) molecu- lar databases. The two exceptions are AsH3, where we use laboratory data from Dana et al. (1993) and Mandin et al. (1995), and GeH4, where we use the Spherical Top Data System (Wenger and Champion, 1998, see Section 6.1 for more details). Collision-induced absorption data is taken from Borysow and Frommhold (1986), Borysow and Frommhold (1987), Borysow et al. (1988) and Borysow (1991). We add a conservative 5% forward modelling error to the retrieval calculations, to take into account inaccu- racies in the model (e.g. line data). NEMESIS has the ability to model a multiple scatter- ing atmosphere using the plane-parallel matrix operator method of Plass et al. (1973), and we show in Section 4 that the scattering properties of Jupiter's clouds have significant effects on the retrieved gaseous abundances. Unless otherwise stated, we use the simple cloud model from Giles et al. (2015): a single, spectrally-flat, compact cloud layer, located at 0.8 bar, with a single-scattering albedo ω = 0.9 and a Henyey-Greenstein asymmetry pa- rameter g = 0.8. The reference atmospheric profile is described as in Giles et al. (2015). The gaseous species with spectral features in the 5-m region are NH3, PH3, CO, GeH4, AsH3, H2O and CH3D. Unless otherwise stated, the abundances of 3 each of these gases are allowed to vary via a single scaling parameter. 4. Scattering properties of upper cloud The high spectral resolution of the CRIRES instru- ment allows us to resolve the individual absorption lines of molecular species in Jupiter's troposphere. Using the NEMESIS retrieval algorithm, we can retrieve the abun- dances of gases, and map how these abundances vary across the planet. The high spectral resolution helps to reduce the degeneracy between different gaseous species. However, there is a degeneracy between the gaseous abun- dances and scattering parameters of Jupiter's tropospheric clouds: specifically, the asymmetry parameter, g, which defines the extent to which the cloud particles are forward scattering. An example of this effect is shown in Figure 3. This figure shows three prominent CH3D absorption features, marked by the horizontal double-headed arrows. The CRIRES data are shown in black, and is from the Equato- rial Zone of the planet. The Equatorial Zone was chosen as it is one of the cloudiest regions of the planet. Small seg- ments of data with high error bars (due to telluric contami- nation) have been removed. The three absorption features were fitted simultaneously, and the coloured lines show the fits that can be obtained using two different assumptions about the scattering properties of the single cloud deck. Both cases can produce a good fit to the broad CH3D fea- tures, but the retrieved CH3D abundances are significantly different: 0.16 ppm for the case where g = 0.4 and 0.08 ppm for the case where g = 0.9. This effect is due to the different amount of reflected sunlight in the two models. A higher asymmetry parame- ter (more forward scattering) leads to a higher fraction of radiation originating from the planet, and a lower fraction from reflected sunlight. In contrast, a lower asymmetry pa- rameter (less forward scattering) leads to a higher fraction due to reflected sunlight. A higher g value therefore leads to stronger absorption features for a given gaseous abun- dance, and hence lower retrieved abundances. It should be noted that this is dependent on the vertical location of the cloud; if the cloud were located below the main line forming region, the effect would be less pronounced. How- ever, Giles et al. (2015) showed that the main tropospheric cloud deck must be located well above the line forming in order to account for the highly variable 5-m radiation. This degeneracy between cloud scattering properties and retrieved gaseous abundances is only present in re- gions of the planet with thick cloud cover (zones like the Equatorial Zone). The belts of the planet are relatively cloud-free, so changing the cloud parameters has little im- pact on the retrievals. This means that different assump- tions about the cloud properties can affect the latitudinal profiles of the gases; one set of parameters could suggest that a molecular species has a constant abundance across the planet, while another set of parameters could suggest g 0.40 0.45 0.50 0.55 0.60 0.65 0.70 0.75 0.80 0.85 0.90 χ2/N 2.25 2.11 1.95 1.80 1.66 1.54 1.43 1.31 1.19 1.52 2.40 Table 1: The goodness-of-fit values (χ2/N) for a range of different asymmetry parameters. χ2/N is minimised when g = 0.8. that the abundance varies between belts and zones. Select- ing the correct scattering parameters is therefore of critical importance. In Giles et al. (2015) we constrained the scattering parameters of Jupiter's tropospheric clouds using limb darkening observations from the VIMS instrument on the Cassini spacecraft. This analysis ruled out strongly for- ward scattering particles (g > 0.9), but a broad range of values were consistent with the data. In order to fur- ther constrain the asymmetry parameter properties of the clouds, we can make use of the Fraunhofer lines that are present in the CRIRES data because of the reflected sun- light. These narrow solar absorption lines can be seen in Figure 3 (marked by the vertical single-headed arrows). The strength of these features also depends on the scat- tering properties: lower g values lead to higher fractions of reflected sunlight and deeper absorption features. Bjo- raker et al. (2015) also made use of a Fraunhofer line at 4.67 m to determine the reflectance of the upper tropo- spheric clouds. We ran retrievals using a range of different g values in order to determine which value gave the best fit to the CRIRES data from the EqZ. The entire 5-m spectral range was used, excluding regions with high error bars due to telluric contamination. For each value of g, the goodness-of-fit was calculated and the results are given in Table 1. The best fit was obtained for g = 0.8, which is consistent with previous analyses (Giles et al., 2015; Irwin et al., 2008). In the following sections, we will use a cloud with an asymmetry parameter g = 0.8. We also make the assumption that the asymmetry parameter does not vary with latitude. In each case, we will investigate the extent to which the results are dependent on the assumed asymmetry parameter. 4 Figure 3: Fits to three CH3D absorption features using two different values of the cloud particle asymmetry parameter, g. The observational data (black) is from the Equatorial Zone of Jupiter. The location of the CH3D absorption features are marked by the horizontal double-headed arrows. The location of the most prominent Fraunhofer lines are marked by the vertical single-headed arrows. 5. CH3D and the deep cloud structure 5.1. Introduction Methane (CH4) was one of the first gases to be discov- ered in Jupiter's atmosphere when Wildt (1932) identi- fied previously unknown spectral features as methane and ammonia absorption lines. It is the third most abundant gaseous species in Jupiter's upper atmosphere, with a vol- ume mixing ratio (VMR) of 2.04×10−3 in the troposphere detected by the Galileo Probe Mass Spectrometer (Wong et al., 2004). Unlike many other gaseous species, this abun- dance is not expected to have any spatial variability in the troposphere. This is because CH4 is chemically stable throughout the troposphere and does not condense at the temperatures found in Jupiter's atmosphere (Taylor et al., 2004), which means that it should be well-mixed. As a methane isotopologue, CH3D (deuterated methane) is also constant throughout Jupiter's tropo- sphere. Unlike CH4, no in situ measurements have been made of the jovian CH3D abundance, but a tropospheric abundance of 0.16±0.05 ppm has been estimated from ob- servations made with the Infrared Space Observatory (Lel- louch et al., 2001). The CH3D abundance is particularly relevant to our study because CH3D has several strong absorption features in the 5-m window. It is the spatial homogeneity of CH3D that provides an opportunity to study Jupiter's deep cloud structure. Be- cause the CH3D abundance is expected to be constant with latitude, any observed variation in the spectral line- shape must be due to another factor. One way of reproduc- ing variable lineshapes is by varying the deep tropospheric cloud structure (Bjoraker et al., 2015). In this section, we will search for spatial variability in the CH3D lineshapes, and will explore how this can be interpreted in terms of cloud structure. 5 5.2. Analysis 5.2.1. CH3D There are several CH3D features in the 5-m window. Three of the strongest features are shown in Figure 4. These three features are within the same wavelength seg- ment, which means they are from the same longitude (i.e. were observed simultaneously) and can therefore be anal- ysed simultaneously. The CRIRES data are shown in black, and the arrows show the locations of the CH3D features. The upper panel shows the data from the cool Equatorial Zone (EqZ, 5◦S-4◦N), while the lower panel shows data from the warm South Equatorial Belt (SEB, 16◦S-6◦S). Figure 4 shows that the CH3D features are con- siderably narrower in the EqZ than they are in the SEB. This difference was further explored by performing re- trievals with the NEMESIS retrieval algorithm. Based on the conclusion of Section 4, we initially assumed a cloud structure made up of a single compact cloud layer at 0.8 bar, with scattering properties ω = 0.9 and g = 0.8. The optical thickness of this cloud was allowed to vary, along with the abundances of the molecular species as described in Section 3. For the SEB, the retrieved CH3D abundance was 0.15 ppm, approximately the same as the value ob- tained by Lellouch et al. (2001). However, the EqZ pro- duced a much lower value of 0.09 ppm. As a further test, additional retrievals were run where the EqZ CH3D abundance was fixed to the higher level and the SEB abundance was fixed to the lower level. The results of this are shown by the coloured lines in Figure 4 and they show that the EqZ and the SEB cannot be fit using the same CH3D abundance. A VMR of 0.15 ppm fits the SEB data, but is too broad for the EqZ data, and a VMR of 0.09 ppm fits the EqZ data but is too narrow for the SEB. However, a factor ∼2 difference in CH3D abundance for different regions of the planet is not physically plausible. 4.6264.6284.6304.632Wavelength (µm)01234Radiance (µWcm−2sr−1µm−1)4.6424.6444.6464.648Wavelength (µm)4.6604.6644.668Wavelength (µm)g = 0.4g = 0.9 Figure 4: CH3D absorption feature fits in two different regions of the planet. CRIRES data is shown in black, and the CH3D features are shown by the horizontal arrows. Each colour shows the fit obtained using different CH3D abundances: 0.09 ppm (the best fit value for the EqZ) and 0.15 ppm (the best fit value for the SEB). 6 4.6264.6284.6304.632Wavelength (µm)01234Radiance (µWcm−2sr−1µm−1)(a) Equatorial Zone4.6424.6444.6464.648Wavelength (µm)4.6604.6644.668Wavelength (µm)4.6264.6284.6304.632Wavelength (µm)0102030405060Radiance (µWcm−2sr−1µm−1)(b) South Equatorial Belt4.6424.6444.6464.648Wavelength (µm)4.6604.6644.668Wavelength (µm)VMR = 0.09 ppmVMR = 0.15 ppm As described in the previous section, methane and its iso- topologues are expected to be well-mixed in the tropo- sphere, with no spatial variability. We must therefore con- sider an alternative explanation for this change in CH3D lineshape. 5.2.2. Deep cloud structure One possibility is spatial variability in the deep cloud structure. As described by Bjoraker et al. (2015), this can be modelled by considering the presence of deep clouds at the pressure levels where water is expected to condense (∼5 bar). If there is very little cloud opacity at these pres- sures, the radiation is originating from deep in the atmo- sphere, where the pressure is higher. Pressure broadening therefore leads to a lineshape with broad wings. Inserting an optically thick cloud layer at 5 bar has two effects on the spectrum: (i) it reduces the continuum radiance, de- creasing the apparent depth of the absorption feature and (ii) it moves the weighting function upwards to a region of lower pressures and cooler temperatures, leading to a narrower lineshape. Since the presence of a deep cloud can act to narrow the lineshape, this suggests that we may be able to fit the narrow EqZ absorption features while keeping the CH3D abundance fixed at the retrieved value for the SEB. To test this we inserted a deep cloud deck at 5 bar. We assumed the same simple properties as the upper cloud: compact, spectrally flat, ω = 0.9 and g = 0.8. We then performed a retrieval where the opacity of this deep cloud was allowed to vary. The CH3D abundance was held fixed at 0.149 ppm and, as before, the upper cloud and other gaseous species were also allowed to vary. The results are shown by the yellow line in Figure 5. For comparison, we also show the case where there is no deep cloud in blue (identical to the blue line in Figure 4a). The comparison between the blue and yellow lines shows that this addition considerably improves the fit. The re- trieved deep cloud opacity in the EqZ is 230, making the cloud essentially opaque. This agrees with Bjoraker et al. (2015), where the deep cloud in the EqZ was assumed to be opaque. As with Bjoraker et al. (2015), this shows that the difference in CH3D lineshape between the EqZ and the SEB can be accounted for by the variations in the deep cloud structure; the deep cloud is transparent in the SEB and opaque in the EqZ. Based on Bjoraker et al. (2015), the initial analysis was performed using an opaque cloud at 5 bar. We carried out further retrievals in order to test the sensitivity to the cloud location. If the cloud is placed too deep, it falls outside the pressure range probed by 5-m spectra, and can no longer influence the spectral shape. If the cloud is placed too high, it simply plays the same role as the origi- nal 0.8-bar cloud and so does not improve the fit. However, between these two extremes, we find that a broad range of cloud locations from 2 to 7 bar can provide a good fit to the data, with deeper pressure levels requiring higher cloud opacities. In the absence of further constraint, we continue the analysis using the original 5 bar value. 5.2.3. Latitudinal retrieval In the previous sections, we were just comparing two dis- crete parts of the planet, the EqZ and the SEB. We now extend that analysis to perform a full latitudinal retrieval of the deep cloud opacity, using the same CH3D features as in the previous sections. Spectra were smoothed in the spatial direction with a width of 5 pixels in order to re- duce the noise, and the sampling rate was 3 pixels. For each smoothed spectrum, we ran a retrieval following the same procedure as in Section 5.2.2. The retrieved cloud optical thicknesses are shown in Figure 6. In the belts, the retrieved deep cloud opacity is ∼0.1 (relatively transpar- ent), and for the zones the retrieved deep cloud opacity is ∼100 (relatively opaque). The absolute values of the both cloud optical thickness are very dependent on (i) the accuracy of the radiometric calibration, (ii) the cloud locations and (iii) the choice of cloud scattering parameters. The error bars shown in Fig- ure 6 do not include these error sources. Changes to either of the first two factors result in a cloud opacity profile that is either scaled up or scaled down, but the relative shape of the latitudinal profile remains constant. The third factor can affect the relative shape of the latitudinal profile, as described in Section 5.2.4. 5.2.4. Sensitivity to the scattering parameters Section 4 showed that decreasing the asymmetry pa- rameter, g, can act to significantly increase the retrieved gaseous abundances in the cloudy parts of the planet while having a negligible effect on the abundances in the cloud- free regions. This therefore has the potential to solve the problem posed in Section 5.2.1 without having to invoke the presence of a deep cloud deck. With g = 0.8, the re- trieved CH3D abundance in the EqZ was 0.09 ppm, com- pared to 0.15 ppm in the SEB. If instead, we set g = 0.4, the EqZ abundance increases to 0.16, while the SEB abun- dance remains 0.15 ppm, removing the discrepancy be- tween the two regions. This is shown further by Figure 7, which shows the same latitudinal cloud profiles as Fig- ure 6 for the case where g = 0.4. Comparing these two figures shows that the change in scattering parameters sig- nificantly affects the latitudinal profile. The deep cloud is relatively transparent everywhere, with little difference be- tween the belts and the zones. It is therefore clear that the conclusions about the existence of a cloud deck at 5 bar are fairly sensitive to the assumptions made about the scattering properties of the cloud deck at 0.8 bar. 5.3. Summary In this section, we have shown that the shape of the CH3D 5-m absorption features varies with latitude. We initially showed that this can be modelled in two different ways; either by allowing the CH3D abundance to vary spa- tially, or by allowing the opacity of a 5-bar tropospheric 7 Figure 5: Retrievals of three CH3D features in the cool EqZ, with varying the cloud structure. The blue line shows the best fit when there is no deep cloud present, and the yellow line shows the fit where the opacity of a deep cloud at 5 bar is allowed to vary. In both cases, the CH3D abundance is fixed at 0.16 ppm. Figure 6: Retrieved cloud optical thicknesses as a function of latitude. The retrievals show that the cool zones are fit using an opaque deep cloud, while the warm belts are fit using an optically thin deep cloud. Figure 7: Retrieved cloud optical thicknesses as a function of latitude, using g = 0.4 instead of g = 0.8 for the upper cloud. 8 4.6264.6284.6304.632Wavelength (µm)01234Radiance (µWcm−2sr−1µm−1)Equatorial Zone4.6424.6444.6464.648Wavelength (µm)4.6604.6644.668Wavelength (µm)No deep cloudVariable deep cloud(a) Upper Cloud-60-40-200204060Latitude051015Optical Thickness(b) Deep Cloud-60-40-200204060Latitude10-210-1100101102103Optical Thickness(a) Upper Cloud-60-40-200204060Latitude051015Optical Thickness(b) Deep Cloud-60-40-200204060Latitude10-210-1100101102103Optical Thickness cloud to vary. We ruled out the former option on the grounds that CH3D should have a constant VMR through- out the troposphere, so we therefore accept the latter op- tion. This suggests that there is an optically thick deep cloud layer present in the EqZ. We then performed a full latitudinal retrieval, where the opacity of this deep cloud was independently retrieved at each latitude, and we found that the deep cloud is opaque in the zones and transparent in the belts. It should be noted, however, that these conclusions are sensitive to the scattering properties of the upper cloud deck. The above conclusions are based on an asymmetry parameter of g = 0.8, which was the best-fit value found in Section 4. However, if the asymmetry parameter is re- duced to g = 0.4, then the conclusions about the presence of a deep cloud are no longer valid. While g = 0.8 did produce a better fit to the data than g = 0.4, a value of 0.4 is still within a plausible range of values. We therefore conclude that while there is some evidence for the presence of a deep cloud, it is weakened by the degeneracies with the cloud scattering properties. 6. GeH4 6.1. Introduction Germane (GeH4) is a molecular species that only ex- ists in thermochemical equilibrium deep in Jupiter's at- mosphere (Taylor et al., 2004). In the 298-2000 K range, GeH4 is modelled to be the second most abundant Ge species (after GeS), but as the altitude increases and the temperature decreases, it is destroyed in favour of either GeS or GeSe (Fegley and Lodders, 1994). Because of this, GeH4 was not expected to be present in observable quan- tities in the upper troposphere at the pressure levels that are probed by the majority of the infrared spectrum. How- ever, the 5-m atmospheric window probes much deeper into the atmosphere, leading Corice and Fox (1972) to pre- dict that GeH4 absorption features might be visible in this part of the spectrum. GeH4 has two infrared active fundamental bands, ν3 (4.737 m) and ν4 (12.210 m). Fortunately, ν3 is lo- cated in the middle of the 5-m window, providing the perfect opportunity to detect the disequilibrium species. It was the Q-branch of the ν3 fundamental that was first identified as a jovian GeH4 feature by Fink et al. (1978), using observations from the Kuiper Airborne Observatory (KAO). They derived a tropospheric GeH4 abundance of 0.6 ppb. The same spectral feature was subsequently ob- served in data from the Infrared Interferometer Spectrom- eter (IRIS) on the Voyager 1 spacecraft (Kunde et al., 1982; Drossart et al., 1982). Using new observations from KAO, Bjoraker et al. (1986) identified additional GeH4 features, corresponding to the R3 and R6 transitions. In addition, they identified the Q-branch of the ν1 fundamen- tal at 4.738 m; although this band is 'forbidden' under symmetry considerations, weak absorption lines were iden- tified in laboratory spectra by Lepage et al. (1981). For these earlier analyses, the available spectroscopic data was restricted to a single isotopic species, 74GeH4. In fact, there are five naturally occurring isotopes of germa- nium: 70Ge, 72Ge, 73Ge, 74Ge and 76Ge. The terrestrial relative abundances are 20.6%, 27.5%, 7.8%, 36.5% and 7.7% respectively (Berglund and Wieser, 2011). B´ezard et al. (2002) was the first study to include line data for all five isotopic species of GeH4. They used the- oretical line data which they scaled to match observations for a single isotope, and assumed telluric relative abun- dances. Their derived VMR of 0.45 ppb was 35 -- 50% lower than previous estimates, which they attributed to the use of the additional isotopes. We take a similar approach to the GeH4 line data as B´ezard et al. (2002). We generated theoretical line lists for each isotope using the Spherical Top Data System (Wenger and Champion, 1998), and then scaled these to match the observed 74GeH4 data from the GEISA database (Jacquinet-Husson et al., 1999). We also assume telluric relative abundances. As in previous stud- ies, we assume that GeH4 is well-mixed in the troposphere. Previous studies of GeH4 have generally been restricted to a single region of the planet: either an average over the centre of the disc (Fink et al., 1978; Bjoraker et al., 1986) or a focus on the bright NEB (Kunde et al., 1982; B´ezard et al., 2002). The exception to this is Drossart et al. (1982), who found that within a factor of 2, the same GeH4 abundance was able to fit the range of spectra from 30◦S to 30◦N. We use the improved spectral resolution afforded by CRIRES to repeat this latitudinal study and to extend it to higher latitudes. 6.2. Analysis 6.2.1. Identification of GeH4 features Our first task is to identify the GeH4 absorption features in the 5-m spectra. Based on previous studies, we know that there is a strong Q-branch feature at 4.737 m. In a preliminary investigation, we determined that this ab- sorption feature was particularly clear in the bright SEB; we therefore used this spatial region in order to search for additional GeH4 absorption features. The identifiable GeH4 absorption features in the CRIRES spectra are shown in Figure 8. The NEMESIS retrieval algorithm was used to model the data, and the best-fit is shown in red. In order to highlight the shape and location of the GeH4 absorption features, the blue line shows what the spectrum would look like in the absence of GeH4. Each feature is labelled, and is comprised of a blend of the five isotopologues. The Q, R3 and R6 bands have been identified in previous studies (Fink et al., 1978; Bjoraker et al., 1986). In addi- tion, we identify the R7 line, located at 4.639 m. From each of these four spectral features, we retrieve the follow- ing GeH4 abundances: 0.51 ppb (R7), 0.62 ppb (R6), 0.45 ppb (R3), 0.58 ppb (Q). These small differences could re- flect genuine differences, due to each line probing a slightly different pressure level, or they could be due to inaccura- cies in the line data. As each segment was observed at a 9 Figure 8: Identification of GeH4 absorption features in the 5-m window. CRIRES data is shown in black and is taken from the SEB. The best fits obtained from two retrievals are shown: with GeH4 present (red) and with no GeH4 present (blue). different time, and therefore a different longitude, a simul- taneous retrieval of all four features is not possible. 6.2.2. Spatial variability Having identified GeH4 absorption features in one dis- crete region of the planet (the SEB), we now consider how this varies with latitude. We will focus on the most promi- nent GeH4 feature, the Q-branch at 4.734-4.742 m. Figure 9 shows the same absorption feature for three different regions of the planet: the EqZ, the SEB and a "Northern Region", corresponding to 50-55◦N. Compar- ing the data, shown in black, from the three regions high- lights some differences: the absorption feature appears to be deepest in the SEB and shallowest in the EqZ. This would appear to suggest that there is less GeH4 present in the EqZ. However, the previous sections have already demonstrated a degeneracy between cloud properties and retrieved gaseous abundances. Section 4 showed that the depth of an absorption feature can depend on the scat- tering properties of the main cloud deck, and Section 5 showed that it can also depend on the presence/absence of a deep cloud deck. Since the primary difference in line shape is between the EqZ and SEB, which we expect to have very different cloud structures, it is likely that clouds play an important role. In order to explore whether or not the CRIRES obser- vations provide evidence for spatial variability in GeH4, we performed a series of retrievals on these three spectra, making different assumptions about the cloud structure. The red line shows the case where there is a single upper cloud deck, with an asymmetry parameter g = 0.8. The retrieved GeH4 abundances are 0.19 ppb (EqZ), 0.58 ppb (SEB) and 0.44 ppb (Northern Region); in this case, the shallow feature in the EqZ does indeed lead to a lower retrieved abundance. asymmetry parameter of the cloud is changed to 0.4. Be- cause the SEB and Northern Region have low optical thick- nesses, this has a negligible effect of those retrievals (now 0.59 ppb and 0.47 ppb respectively). There is a significant effect for the EqZ, increasing the retrieved abundance from 0.19 ppb to 0.47 ppb. Changing the asymmetry parame- ter (i.e. making the cloud particles less forward scattering) can therefore account for much of the apparent difference between the EqZ and the SEB. Alternatively, the difference can be accounted for by the presence/absence of a deep cloud. In Section 5, we fixed the CH3D abundance to the SEB value, and we were able to fit the EqZ by adding a deeper cloud deck. We repeat the same process here: upper cloud is returned to an asym- metry parameter of 0.8, the GeH4 abundance is fixed to the SEB value (0.58 ppb) and the opacity of a deep cloud at 5 bar is allowed to vary in the retrievals. The green line in Figure 9 shows the fits obtained for this case. The retrieved deep cloud is opaque in the EqZ and transparent in the SEB, just as in Section 5. This shows that the deep cloud structure studied in Section 5 can account for the difference in line shape between the EqZ and SEB. 6.2.3. Latitudinal retrieval Section 6.2.2 showed that there is no evidence for spatial variability in GeH4 between three discrete regions of the planet, as any differences in line shape can be accounted for by changes in the cloud structure alone. In this sec- tion, we extend the analysis to cover the full range of lat- itudes available. Figure 10 shows the GeH4 abundance as a function of latitude, retrieved with different assump- tions. Alongside this is the goodness-of-fit (χ2/n), which describes the quality of the fit at each latitude point. As with Figure 9, the three colours correspond to different assumptions made in the retrievals. The yellow line in Figure 9 shows the case where the The red line corresponds to the simplest case, where 10 4.6324.6364.6404.644Wavelength (µm)0102030405060Radiance (µWcm−2sr−1µm−1)R74.6484.6524.656Wavelength (µm)R64.6844.6884.692Wavelength (µm)R34.7324.7364.7404.744Wavelength (µm)QWith GeH4Without GeH4 Figure 9: GeH4 absorption feature fits in the Equatorial Zone and South Equatorial Belt. The CRIRES oservations are shown in black. The coloured lines show the best fit that can be achieved using different assumptions. Red: single cloud deck with g = 0.8, GeH4 allowed to vary. Yellow: single cloud deck with g = 0.4, GeH4 allowed to vary. Green: upper cloud deck with g = 0.8, additional deep cloud deck, GeH4 held fixed. there is a single upper cloud deck (with g = 0.8) and the GeH4 abundance is allowed to vary, i.e. the same as the red line in Figure 9. Section 6.2.2 showed that this simple case led to a variability in the retrieved GeH4 abundance, with a depletion in the EqZ and an enhancement in the SEB. Figure 10 shows that this applies at all latitudes, and so this simple assumption leads to an apparent belt-zone variability. The black line corresponds to the case where an addi- tional deep cloud is included in the model and its latitu- dinal profile is fixed to the values retrieved in Section 5 (shown in Figure 6). It should be noted that this previ- ously determined deep cloud latitudinal profile relates to a different longitude, so it is not necessarily expected to be identical to the cloud profile at this longitude. Nev- ertheless, the overall latitudinal trend is likely to be very similar and it provides an insight into the role of the deep cloud on the retrieved abundances. Once again, the GeH4 abundance is allowed to vary in the retrievals. It can be seen in Figure 10a that the presence of this deep cloud increases the retrieved abundances in the zones, decreas- ing the apparent belt-zone variability shown by the red line. Figure 10b shows that this change does not alter the quality of the fits. For the green line, the GeH4 abundance is held fixed as a function of latitude and the opacity of the deep cloud is allowed to vary, instead of being held fixed at the values determined in Section 5. This is the same as the green lines in Figure 9. As in Section 6.2.2, the GeH4 abundance cho- sen is the best-fit value from the SEB (0.58 ppb). Since the SEB is relatively cloud-free, there are no complica- tions from the cloud structure and this should represent the 'true' abundance. The goodness-of-fit in Figure 10 shows that this constant abundance is able to provide a similar fit to the data at all latitudes. Section 6.2.2 pre- viously showed that variability in the deep cloud opacity could account for the difference in line shape between three discrete regions of the planet, and Figure 10 shows that 11 Figure 10: GeH4 abundance and goodness-of-fit values as a function of latitude, for three different assumptions. Red: single cloud deck with g = 0.8, GeH4 allowed to vary. Black: additional deep cloud deck with opacities derived from Section 5, GeH4 allowed to vary. Green: deep cloud deck allowed to vary, GeH4 held fixed at 0.58 ppb. (a) Equatorial Zone4.7304.7354.7404.745Wavelength (µm)0.00.51.01.52.02.53.03.5Radiance (µWcm−2sr−1µm−1)(b) South Equatorial Belt4.7304.7354.7404.745Wavelength (µm)01020304050Radiance (µWcm−2sr−1µm−1)(c) Northern Region4.7304.7354.7404.745Wavelength (µm)0246810Radiance (µWcm−2sr−1µm−1)(a) GeH4 abundance-90-60-300306090Latitude0.00.20.40.60.8V.M.R (ppb)(b) Goodness−of−fit−90−60−300306090Latitude0246810χ2/n this remains true across the full range of latitudes. 6.3. Summary In this section, we have identified four GeH4 absorption features in the CRIRES observations, one of which has not been previously identified. Using the strongest absorption feature, we searched for latitudinal variability and found that there are some differences in the line depth between the EqZ, the SEB and the Northern Region of the planet. However, this difference in line shape does not necessar- ily indicate a difference in the abundance of GeH4. We found that the variation in line shape could be accounted for by Jupiter's tropospheric cloud structure. This anal- ysis was then extended to all latitudes, and it was found that a single fixed GeH4 abundance could be used to fit all spectra without compromising the quality of the fits. The CRIRES observations therefore do not provide any evidence for spatial variability in GeH4. This highlights the high level of degeneracy present in the problem. The deep cloud structure, which has not been included in an any previous studies of GeH4, significantly complicates the analysis. 7. AsH3 7.1. Introduction Arsine (AsH3) is the most abundant arsenic gas on Jupiter. Deep in the atmosphere it exists in thermochem- ical equilibrium but, in the absence of vertical motion, it precipitates into As4(s) or As2S2(s) at ∼400 K (Fegley and Lodders, 1994). As with GeH4, this means that the 5-m atmospheric window is an ideal spectral region to search for the species. AsH3 has two fundamental bands within the 5-m win- dow: ν1 at 4.728 m and ν3 at 4.704 m. Of these, the ν3 band is stronger and its Q-branch led to the first detec- tion of AsH3 on Jupiter by Noll et al. (1989), using spec- tra obtained from the United Kingdom Infrared Telescope. They initially estimated the deep molar abundance to be 0.7+0.7−0.4 ppb, which was then revised down to 0.22 ± 0.11 ppb the following year, after acquisition of new airborne and ground-based spectra and additional laboratory line data (Noll et al., 1990). More recently, B´ezard et al. (2002) used data from the Canada-France-Hawaii Telescope to ob- tain an abundance of 0.24 ppb. The Q-branch of the ν3 band remains the only AsH3 feature to have been identi- fied on Jupiter; we will search for evidence of additional AsH3 absorption lines. As with GeH4, we assume that AsH3 is well-mixed in the troposphere. Out of the previous studies of AsH3, only Noll et al. (1990) studied more than one region of the planet. They made observations of both the EqZ and the NEB and found that these spectra matched to within the noise levels. As the CRIRES observations provide both high spectral res- olution and high spatial resolution, we will extend this analysis to cover all latitudes. 7.2. Analysis 7.2.1. Identification of AsH3 features As with GeH4, we first seek to simply identify the AsH3 absorption features in the 5-m spectrum. The only pre- viously identified feature is the Q-branch at 4.704 m; in a preliminary investigation we determined that this feature was particularly strong at high latitudes (50-55◦N) so we use this 'Northern Region' in order to search for additional absorption lines. We were able to identify two additional absorption fea- tures, corresponding to the R1 and P6 lines of the ν3 band (Dana et al., 1993). These lines are shown in Fig- ure 11, alongside the Q-branch. The CRIRES data is shown in black. The blue lines show the best fit that can be achieved when no AsH3 is present in the atmosphere and the red lines show the best fit that can be achieved when the AsH3 abundance is allowed to vary. As in Sec- tion 6.2.1, the cloud opacity of the single cloud deck and the gaseous abundances were also allowed to vary. In each case, the addition of AsH3 significantly improves the fit. We obtain the following abundances from each retrieval: 0.60 ppb (R1), 0.49 ppb (Q) and 0.46 ppb (P6). 7.2.2. Spatial variability We now consider how the AsH3 lineshape varies with lat- itude, and how this affects the retrieved abundances. As in previous studies, we focus on the most prominent AsH3 absorption feature in the 5-m window, the Q-branch of the ν3 fundamental, located at 4.704 m. Figure 12 shows this spectral region for the three regions of the planet con- sidered in Section 6. Unlike the previous GeH4 section, the EqZ and the SEB have similar spectral shapes. In this case, however, there is significant variation between these spectra and the spectrum from the Northern Region. The absorption feature at 4.704 m is considerably flatter in Northern Region than it is in the SEB and the EqZ, sug- gesting a higher abundance of AsH3 at high latitudes. This is confirmed by a simple NEMESIS retrieval, shown in red in Figure 12. This fit is analogous to the red line in Figure 9, where the cloud structure consists of a single upper cloud with g = 0.8. With these assumptions, we obtain the following retrieved abundances for AsH3: 0.01 ppb in the EqZ, 0.02 ppb in the SEB and 0.49 ppb in the Northern Region. In the case of the EqZ and the SEB, the retrieved abundances are so low that they are essentially zero. In Section 6, we found that the apparent change in GeH4 lineshape could be accounted for by variations in the cloud structure, rather than changes in the GeH4 abun- dance itself. However, clouds are unlikely to be responsi- ble here, since the EqZ and the SEB have very different cloud structures, and yet the AsH3 features have similar spectral shapes. The observed effect is also unlikely to be due to limb darkening at high altitudes. Firstly, previous studies have shown that Jupiter's tropospheric clouds are highly scattering, which minimises the effect of limb dark- ening (Roos-Serote and Irwin, 2006; Giles et al., 2015). 12 Figure 11: Identification of AsH3 absorption features in the 5-m window. CRIRES data is shown in black, and is taken from the Northern Region of the planet. The best fits obtained from two retrievals are shown: with AsH3 present (red) and with no AsH3 present (blue). Figure 12: AsH3 absorption feature fits at different latitudes. The CRIRES observations are shown in black. The coloured lines show the best fit that can be achieved using different assumptions. Red: single cloud deck with g = 0.8, AsH3 allowed to vary. Yellow: single cloud deck with g = 0.4, AsH3 allowed to vary. Green: upper cloud deck with g = 0.8, additional deep cloud deck, AsH3 held fixed. 13 4.6644.6684.672Wavelength (µm)051015Radiance (µWcm−2sr−1µm−1)R14.6924.6964.7004.704Wavelength (µm)Q4.8004.804Wavelength (µm)P6With AsH3Without AsH3(a) Equatorial Zone4.6854.6904.6954.7004.705Wavelength (µm)012345Radiance (µWcm−2sr−1µm−1)(b) South Equatorial Belt4.6854.6904.6954.7004.705Wavelength (µm)0102030405060Radiance (µWcm−2sr−1µm−1)(c) Northern Region4.6854.6904.6954.7004.705Wavelength (µm)02468101214Radiance (µWcm−2sr−1µm−1) Secondly, even if there was a significant limb darkening effect, we would to see the opposite phenomenon. As we move to higher latitudes, the emission angle increases, and we therefore probe higher in the atmosphere, where (as a disequilibrium species) AsH3 is less abundant, not more. Nevertheless, in order to fully explore any degeneracy, we repeated the different retrievals performed in Section 6. The yellow line in Figure 12 shows the fits that are ob- tained when the asymmetry parameter of the upper cloud is changed from 0.8 to 0.4. The retrieved abundance in the EqZ is increased from 0.01 ppb to 0.05 ppb, and the SEB and Northern abundances remain essentially unchanged at 0.02 ppb and 0.50 ppb respectively. We are unable to close the gap between the equatorial regions and the high lat- itudes by changing the scattering properties of the upper cloud. We then consider the effect of a deep cloud. The AsH3 were held constant at the SEB level (0.02 ppb) and the opacity of a deep cloud was allowed to vary. The results are shown by the green line in Figure 12. As in Section 6, the EqZ and SEB can be fit using the same AsH3 abun- dance, provided that the deep cloud is optically thick in the EqZ. However, it is clear from Figure 12(c) that this constant abundance cannot be applied to the Northern Re- gion of the planet. We therefore conclude that the spatial variation in the AsH3 lineshape is indicative of genuine latitudinal variability in the abundance of AsH3. 7.2.3. Latitudinal retrieval Having established that there is evidence for spatial vari- ability in AsH3, we now extend the analysis from three dis- crete sections to a full latitudinal retrieval of AsH3. Fig- ure 13 shows the latitudinal distribution of AsH3 along- side the goodness-of-fit values. The three different colours correspond to the same assumptions described in Sec- tion 6.2.3. For the latitudinal distribution shown by the red line, there is a single upper cloud deck with g = 0.8 and the AsH3 abundance is allowed to vary, i.e. the same as the red line in Figure 12. Section 7.2.2 showed that this simple case produced similar retrieved abundances in the EqZ and SEB and an enhanced abundance in the Northern Region. This pattern is confirmed by the full latitudinal retrieval; there is a roughly symmetrical distribution, with an in- creased abundance at high latitudes. The black line shows the case where an additional deep cloud is included in the model, with the latitudinal pro- file obtained from Section 5. While the precise values of the retrieved abundances do change slightly with the in- clusion of this additional cloud deck, the overall trend of an enhancement an high latitudes remains true. The er- rors shown in Figure 13 represent the formal errors on the retrieval; the difference between the red and black lines shows that these formal errors are significantly smaller than the errors due to varying assumptions about cloud structure. As with Section 6.2.3 for GeH4, the inclusion of an additional cloud does not alter the quality of the fits. Figure 13: AsH3 abundance and goodness-of-fit values as a function of latitude, for three different assumptions. Red: single cloud deck with g = 0.8, AsH3 allowed to vary. Black: additional deep cloud deck with opacities derived from Section 5, AsH3 allowed to vary. Green: deep cloud deck allowed to vary, AsH3 held fixed at 0.02 ppb. This is in contrast to the final case, shown by the green line. In this case, the AsH3 abundance is held fixed at the SEB value of 0.02 ppb and the deep cloud is allowed to vary. Section 7.2.2 showed that this can provide a good fit in the EqZ and SEB, but fails to provide a good fit for the Northern Region. Figure 13 extends this to all lat- itudes and shows that this constant abundance can provide a good fit between 40◦S and 40◦N, i.e. there is no evidence for belt-zone variability. However, the fit worsens consid- erably at high latitudes; an abundance of 0.02 ppb simply cannot fit these regions, regardless of the opacity of the deep cloud. This leads to the conclusion that there is a genuine enhancement at high latitudes. Figure 13 is restricted to the data from 12 November 2012. However, Figure 14 shows that the 1 January 2013 dataset produces consistent results, and that the overall conclusion of an AsH3 enhancement at high latitudes is reproducible. 7.3. Summary In this section, we have identified three AsH3 absorp- tion features in the CRIRES observations, two of which have not been previously identified. Using the strongest absorption feature, we searched for latitudinal variability and found that there are significant differences in the line shape between spectra from the equatorial regions (EqZ and SEB) and spectra from the far northern latitudes. These differences cannot be accounted for by Jupiter's cloud structure, and we therefore conclude that there is ev- idence for latitudinal variability in the abundance of AsH3. 14 (a) AsH3 abundance-90-60-300306090Latitude0.00.20.40.6V.M.R (ppb)(b) Goodness−of−fit−90−60−300306090Latitude051015χ2/n profile via a single scaling parameter; because the CRIRES data is from the 5-m region, we are probing the deep abundance. Several previous studies have also considered the pos- sibility of spatial variability in the deep PH3 abundance, as observed in the 5-m window. Drossart et al. (1982) and Lellouch et al. (1989) searched for a correlation be- tween cloud opacity and PH3 abundance, but did not find any evidence. More recently, Drossart et al. (1990) and Giles et al. (2015) found evidence for a PH3 enhance- ment at high latitudes compared to equatorial- and mid- latitudes. The spatially resolved 10-m studies show a dif- ferent pattern; the retrieved fractional scale height shows a global maximum over the equator and global minima at the NEB and SEB (Irwin et al., 2004; Fletcher et al., 2009). 8.2. Analysis 8.2.1. Spatial variability For GeH4 and AsH3, the choice of spectral region was clear, as there were very few absorption features in the 5-m window. In contrast, PH3 is one of the primary absorbers at 5-m, and there are many absorption fea- tures. In particular, the short-wavelength part of the 5-m spectrum (4.50-5.00 m) is dominated by PH3 absorption. This might sound ideal, but in fact this prevents the accu- rate retrieval of PH3, as it becomes difficult to distinguish between PH3 and other broad features such as the cloud opacity and the water vapour abundance. Instead, we look towards the long-wavelength edge, where PH3 absorption contributes less to the continuum. We find that there is a clear, well-separated absorption feature at 5.071 m and we will use this as the basis of our study of the deep PH3 abundance in the troposphere. Figure 15 shows this absorption feature for the same three spatial regions used in Sections 6 and 7. The CRIRES spectra in Figure 15 show variation in the depth of the PH3 absorption feature; the feature is considerably deeper in the Northern Region than in the EqZ, while the SEB spectrum is between these two extremes. The coloured lines show the best fit obtained for each spectrum, with the same assumptions as in Sections 6 and 7: the red line shows the case where there is a sin- gle upper cloud with g = 0.8 and the yellow line shows the case where there is a single upper cloud with g = 0.4. In the first case (red), the retrieved abundances in the EqZ, SEB and Northern Region are 0.49 ppm, 0.76 ppm and 1.22 ppm respectively. In the second case (yellow), these become 0.81 ppm, 0.75 ppm and 1.21 ppm. As in Section 7, altering the scattering properties of the cloud can account for the discrepancy between the EqZ and the SEB but not between the equatorial regions and the polar regions. The green line in Figure 15 shows the case where the upper cloud has g = 0.8, the PH3 abundance is fixed to the SEB value and the opacity of a second deep cloud Figure 14: AsH3 abundance as a function of latitude from the 12 November 2012 dataset (black) and the 1 January 2013 dataset (red). In both cases, a single cloud deck with g = 0.8 is assumed and the AsH3 abundance is allowed to vary. The two datasets produce consistent results. This was further shown by a latitudinal retrieval, which demonstrated that a constant AsH3 abundance is incon- sistent with the observations. Instead, the retrieved AsH3 abundances have a roughly symmetrical latitudinal profile, with a minimum in the equatorial regions and enhanced abundances at high latitudes. 8. PH3 8.1. Introduction Like GeH4 and AsH3, phosphine (PH3) is a disequilib- rium species in Jupiter's troposphere. Deep in the at- mosphere it is the primary phosphorus gas but at higher altitudes thermal decomposition produces PH, which then reacts with OH radicals to form P4O6 (Fegley and Lodders, 1994). However, the solar abundance of P is much greater than the solar abundances of Ge or As (Grevesse et al., 2007). This means that even though the VMR of PH3 de- creases rapidly with altitude, it is still present in observ- able quantities in the upper troposphere and is therefore detectable outside the 5-m window. Strong PH3 absorption features can be seen at ∼ 5 m and at ∼10 m. PH3 was simultaneously detected in both of these spectral regions. Ridgway et al. (1976) noted that including a solar abundance of PH3 in their model im- proved their fit to 10-m observations of Jupiter from the McMath Solar Observatory. This was then confirmed by Larson et al. (1977), who also found that a roughly so- lar abundance of PH3 was able to fit their 5-m airborne observations. The 5-m and 10-m spectral regions probe different pressure levels in Jupiter's troposphere. At 5-m, the spec- trum is primarily sensitive to the lower troposphere (p>1 bar) whereas the 10-m observations are mostly sensitive to the upper troposphere, at pressures p<1 bar. This pres- sure difference allows us to start to constrain the vertical profile of PH3. Fletcher et al. (2009) modelled Jupiter's PH3 abundance using a constant abundance of 1.86 ppm up to 1 bar, above which it drops off with a constant frac- tional scale height of 0.3. In this section, we will vary this 15 AsH3 abundance-90-60-300306090Latitude0.00.20.40.6V.M.R (ppb) Figure 15: PH3 feature fits at different latitudes. The CRIRES observations are shown in black. The coloured lines show the best fit that can be achieved using different assumptions. Red: single cloud deck with g = 0.8, PH3 allowed to vary. Yellow: single cloud deck with g = 0.4, PH3 allowed to vary. Green: upper cloud deck with g = 0.8, additional deep cloud deck, PH3 held fixed. is allowed to vary. Again, the results are the same as for AsH3 in Section 7. The presence of an optically thick deep cloud in the EqZ can account for the change in spectral shape between the EqZ and the SEB, but the same PH3 abundance cannot be used to fit the regions at high lat- itude. We conclude that there is evidence for latitudinal variability in PH3. 8.2.2. Latitudinal retrieval Having previously focussed on three discrete sections of the planet, we now turn to a full latitudinal retrieval of PH3. he results of these retrievals are shown in Figure 16; these results take the same form as Figure 10 (for GeH4) and Figure 13 (for AsH3) and the same conclusions can be drawn for PH3 as for AsH3. As with AsH3, the black and red lines in Figure 16 both show a polar enhancement in PH3, whether or not the additional deep cloud from Section 5 is included in the model. The case where the PH3 abundance is held fixed at 0.76 ppm is shown in green, and this also follows the same pattern as AsH3; at low latitudes, this model provides a good fit (i.e. there is no evidence for belt-zone variability), but the fit starts to worsen at high latitudes (in this case, above ∼60◦N). This agrees with the conclusion made in Section 8.2.1 that there in evidence for an enhancement in PH3 at high latitudes, with the retrieved abundance varying from ∼0.5 ppm near the equator to 1.5 ppm at 80◦N. Figure 16 is restricted to the data from 12 November 2012. However, as with AsH3, the 1 January 2013 dataset produces consistent results (see Figure 17), showing that the overall conclusion of an PH3 enhancement at high lat- itudes is reproducible. 8.3. Summary In this section, we considered how the strength of a well-isolated PH3 absorption line varied with latitude. We found that the depth of the absorption feature was greater 16 Figure 16: PH3 abundance and goodness-of-fit values as a function of latitude, for three different assumptions. Red: single cloud deck with g = 0.8, PH3 allowed to vary. Black: additional deep cloud deck with opacities derived from Sec- tion 5, PH3 allowed to vary. Green: deep cloud deck al- lowed to vary, PH3 held fixed at 0.76 ppm. (a) Equatorial Zone5.0655.0705.0755.080Wavelength (µm)01234Radiance (µWcm−2sr−1µm−1)(b) South Equatorial Belt5.0655.0705.0755.080Wavelength (µm)0102030405060Radiance (µWcm−2sr−1µm−1)(c) Northern Region5.0655.0705.0755.080Wavelength (µm)02468101214Radiance (µWcm−2sr−1µm−1)(a) PH3 abundance-90-60-300306090Latitude0.00.51.01.5V.M.R (ppm)(b) Goodness−of−fit−90−60−300306090Latitude012345χ2/n robustness of this conclusion. In the future, this could be improved by simultaneous modelling of the 5-m region and the near-infrared reflected sunlight spectrum. Both of these degeneracies complicate the retrievals of other molecular species. Since the cloud structure varies between belts and zones, these degeneracies especially limit our ability to come to reliable conclusions about belt- zone variability in gaseous abundances. This is further demonstrated in Section 6, where we showed that we were unable to distinguish between variability in GeH4 abun- dance and variability in clouds. 9.2. Disequilibrium species 9.2.1. Abundances In Sections 6-8, we measured the abundances of three disequilibrium species in Jupiter's troposphere: GeH4, AsH3 and PH3. For GeH4, we initially retrieved abun- dances that varied between 0.19 ppb (EqZ) and 0.58 ppb (SEB). By including a variable deep cloud, we were able to fit all spatial regions using a tropospheric abundance of 0.58 ppb. This is similar to previous results, which range from 0.45 ppb (B´ezard et al., 2002) to 1.0 ppb (Drossart et al., 1982). This is significantly lower than the solar abundance of 6.7 ppb (Grevesse et al., 2007), but this is unsurprising as GeS, rather than GeH4, is the most abun- dant Ge species (Fegley and Lodders, 1994). Unlike GeH4, measurements of the AsH3 absorption fea- ture showed evidence for an enhancement at high latitudes. Using a one-cloud model, the retrieved abundance varies from as low as 0.01 ppb in parts of the equatorial region to 0.55 ppb at 80◦N. Previous measurements of the AsH3 abundance have relied exclusively on data from the low latitudes of the planet, and have led to estimates of 0.2- 0.3 ppb (B´ezard et al., 1989; Noll et al., 1990; B´ezard et al., 2002). Wang et al. (2016) notes that AsH3 is expected to be the primary As gas in the troposphere, and that these previous studies therefore suggest that the jovian abun- dance of As is lower than the solar value (a solar abundance would lead to a volume mixing ratio of 0.34 ppb, Grevesse et al., 2007). This is surprising, as other heavy elements show an enrichment relative to solar abundances (e.g. Tay- lor et al., 2004). Our results show that in some parts of the planet, the observed AsH3 abundance is subsolar, while in other regions, AsH3 shows an enrichment of up to 1.6 times the solar value. It is possible that this higher abundance is more representative of the deep abundance of As. As with AsH3, the retrieved PH3 abundance varies with latitude, ranging from 0.5 ppm in the equatorial regions to 1.4 ppm at 80◦N. These values are consistent with previous 5-m observations of PH3: 0.7 ppm (Bjoraker et al., 1986), 0.77 ppm (Irwin et al., 1998) and 0.76-0.90 ppm (Giles et al., 2015). As noted in Giles et al. (2015), measurements at 5-m appear to consistently lead to lower retrieved abundances than measurements at 10-m (e.g. Fletcher et al., 2009); these differences are potentially due to errors in line data or cloud modelling uncertainties. Assuming a Figure 17: PH3 abundance as a function of latitude from the 12 November 2012 dataset (black) and the 1 January 2013 dataset (red). In both cases, a single cloud deck with g = 0.8 is assumed and the AsH3 abundance is allowed to vary. The two datasets produce consistent results. for the high northern latitudes than for the equatorial lat- itudes. As with AsH3, we concluded that this was due to higher PH3 abundances at high latitudes. We then per- formed a latitudinal retrieval, and found that this showed a minimum near the equator and an enhancement at high latitudes. 9. Discussion 9.1. Cloud structure and degeneracies In Sections 4 and 5 we showed that there are signif- icant degeneracies between Jupiter's tropospheric clouds and the gaseous composition. Firstly, Section 4 showed that the retrieved abundances are sensitive to the cloud scattering properties; a high asymmetry parameter (more forward scattering, less reflected sunlight) leads to a low abundance, and a low asymmetry parameter (less forward scattering, more reflective sunlight) leads to a high abun- dance. The scattering properties also control the strength of the Fraunhofer lines in the spectrum, and by looking at the strength of these lines in the observations, we estimate that a value of g = 0.8 provides the best fit to the data. However, lower values can also provide an adequate fit, meaning that the degeneracy is not entirely broken. Secondly, Section 5 showed that the deep cloud struc- ture is also degenerate with the retrieved gaseous abun- dances. Increasing the optical thickness of a deep cloud has a similar effect on the spectrum as decreasing the molec- ular abundance. An optically thin deep cloud and a low abundance can produce a similar spectral shape to an op- tically thick cloud and a high abundance. In the case of CH3D, this proves useful. The EqZ and SEB have differ- ent spectral shapes, yet we expect the CH3D abundance to be constant. As described in Bjoraker et al. (2015), the presence of a cloud at 5 bar that is optically thick in the zones and optically thin in the belts can solve this 'prob- lem'. These observations therefore provide some evidence for the existence of a deep cloud; based on the location, this cloud is likely to be composed of water ice. However, the degeneracy with the scattering properties limits the 17 PH3 abundance-90-60-300306090Latitude0.00.51.01.5V.M.R (ppm) solar abundance of P would lead to a volume mixing ratio of 0.43 ppm (Grevesse et al., 2007), so our analysis shows a variation from roughly solar levels to an enrichment of up to 3.3 times more the solar value, which is consistent with the observed enhancement of other heavy elements in Jupiter's atmosphere. 9.2.2. Latitudinal profiles In Sections 6-8, we studied the latitudinal distributions of GeH4, AsH3 and PH3. We found no evidence for spatial variability in the abundance of GeH4. This result for GeH4 agrees with the previous study of Drossart et al. (1982) who found no evidence for variability in the 30◦S to 30◦N region, and our analysis shows that this remains true at higher latitudes. This is in contrast to the results for AsH3 and PH3. While there is no evidence for any belt/zone variability, both of these species exhibit an enhancement towards the poles. An enhancement at high latitudes has previously been reported for PH3 (Drossart et al., 1990; Giles et al., 2015), but this is the first time that it has been seen in AsH3, as the only previous studies to compare different regions of the planet focussed exclusively on the equatorial latitudes (Noll et al., 1990). Our results contrast with the theoretical predictions In their work, they use a from Wang et al. (2016). diffusion-kinetics code to investigate how the abundances of disequilibrium species in Jupiter's troposphere depend on the vertical eddy diffusion coefficient, K eddy. They find that GeH4 has a different behaviour to AsH3 and PH3, which agrees with our observations. The GeH4 abundance is found to be very sensitive to the eddy diffusion coeffi- cient, with a higher value of K eddy leading to a higher tro- pospheric abundance. In contrast, PH3 and AsH3 do not appear to have any dependence on K eddy (for a physically plausible range of K eddy). These differences are due to the interplay between the quench level for the gas (the pressure at which the abundance is "frozen in") and the equilibrium vertical profile. If the quench level is deep in the regime where the gas is well-mixed, then a small change in K eddy will not alter the observed abundance. If the quench level is in a regime where the equilibrium abundance is rapidly changing with altitude, then a small change in K eddy can have a significant effect on the observed abundance. Wang et al. (2016) considers how abundances vary as a function of K eddy, while a previous study by the same au- thors (Wang et al., 2015) looks at how K eddy varies with latitude. In Wang et al. (2015), they suggest that rota- tion plays an increasingly important role in suppressing turbulent convection at higher latitudes and that K eddy should decrease towards the poles. This also agrees with earlier theoretical work by Flasar and Gierasch (1978). Incorporating this finding, Wang et al. (2016) therefore predict that GeH4 should decrease in abundance towards the poles, while AsH3 and PH3 should remain constant. In contrast, our observations show that GeH4 is constant with latitude, while AsH3 and PH3 increase towards the poles. It is unclear what the cause of this discrepancy is. One possibility is that it is due to photolytic destruction, which was not considered in Wang et al. (2016). Many stud- ies have concluded that ultraviolet photons are a signif- icant contribution to the destruction of PH3 in the up- per troposphere (e.g. Strobel, 1977; Ferris et al., 1984). It has also been suggested as a destruction mechanism for GeH4 (Guillemin et al., 1995) and could plausibly apply to AsH3 too. Geometry effects mean that higher latitudes re- ceive fewer UV photons from the sun, and the higher opac- ity of stratospheric aerosols in the polar regions also acts as a shield to UV light (Zhang et al., 2013). Both of these effects could lead to a lower rate of photolytic destruction at high latitudes, and therefore a higher abundance. By increasing the abundances of all three species at high lati- tudes, the latitudinal profiles from Wang et al. (2016) can be modified to match our observations. However, the rea- son that photolysis was neglected by Wang et al. (2016) is that it is expected to primarily affect the upper tropo- sphere (p<0.5 bar), not the deeper regions probed in the 5-m region (Strobel, 1977). Alternatively, it is possible that the vertical mixing strength does not in fact decrease towards the poles. Wang et al. (2015) and Flasar and Gierasch (1978) focus on small-scale turbulent convection, and how this is affected by the increasing importance of rotation at high latitudes. While turbulent convection is inhibited towards the poles, it is possible that planetary-scale vertical motion increases. Solar heating is at a minimum at the poles, which reduces the static stability of the atmosphere and convection be- comes uninhibited (Irwin, 2003). However, Wang et al. (2016) showed that PH3 and AsH3 are not sensitive to the strength of vertical mixing, while GeH4 is. Even if the trend reported in Wang et al. (2015) and Flasar and Gierasch (1978) were to be reversed entirely, we would ex- pect to see an increase of GeH4 at the poles and a flat latitudinal distribution of AsH3 and PH3, which is not what we observe. It is also possible that the sensitivities of the disequilib- rium species to K eddy, as described in Wang et al. (2015), are not robust. In several cases, the authors note a lack of available kinetic data which limit the confidence in the results. However, these diffusion-kinetics calculations do provide a useful explanation for the difference in behaviour between GeH4 and AsH3/PH3. In addition, changes to the chemistry model cannot single-handedly account for the discrepancy between the theory and the observations. If AsH3 and PH3 were more sensitive to K eddy, they would be predicted to decrease, not increase, towards the poles, since K eddy decreases at high latitudes (Wang et al., 2015). Finally, there is the possibility that our observational conclusions are incorrect. We have shown that there are several degeneracies which complicate the retrievals of gaseous abundances in Jupiter's troposphere. Neverthe- less, we are confident in the overall latitudinal trends that we report. The differences between the equatorial latitudes and the high latitudes are extremely apparent in the raw 18 observations themselves, even before we begin to model the spectra (Figures 12 and 15). 10. Conclusions The CRIRES instrument on the VLT was used to make high-resolution (R=96,000) observations of Jupiter in the 5-m window, where the planet's atmosphere is optically thin. This enabled us to spectrally resolve the line shapes of four minor species in Jupiters troposphere: CH3D, GeH4, AsH3 and PH3. The slit was aligned north-south along Jupiter's central meridian, allowing us to search for latitudinal variability in these line shapes. The spectra were analysed using the NEMESIS radiative transfer code and retrieval algorithm. We make the following conclu- sions: CRIRES observations were limited to central-meridian scans of the planet. Future work to explore Jupiter's deep atmosphere must utilise spatially-resolved spectral map- ping to relate the dynamics of distinct atmospheric fea- tures to the distributions of these disequilibrium species. Acknowledgements Giles was supported via a Royal Society studentship, and Fletcher was supported via a Royal Society Research Fellowship at the University of Leicester. Irwin acknowl- edges the support of the United Kingdom Science and Technology Facilities Council. This work is based on ob- servations collected at the European Organisation for As- tronomical Research in the Southern Hemisphere under ESO programme 090.C-0053(A). 1. There are degeneracies between the cloud structure and scattering properties that complicate the re- trievals of tropospheric gaseous abundances. These degeneracies can sometimes explain apparent belt- zone variations in composition. Any future stud- ies into belt-zone variability must therefore carefully assess the degeneracies with the assumed scattering properties before any conclusions about spatial varia- tions in gaseous abundances can be drawn. 2. As seen in Bjoraker et al. (2015), there is some evi- dence for the existence of a highly variable cloud deck at ∼5 bar, but this conclusion is limited by degenera- cies with cloud scattering properties. 3. We are able to fit all spatial regions with a GeH4 abun- dance of 0.58 ppb, which is consistent with previous studies. We find that the AsH3 abundance varies be- tween 0.01 ppb and 0.55 ppb from equator to pole. This is the first time that supersolar abundances of AsH3 have been measured in Jupiter's atmosphere. We measure a PH3 abundance that varies between 0.5 and 1.5 ppm, which is consistent with previous measurements. 4. We found that GeH4 appears to have an abun- dance that is constant with latitude, while PH3 and AsH3 show evidence for an enhancement at high lat- itudes. This partially agrees with a theoretical study from Wang et al. (2016) which suggests that GeH4 should have a different latitudinal profile to PH3 and AsH3. However, our observations disagree with the shapes of those profiles, as they predict that GeH4 should decrease towards the poles while PH3 and AsH3 are flat. The explanation for this difference is not clear, but it could be due to the effect of pho- tolytic destruction, increased global-scale convection at the poles, missing chemical kinetic data, or a com- bination thereof. This study has demonstrated the utility of high- resolution 5-m observations in studying Jupiter's mid- tropospheric composition, revealing the latitudinal distri- butions of arsine and germane for the first time. The References References Ballester, P., Banse, K., Castro, S., Hanuschik, R., Hook, R., Izzo, C., Jung, Y., Kaufer, A., Larsen, J., Licha, T., et al., 2006. Data reduction pipelines for the Very Large Telescope, in: Proc. SPIE. Berglund, M., Wieser, M.E., 2011. Isotopic compositions of the ele- ments 2009 (IUPAC Technical Report). Pure and Applied Chem- istry 83, 397 -- 410. B´ezard, B., Drossart, P., Lellouch, E., Tarrago, G., Maillard, J., 1989. Detection of arsine in Saturn. The Astrophysical Journal 346, 509 -- 513. B´ezard, B., Lellouch, E., Strobel, D., Maillard, J.P., Drossart, P., 2002. Carbon monoxide on Jupiter: evidence for both internal and external sources. Icarus 159, 95 -- 111. Bjoraker, G., Wong, M., de Pater, I., ´Ad´amkovics, M., 2015. Jupiter's deep cloud structure revealed using keck observations of spectrally resolved line shapes. The Astrophysical Journal 810, 122. Bjoraker, G.L., Larson, H.P., Kunde, V.G., 1986. The gas composi- tion of Jupiter derived from 5-m airborne spectroscopic observa- tions. Icarus 66, 579 -- 609. Borysow, A., 1991. Modeling of collision-induced infrared absorption spectra of H2-H2 pairs in the fundamental band at temperatures from 20 to 300 K. Icarus 92, 273 -- 279. Borysow, A., Frommhold, L., 1986. Theoretical collision-induced rototranslational absorption spectra for the outer planets: H2- CH4 pairs. The Astrophysical Journal 304, 849 -- 865. Borysow, A., Frommhold, L., 1987. Collision-induced rototransla- tional absorption spectra of CH4-CH4 pairs at temperatures from 50 to 300 K. The Astrophysical Journal 318, 940 -- 943. Borysow, J., Frommhold, L., Birnbaum, G., 1988. Collison-induced rototranslational absorption spectra of H2-He pairs at tempera- tures from 40 to 3000 K. The Astrophysical Journal 326, 509 -- 515. Corice, R.J., Fox, K., 1972. The hypothetical chemical and spectro- scopic activity of germane in the atmosphere of Jupiter. Icarus 16, 388 -- 391. Cutri, R., Skrutskie, M., Van Dyk, S., Beichman, C., Carpenter, J., Chester, T., Cambresy, L., Evans, T., Fowler, J., Gizis, J., et al., 2003. 2MASS all sky catalog of point sources. . Dana, V., Mandin, J., Tarrago, G., Olson, W., Bezard, B., 1993. Absolute infrared intensities in the fundamentals ν1 and ν3 of arsine. Journal of molecular spectroscopy 159, 468 -- 480. Drossart, P., Encrenaz, T., Kunde, V., Hanel, R., Combes, M., 1982. An estimate of the PH3, CH3D, and GeH4 abundances on Jupiter from the Voyager IRIS data at 4.5 m. Icarus 49, 416 -- 426. Drossart, P., Lellouch, E., B´ezard, B., Maillard, J.P., Tarrago, G., 1990. Jupiter: Evidence for a phosphine enhancement at high northern latitudes. Icarus 83, 248 -- 253. 19 Lepage, P., Champion, J.P., Robiette, A.G., 1981. Analysis of the ν3 and ν1 infrared bands of GeH4. Journal of Molecular Spectroscopy 89, 440 -- 448. Lewis, J.S., Fegley, M.B., 1984. Vertical distribution of disequilib- rium species in Jupiter's troposphere. Space Science Reviews 39, 163 -- 192. Mandin, J., Dana, V., Tarrago, G., Klee, S., Winnewisser, B., 1995. Line positions and intensities in the vibrational system 2ν2/ν2+ν4/2ν4/ν1/ν3 of arsine. Journal of Molecular Spec- troscopy 172, 319 -- 329. Noll, K.S., Geballe, T., Knacke, R., 1989. Arsine in Saturn and Jupiter. The Astrophysical Journal 338, L71 -- L74. Noll, K.S., Larson, H.P., Geballe, T., 1990. The abundance of AsH3 in Jupiter. Icarus 83, 494 -- 499. Plass, G.N., Kattawar, G.W., Catchings, F.E., 1973. Matrix operator theory of radiative transfer. 1: Rayleigh scattering. Applied Optics 12, 314 -- 329. Ridgway, S., Wallace, L., Smith, G., 1976. The 800-1200 inverse centimeter absorption spectrum of Jupiter. The Astrophysical Journal 207, 1002 -- 1006. Roos-Serote, M., Irwin, P., 2006. Scattering properties and loca- tion of the Jovian 5-micron absorber from Galileo/NIMS limb- darkening observations. Journal of Quantitative Spectroscopy and Radiative Transfer 101, 448 -- 461. Rothman, L.S., Gordon, I.E., Barbe, A., Benner, D.C., Bernath, P.F., Birk, M., Boudon, V., Brown, L.R., Campargue, A., Cham- pion, J.P., et al., 2009. The HITRAN 2008 molecular spectroscopic database. Journal of Quantitative Spectroscopy and Radiative Transfer 110, 533 -- 572. Strobel, D., 1977. NH3 and PH3 photochemistry in the jovian at- mosphere. The Astrophysical Journal 214, L97 -- L99. Taylor, F., Atreya, S., Encrenaz, T., Hunten, D., Irwin, P., Owen, T., 2004. Jupiter: The Planet, Satellites and Magnetosphere. Cambridge University Press. chapter The Composition of the At- mosphere of Jupiter. pp. 59 -- 78. Wang, D., Gierasch, P.J., Lunine, J.I., Mousis, O., 2015. New insights on Jupiter's deep water abundance from disequilibrium species. Icarus 250, 154 -- 164. Wang, D., Lunine, J.I., Mousis, O., 2016. Modeling the disequilib- rium species for Jupiter and Saturn: Implications for Juno and Saturn entry probe. Icarus 276, 21 -- 38. Wenger, C., Champion, J., 1998. Spherical top data system (STDS) software for the simulation of spherical top spectra. Journal of Quantitative Spectroscopy and Radiative Transfer 59, 471 -- 480. Westphal, J., 1969. Observations of localised 5-micron radiation from Jupiter. The Astrophysical Journal 157, L63 -- L64. Wildt, R., 1932. Absorptionsspektren und atmospharen der gros; en planeten. Veroeffentlichungen der Universitaets-Sternwarte zu Goettingen 2, 171. Wong, M.H., Mahaffy, P.R., Atreya, S.K., Niemann, H.B., Owen, T.C., 2004. Updated Galileo probe mass spectrometer measure- ments of carbon, oxygen, nitrogen, and sulfur on Jupiter. Icarus 171, 153 -- 170. Zhang, X., West, R., Banfield, D., Yung, Y., 2013. Stratospheric aerosols on Jupiter from Cassini observations. Icarus 226, 159 -- 171. Encrenaz, T., de Graauw, T., Schaeidt, S., Lellouch, E., Feuchtgru- ber, H., Beintema, D., B´ezard, B., Drossart, P., Griffin, M., Heras, A., et al., 1996. First results of ISO-SWS observations of Jupiter. Astronomy and Astrophysics 315, L397 -- L400. Fegley, B., Lodders, K., 1994. Chemical models of the deep atmo- spheres of Jupiter and Saturn. Icarus 110, 117 -- 154. Ferris, J., Bossard, A., Khwaja, H., 1984. Mechanism of phos- phine photolysis. application to jovian atmospheric photochem- istry. Journal of the American Chemical Society 106, 318 -- 324. Fink, U., Larson, H.P., Treffers, R.R., 1978. Germane in the atmo- sphere of Jupiter. Icarus 34, 344 -- 354. Flasar, F.M., Gierasch, P.J., 1978. Turbulent convection within rapidly rotating superadiabatic fluids with horizontal tempera- ture gradients. Geophysical & Astrophysical Fluid Dynamics 10, 175 -- 212. Fletcher, L., Orton, G., Teanby, N., Irwin, P., 2009. Phosphine on Jupiter and Saturn from Cassini/CIRS. Icarus 202, 543 -- 564. Giles, R.S., Fletcher, L.N., Irwin, P.G., 2015. Cloud structure and composition of Jupiter's troposphere from 5-m Cassini VIMS spectroscopy. Icarus 257, 457 -- 470. Gillett, F., Low, F., Stein, W., 1969. The 2.8-14 micron spectrum of Jupiter. The Astrophysical Journal 157, 925. Grevesse, N., Asplund, M., Sauval, A., 2007. The solar chemical composition. Space Science Reviews 130, 105 -- 114. Guillemin, J.C., Lassalle, L., Janati, T., 1995. Germane photochem- istry. photolysis of gas mixtures of planetary interest. Planetary and Space Science 43, 75 -- 81. Irwin, P., Parrish, P., Fouchet, T., Calcutt, S., Taylor, F., Simon- Miller, A., Nixon, C., 2004. Retrievals of jovian tropospheric phos- phine from Cassini/CIRS. Icarus 172, 37 -- 49. Irwin, P., Teanby, N., de Kok, R., Fletcher, L., Howett, C., Tsang, C., Wilson, C., Calcutt, S., Nixon, C., Parrish, P., 2008. The NEMESIS planetary atmosphere radiative transfer and retrieval tool. Journal of Quantitative Spectroscopy and Radiative Transfer 109, 1136 -- 1150. Irwin, P., Weir, A., Smith, S., Taylor, F., Lambert, A., Calcutt, S., Cameron-Smith, P., Carlson, R., Baines, K., Orton, G., et al., 1998. Cloud structure and atmospheric composition of Jupiter retrieved from Galileo near-infrared mapping spectrometer real- time spectra. Journal of Geophysical Research: Planets (1991 -- 2012) 103, 23001 -- 23021. Irwin, P.G., 2003. Giant Planets of Our Solar System. Springer Praxis. chapter Mean and eddy circulation of the giant planet atmospheres. Jacquinet-Husson, N., Arie, E., Ballard, J., Barbe, A., Bjoraker, G., Bonnet, B., Brown, L., Camy-Peyret, C., Champion, J., Chedin, A., et al., 1999. The 1997 spectroscopic GEISA databank. Journal of Quantitative Spectroscopy and Radiative Transfer 62, 205 -- 254. Jacquinet-Husson, N., Crepeau, L., Armante, R., Boutammine, C., Ch´edin, A., Scott, N., Crevoisier, C., Capelle, V., Boone, C., Poulet-Crovisier, N., et al., 2011. The 2009 edition of the GEISA spectroscopic database. Journal of Quantitative Spectroscopy and Radiative Transfer 112, 2395 -- 2445. Kaufl, H.U., Ballester, P., Biereichel, P., Delabre, B., Donaldson, R., Dorn, R., Fedrigo, E., Finger, G., Fischer, G., Franza, F., et al., 2004. CRIRES: a high resolution infrared spectrograph for ESO's VLT, in: Proc. SPIE. Kunde, V., Hanel, R., Maguire, W., Gautier, D., Baluteau, J., Marten, A., Chedin, A., Husson, N., Scott, N., 1982. The tropo- spheric gas composition of Jupiter's North Equatorial Belt (NH3, PH3, CH3D, GeH4, H2) and the jovian D/H isotopic ratio. The Astrophysical Journal 263, 443 -- 467. Larson, H., Fink, U., Treffers, R., 1977. Phosphine in Jupiter's at- mosphere - the evidence from high-altitude observations at 5 mi- crometers. The Astrophysical Journal 211, 972 -- 979. Lellouch, E., B´ezard, B., Fouchet, T., Feuchtgruber, H., Encrenaz, T., de Graauw, T., 2001. The deuterium abundance in Jupiter and Saturn from ISO-SWS observations. Astronomy and Astrophysics 370, 610 -- 622. Lellouch, E., Drossart, P., Encrenaz, T., 1989. A new analysis of the jovian 5-m Voyager/IRIS spectra. Icarus 77, 457 -- 465. 20
1606.00023
1
1606
2016-05-31T20:05:34
HATS-25b through HATS-30b: A Half-dozen New Inflated Transiting Hot Jupiters from the HATSouth Survey
[ "astro-ph.EP" ]
We report six new inflated hot Jupiters (HATS-25b through HATS-30b) discovered using the HATSouth global network of automated telescopes. The planets orbit stars with $V$ magnitudes in the range $\sim 12-14$ and have masses in the largely populated $0.5M_J-0.7M_J$ region of parameter space but span a wide variety of radii, from $1.17R_J$ to $1.75 R_J$. HATS-25b, HATS-28b, HATS-29b and HATS-30b are typical inflated hot Jupiters ($R_p = 1.17-1.26R_J$) orbiting G-type stars in short period ($P=3.2-4.6$ days) orbits. However, HATS-26b ($R_p = 1.75R_J$, $P = 3.3024$ days) and HATS-27b ($R_p=1.50R_J$, $P=4.6370$ days) stand out as highly inflated planets orbiting slightly evolved F stars just after and in the turn-off points, respectively, which are among the least dense hot Jupiters, with densities of $0.153$ g cm$^{-3}$ and $0.180$ g cm$^{-3}$, respectively. All the presented exoplanets but HATS-27b are good targets for future atmospheric characterization studies, while HATS-27b is a prime target for Rossiter-McLaughlin monitoring in order to determine its spin-orbit alignment given the brightness ($V = 12.8$) and stellar rotational velocity ($v \sin i \approx 9.3$ km/s) of the host star. These discoveries significantly increase the number of inflated hot Jupiters known, contributing to our understanding of the mechanism(s) responsible for hot Jupiter inflation.
astro-ph.EP
astro-ph
Draft version 01 June 2, 2016 Preprint typeset using LATEX style emulateapj v. 2/16/10 6 1 0 2 y a M 1 3 . ] P E h p - o r t s a [ 1 v 3 2 0 0 0 . 6 0 6 1 : v i X r a HATS-25b THROUGH HATS-30b: A HALF-DOZEN NEW INFLATED TRANSITING HOT JUPITERS FROM THE HATSOUTH SURVEY† N. Espinoza1,2, D. Bayliss3, J. D. Hartman4, G. ´A. Bakos4,*,**, A. Jord´an1,2, G. Zhou5, L. Mancini6, R. Brahm1,2, S. Ciceri6, W. Bhatti4, Z. Csubry4, M. Rabus1,6, K. Penev4, J. Bento7, M. de Val-Borro4, T. Henning6, B. Schmidt7, V. Suc1, D. J. Wright8,9, C.G. Tinney8,9, T.G. Tan10, R. Noyes5 Draft version 01 June 2, 2016 ABSTRACT We report six new inflated hot Jupiters (HATS-25b through HATS-30b) discovered using the HAT- South global network of automated telescopes. The planets orbit stars with V magnitudes in the range ∼ 12 − 14 and have masses in the largely populated 0.5MJ − 0.7MJ region of parameter space but span a wide variety of radii, from 1.17RJ to 1.75RJ. HATS-25b, HATS-28b, HATS-29b and HATS-30b are typical inflated hot Jupiters (Rp = 1.17 − 1.26RJ) orbiting G-type stars in short period (P = 3.2 − 4.6 days) orbits. However, HATS-26b (Rp = 1.75RJ, P = 3.3024 days) and HATS-27b (Rp = 1.50RJ, P = 4.6370 days) stand out as highly inflated planets orbiting slightly evolved F stars just after and in the turn-off points, respectively, which are among the least dense hot Jupiters, with densities of 0.153 g cm−3and 0.180 g cm−3, respectively. All the presented exoplanets but HATS-27b are good targets for future atmospheric characterization studies, while HATS-27b is a prime target for Rossiter-McLaughlin monitoring in order to determine its spin-orbit alignment given the brightness (V = 12.8) and stellar rotational velocity (v sin i ≈ 9.3 km/s) of the host star. These discoveries significantly increase the number of inflated hot Jupiters known, contributing to our understanding of the mechanism(s) responsible for hot Jupiter inflation. Subject headings: planetary systems - stars: individual ( HATS-25, GSC 6716-01190, HATS-26, GSC 6614-01083 HATS-27, GSC 8245-02236 HATS-28, GSC 8382-00661 HATS- 29, GSC 8763-00475 HATS-30, GSC 8471-00231 ) techniques: spectroscopic, photometric 1. INTRODUCTION Since the observation of the transit of HD209458b, the first exoplanet to be observed to transit its host star by Charbonneau et al. (2000) and Henry et al. (2000), the field of transiting extrasolar planets has evolved tremen- 1 Instituto de Astrof´ısica, Facultad de F´ısica, Pontificia Uni- versidad Cat´olica de Chile, Av. Vicuna Mackenna 4860, 7820436 Macul, Santiago, Chile; [email protected]. 2 Millennium Institute of Astrophysics, Av. Vicuna Mackenna 4860, 782-0436 Macul, Santiago, Chile. 3 Observatoire Astronomique de l'Universit´e de Geneve, 51 ch. des Maillettes, 1290 Versoix, Switzerland. 4 Department of Astrophysical Sciences, Princeton University, NJ 08544, USA. 5 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA. 6 Max Planck Institute for Astronomy, Heidelberg, Germany. 7 Research School of Astronomy and Astrophysics, Australian National University, Canberra, ACT 2611, Australia. 8 Exoplanetary Science at UNSW, School of Physics, UNSW Australia, 20152, Australia. 9 Australian Centre for Astrobiology, UNSW Australia, 20152, Australia. 10 Perth Exoplanet Survey Telescope, Perth, Australia. * Alfred P. Sloan Research Fellow. ** Packard Fellow. † The HATSouth network is operated by a collaboration con- sisting of Princeton University (PU), the Max Planck Insti- tute fur Astronomie (MPIA), the Australian National Univer- sity (ANU), and the Pontificia Universidad Cat´olica de Chile (PUC). The station at Las Campanas Observatory (LCO) of the Carnegie Institute is operated by PU in conjunction with PUC, the station at the High Energy Spectroscopic Survey (H.E.S.S.) site is operated in conjunction with MPIA, and the station at Siding Spring Observatory (SSO) is operated jointly with ANU. Based in part on observations made with the MPG 2.2 m Tele- scope at the ESO Observatory in La Silla. dously. Transiting planets not only allow us to study the distribution of exoplanetary sizes, but in combina- tion with mass measurements allow us to unveil the wide range of densities for these distant worlds. This is crit- ical data that delivers physical characterisation of these systems. In addition, these systems allow the study of atmospheric properties (see, e.g., Crossfield 2015, and references therein) and the relationship between the or- bits of these systems and the spin of their host stars (Queloz et al. 2000; Ohta et al. 2005; Winn 2007). The so-called "hot Jupiters" (i.e. planets with masses and radii similar to Jupiter, but with periods P < 10 days) have been amongst the most studied exoplanets. Their observed sizes, orbits and compositions have pre- sented mutliple theoretical challenges. One of the most substantial challenges has been to explain the observed "inflated" nature of most of these systems (i.e. the fact that their radii are typically larger than what is expected from models of irradiated planets see, e.g., Baraffe et al. 2003; Fortney et al. 2007). This inflation suggests that additional processes must be at hand helping to avoid the gravitational contraction that self-gravitating bodies are subject to (see, e.g., Spiegel & Burrows 2013, for a comprehensive review of the subject). Another long-lasting puzzle is the exact way in which these exoplanets acquire such close-in orbits. Core- accretion theory predicts these planets would form from a solid ∼ 10M⊕ embryo that then accumulates large amounts of gas from the protoplanetary disk at sev- eral AU from the host star (Lissauer & Stevenson 2007). Once formed they migrate inwards, with the two main 2 Espinoza et al. mechanisms proposed as driving this migration being the planet's interaction with the protoplanetary disk (Goldreich & Tremaine 1980) and/or interaction of the planet with other planetary or stellar objects in the sys- tem (see, e.g., Rasio & Ford 1996; Wu & Lithwick 2011; Fabrycky & Tremaine 2007; Petrovich 2015). The characterisation to make these 2010; example, one popular model (Batygin & Stevenson transiting nature of systems allows observational powerful tests of a variety of models proposed for them. For explaining the inflated nature of hot Jupiters is Ohmic dissi- pation Perna et al. 2010; Batygin et al. 2011; Huang & Cumming 2012; Wu & Lithwick 2013). However, many of the physical parameters that underlie these models – such as wind speeds and planetary magnetic fields – are largely un- known and are only just beginning to be constrained via detailed photometric (see, e.g., Kataria et al. 2016, and references therein) and spectroscopic (Kislyakova et al. 2014; Louden & Wheatley 2015) characterization of transiting systems. Other models (e.g., increased opac- ities in the atmospheres of hot Jupiters Burrows et al. 2007), can be tested by detailed spectral characteri- zation of exoplanet atmospheres, which to date has mainly been provided through the technique of trans- mission spectroscopy. Interestingly, the composition of exoplanets inferred from studying their atmospheres is not only relevant for the problem of inflation or the study of atmospheric abundances in hot Jupiters (see, e.g., Sing et al. 2016), but can also constrain proposed migration mechanisms through the estimation of carbon-to-oxygen ratios (Madhusudhan et al. 2014; Benneke 2015). Detection of more of these characteriz- able systems is thus critical to build the large samples required to test physical models. In this work, we report the discovery of six new, well-characterized transiting hot Jupiters using the HATSouth global network of automated telescopes (Bakos et al. 2013), all of which are inflated and amenable for future atmospheric or Rossiter McLaugh- lin characterization: HATS-25b, HATS-26b, HATS-27b, HATS-28b, HATS-29b and HATS-30b. The structure of the paper is as follows. In Section 2 we summarize the detection of the photometric transit signal and the subsequent spectroscopic and photometric observations of each star to confirm and characterize the planets. In Section 3 we analyze the data to rule out false positive scenarios, and to determine the stellar and planetary pa- rameters. Our findings are discussed in Section 4. 2. OBSERVATIONS 2.1. Photometric detection In Table 1 we summarise the HATSouth discovery data of the six exoplanets presented in this work, all of which used data from the three HATSouth sites, namely, the site at Las Campanas Observatory in Chile (LCO, whose stations are designated HS-1 and HS-2), the site at of the HESS in Namibia (whose stations are designated HS-3 and HS-4) and the site at the Siding Spring Observa- tory (SSO, whose stations are designated HS-5 and HS- 6). The large number of observations for HATS-28 and HATS-29 are due to them being observed as part of the HATSouth "super-fields" program, where observations of the same field are taken with two telescopes from each HATSouth site. The large number of observations for HATS-30 are due to overlaps between its field and adja- cent HATSouth fields. The observations, reductions and analysis of the data were carried out as detailed in Bakos et al. (2013). In summary, the acquired images were obtained with a ca- dence of ≈ 300 s using a r SDSS filter on each of the sites. The images were then reduced and the result- ing lightcurves detrended using the methods described in Hartman et al. (2015). Finally, a Box Least Squares (BLS, Kov´acs et al. 2002) algorithm was ran on the lightcurves in order to search for periodic transit signa- tures. The discovery lightcurves of each of these stars, phased around the best-fit period of the transiting planet candidates, are depicted in Figure 1. In addition to these detections, we also searched for additional signals in the lightcurves in order to search for variability, activity and/or additional transit signals in the candidate systems. To this end, we ran BLS and Generalised Lomb Scargle (GLS, Zechmeister & Kurster 2009) algorithms on the residuals of each lightcurve, ex- ploring each of the significant peaks (which we defined as peaks with false alarm probabilities lower than 0.1%) in each of the periodograms by fitting boxes and sinu- soids, respectively, at those peaks and also inspecting visually the phased lightcurves. By analysing the peri- odograms along with the window functions, all the signif- icant peaks are near prominent sampling frequencies in the window function, or their harmonics, and are likely to be instrumental in origin. We thus conclude that all of the lightcurves do not show any additional signs of variability, activity and/or additional transit signals at least at the mmag level. 2.2. Spectroscopic Observations The spectroscopic observation of our planetary candi- dates is a two-step process. The first step is "reconnais- sance" spectroscopy, which consists of observations used both to rule out false positive scenarios produced by cer- tain configurations of stellar binaries that could mimic the detected transit features, and to estimate rough spec- tral parameters in order to estimate the physical and orbital parameters of the transiting planet candidates. The second step consists of spectroscopic observations that allow us to both confirm the planetary nature of the companion by radial velocity (RV) variations of the star due to the reflex motion produced by the planetary companion (which allows us to estimate its mass) and also to obtain precise stellar parameters from spectro- scopic observables in order to derive absolute parameters of the planetary companion. The spectroscopic observa- tions are summarized in Table 2, and are detailed below. 2.2.1. Reconnaissance spectroscopy The reconnaissance spectroscopy of our candidates was made using the Wide Field Spectrograph (WiFeS, Dopita et al. 2007), located on the ANU 2.3m telescope. Details of the observing strategy, reduction methods and the processing of the spectra for this instrument can be found in Bayliss et al. (2013). In summary, the observ- ing strategy usually consists in taking data with two res- olutions: R = λ/∆λ = 7000 (medium) and R = 3000 HATS-25b–HATS-30b 3 Summary of photometric observations Table 1 Instrument/Fielda Date(s) # Images Cadenceb (s) Filter Precisionc (mmag) HATS-25 HS-2.1/G568 HS-4.1/G568 HS-6.1/G568 LCOGT 1 m+CTIO/sinistro LCOGT 1 m+SSO/SBIG HATS-26 HS-2.3/G606 HS-4.3/G606 HS-6.3/G606 LCOGT 1 m+SAAO/SBIG LCOGT 1 m+SAAO/SBIG LCOGT 1 m+CTIO/sinistro LCOGT 1 m+CTIO/sinistro LCOGT 1 m+SSO/SBIG HATS-27 HS-2.1/G700 HS-4.1/G700 HS-6.1/G700 PEST 0.3 m LCOGT 1 m+SSO/SBIG HATS-28 HS-1.2/G747 HS-2.2/G747 HS-3.2/G747 HS-4.2/G747 HS-5.2/G747 HS-6.2/G747 LCOGT 1 m+CTIO/sinistro LCOGT 1 m+CTIO/sinistro HATS-29 HS-1.1/G747 HS-2.1/G747 HS-3.1/G747 HS-4.1/G747 HS-5.1/G747 HS-6.1/G747 LCOGT 1 m+CTIO/sinistro LCOGT 1 m+CTIO/sinistro HATS-30 HS-2.3/G754 HS-6.3/G754 HS-2.4/G754 HS-4.4/G754 HS-6.4/G754 HS-1.1/G755 HS-3.1/G755 HS-5.1/G755 LCOGT 1 m+SAAO/SBIG LCOGT 1 m+CTIO/sinistro 2011 Mar–2011 Aug 2011 Jul–2011 Aug 2011 May 2015 Feb 23 2015 Mar 16 2012 Feb–2012 Jun 2012 Feb–2012 Jun 2012 Feb–2012 Jun 2015 Mar 16 2015 Mar 26 2015 Apr 19 2015 May 21 2015 Jun 04 2011 Apr–2012 Jul 2011 Jul–2012 Jul 2011 May–2012 Jul 2015 Mar 12 2015 Apr 09 2013 Mar–2013 Oct 2013 Sep–2013 Oct 2013 Apr–2013 Nov 2013 Sep–2013 Nov 2013 Mar–2013 Nov 2013 Sep–2013 Nov 2015 Aug 31 2015 Sep 03 2013 Apr–2013 May 2013 Sep–2013 Oct 2013 Apr–2013 Nov 2013 Sep–2013 Nov 2013 Mar–2013 Nov 2013 Sep–2013 Nov 2015 Jun 01 2015 Jun 24 2012 Sep–2012 Dec 2012 Sep–2012 Dec 2012 Sep–2012 Dec 2012 Sep–2013 Jan 2012 Sep–2012 Dec 2011 Jul–2012 Oct 2011 Jul–2012 Oct 2011 Jul–2012 Oct 2014 Oct 19 2014 Oct 23 5055 841 131 70 104 3134 2761 1170 30 46 93 40 110 4603 3851 1512 141 282 4086 650 9051 1464 6018 1576 38 55 828 1331 9121 1505 6045 1544 90 36 3869 3000 3801 2820 2977 5180 4204 4904 50 56 290 301 289 226 196 291 300 299 199 137 166 165 73 292 301 300 132 75 287 287 297 297 297 290 223 223 289 287 297 297 297 290 166 162 282 285 282 292 285 291 287 296 196 226 r r r i i r r r i i i i i r r r RC i r r r r r r i i r r r r r r i i r r r r r r r r i i 6.9 7.8 9.3 1.1 2.3 7.0 7.1 6.8 1.8 2.0 1.0 1.7 2.9 6.3 7.5 7.1 4.1 2.3 12.8 11.5 12.1 12.5 10.7 11.4 1.4 1.4 7.2 7.5 6.1 8.2 6.4 7.2 1.2 1.0 6.1 6.2 6.0 6.6 5.7 9.2 7.4 6.5 1.2 1.0 a For HATSouth data we list the HATSouth unit, CCD and field name from which the observations are taken. HS-1 and -2 are located at Las Campanas Observatory in Chile, HS-3 and -4 are located at the H.E.S.S. site in Namibia, and HS-5 and -6 are located at Siding Spring Observatory in Australia. Each unit has 4 ccds. Each field corresponds to one of 838 fixed pointings used to cover the full 4π celestial sphere. All data from a given HATSouth field and CCD number are reduced together, while detrending through External Parameter Decorrelation (EPD) is done independently for each unique unit+CCD+field combination. b The median time between consecutive images rounded to the nearest second. Due to factors such as weather, the day–night cycle, guiding and focus corrections the cadence is only approximately uniform over short timescales. c The RMS of the residuals from the best-fit model. 4 Espinoza et al. g a m ∆ g a m ∆ -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 g a m ∆ g a m ∆ g a m ∆ g a m ∆ -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05 -0.04 HATS-25 HATS-26 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.06 -0.04 -0.02 0 0.02 0.04 0.06 Orbital phase HATS-27 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 g a m ∆ g a m ∆ g a m ∆ g a m ∆ -0.4 -0.2 0 0.2 0.4 Orbital phase -0.06 -0.04 -0.02 0 0.02 0.04 0.06 Orbital phase HATS-28 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.06 -0.1 -0.05 Orbital phase HATS-29 0 Orbital phase HATS-30 0.05 0.1 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 Orbital phase g a m ∆ g a m ∆ -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.06 -0.04 -0.02 0 0.02 0.04 0.06 Orbital phase Figure 1. Phase-folded unbinned HATSouth light curves for the six new transiting planet systems. In each case we show two panels. The top panel shows the full light curve, while the bottom panel shows the light curve zoomed-in on the transit. The solid lines show the model fits to the light curves. The dark filled circles in the bottom panels show the light curves binned in phase with a bin size of 0.002. HATS-25b–HATS-30b 5 (low). The former are used to search for RV variations at the ∼ 2 km/s level in order to rule out possible stel- lar companions, while the latter are used to estimate the spectroscopic parameters of the host stars. The results for each star were as follows: • HATS-25: four medium resolution spectra and one low resolution spectrum were obtained. From these, a temperature of 5830 ± 300 K, log(g) of 4.4 ± 0.3, metallicity of [Fe/H] = 0.0 ± 0.5 was de- rived, implying that the star was a G-type star. No RV variations at the ∼ 2 km/s level were found. • HATS-26: two medium resolution spectra and one low resolution spectrum were obtained. No RV variation at the ∼ 2 km/s level was found, and a temperature of 6333 ± 300 K, log(g) of 4.1 ± 0.3 and a metallicity of [Fe/H] = 0.0 ± 0.5 was derived, which pointed to an F-type star. • HATS-27: three medium resolution and one low resolution spectra were obtained. We found no variation at the ∼ 2 km/s level, and a tempera- ture of 6683 ± 300 K, log(g) of 4.5 ± 0.3 and a metallicity of [Fe/H] = 0.0 ± 0.5 was derived for this star, implying it was consistent with being an F-type star. • HATS-28: only one low resolution spectrum was obtained. With it, we derived a temperature of 5800 ± 300 K, log(g) of 4.5 ± 0.3 and a metallicity of [Fe/H] = 0.0 ± 0.5, hinting that this star was a G-type star. • HATS-29: four medium resolution spectra and one low resolution spectrum were obtained. No varia- tions at the ∼ 2 km/s level were found, and we derived a temperature of 5658 ± 300 K, log(g) of 4.5 ± 0.3 and a metallicity of [Fe/H] = 0.0 ± 0.5, for this star, finding it to be a G-type star. • HATS-30: three medium resolution spectra and one low resolution spectrum were obtained. No variations at the ∼ 2 km/s level in the RVs were found. A temperature of 6155 ± 300 K, log(g) of 4.6 ± 0.3 and a metallicity of [Fe/H] = 0.0 ± 0.5 was derived, which suggested the star was either a hot G-type or a cool F-type star. Given these results, our planet candidates were then promoted to our list requiring high-resolution spec- troscopy and high precision photometric follow-up ob- servations, which we now detail. 2.2.2. High-precision spectroscopy High-precision spectroscopy was obtained for our tar- gets with different instruments. Several R = 115000 spectra were taken with the High Accuracy Radial Ve- locity Planet Searcher (HARPS, Mayor et al. 2003) on the ESO 3.6m telescope at La Silla Observatory (LSO) between February 2015 and March 2016 in order to ob- tain high-precision RVs for HATS-25, HATS-26, HATS- 27 and HATS-29. Spectra with R = 48000 were also taken with the FEROS spectrograph (Kaufer & Pasquini 1998) mounted on the MPG 2.2m telescope at LSO be- tween July 2014 and July 2015 in order to both extract precise spectroscopic parameters of the host stars (see Section 3) and obtain precise RVs for all of our targets. In addition, R = 60000 spectra were also taken with the CORALIE (Queloz et al. 2001) spectrograph mounted on the 1.2m Euler telescope at LSO between June and November of 2014 for HATS-26, HATS-27, HATS-29 and HATS-30. The reduction of the CORALIE, FEROS and HARPS spectra followed the procedures described in Jord´an et al. (2014) for CORALIE, and adapted to FEROS and HARPS. Finally, eight R = 70000 spec- tra were obtained for HATS-29 on May 2015 to measure RVs, using the CYCLOPS2 fibre feed with the UCLES spectrograph on the 3.9m Anglo-Australian Telescope (AAT); the data was reduced following the methods de- tailed in Addison et al. (2013). The phased high-precision RV and bisector span (BS) measurements are shown for each system in Figure 2, while the data are listed in Table 8. It is important to note that the large observed scatter and errorbars on the RVs obtained from FEROS for HATS-27 are both due to the hot temperature of the star and due to contamination by scattered moonlight. Despite of this, it is evident that all the candidates show RV variations that are in phase with the photometric ephemeris. In addition, computed correlation coefficients between the RV and the BS mea- surements are all consistent with zero. 2.3. Photometric follow-up observations Photometric follow-up for the six systems was ob- tained in order to (1) rule out possible false positive sce- narios not identified in our reconnaissance spectroscopy (e.g., blended eclipsing binaries, hierarchical triples) that would leave signatures in the transit events (e.g., sig- nificantly different depths between different bands), (2) refine the ephemerides and (3) refine the derived tran- sit parameters obtained from the HATSouth discovery lightcurves. Our photometric follow-up observations are summarized in Table 1 and plotted in Figure 3. Photometry for these six systems was obtained mainly from 1m-class telescopes at different sites of the Las Cumbres Observatory Global Telescope (LCOGT) net- work (Brown et al. 2013), using the i filter (each of the sites used are indicated in Table 1). In particular, one partial transit and a full transit was observed for HATS- 25b on February 2015 and March 2015 respectively, three partial transits were observed for HATS-26b on April, May and June of 2015, one full transit was observed for HATS-27b on April 2015, two partial transits were ob- served for HATS-28b on August and September of 2015, one full transit and a partial transit was observed for HATS-29b on June 2015 and 2014, respectively, and two partial transits were observed for HATS-30b on Octo- ber 2014. In addition, one full transit of HATS-27b was observed using the 0.3m Perth Exoplanet Survey Tele- scope (PEST) on March of 2015. The instrument speci- fications, observing strategies and reduction of the data have been previously described in Bayliss et al. (2015) for the LCOGT data and in Zhou et al. (2014) for the PEST data. 2.4. Lucky imaging observations 6 Espinoza et al. Summary of spectroscopy observations Table 2 Instrument UT Date(s) # Spec. Res. ∆λ/λ/1000 S/N Rangea b γRV (km s−1) HATS-25 ANU 2.3 m/WiFeS ANU 2.3 m/WiFeS ESO 3.6 m/HARPS MPG 2.2 m/FEROS HATS-26 ANU 2.3 m/WiFeS ANU 2.3 m/WiFeS Euler 1.2 m/Coralie MPG 2.2 m/FEROS ESO 3.6 m/HARPS HATS-27 ANU 2.3 m/WiFeS ANU 2.3 m/WiFeS Euler 1.2 m/Coralie MPG 2.2 m/FEROS ESO 3.6 m/HARPS HATS-28 ANU 2.3 m/WiFeS MPG 2.2 m/FEROS HATS-29 ANU 2.3 m/WiFeS ANU 2.3 m/WiFeS ESO 3.6 m/HARPS AAT 3.9 m/CYCLOPS Euler 1.2 m/Coralie MPG 2.2 m/FEROS HATS-30 MPG 2.2 m/FEROS ANU 2.3 m/WiFeS ANU 2.3 m/WiFeS Euler 1.2 m/Coralie 2014 Jun–Aug 2014 Aug 5 2015 Feb–Apr 2015 Apr 9 2014 Jun 3–5 2014 Jun 4 2014 Jun 19–21 2015 Jan–Feb 2015 Feb 14–19 2014 Jun 2 2014 Jun 3–5 2014 Jun 20–21 2014 Jul–2015 Apr 2014 Aug–2016 Mar 2015 Jun 1 2015 Jun–Jul Apr 2014 Dec–2015 Mar 2015 Mar 2 2015 Apr 6–8 2015 May 6–9 2014 Jun 20–21 2015 Jun 13 2014 Oct–Dec 2014 Oct 4 2014 Oct 4–10 2014 Oct–Nov 4 1 8 1 2 1 2 8 4 1 3 3 15 11 1 18 4 1 3 9 4 3 7 1 3 6 7 3 115 48 7 3 60 48 115 3 7 60 48 115 3 48 7 3 115 70 60 48 48 3 7 60 26–152 88 11–23 64 95–107 121 17–19 56–74 19–23 50 4.6–12 21–22 18–92 4–25 38 17–52 3.1–31 45 12–23 16–30 16–19 48–50 60–96 233 87–118 22–30 30.0 · · · 31.663 31.649 -14.4 · · · -12.489 -12.516 -12.561 · · · -7.6 -3.521 -3.525 -3.582 · · · -8.651 -17.5 · · · -19.719 -19.722 -19.698 -19.670 -0.079 · · · 1.4 -0.112 RV Precisionc (m s−1) 4000 · · · 8.8 20 4000 · · · 5.2 21.0 21.3 · · · 4000 66 78 35 · · · 38 4000 · · · 18 40 11 20 8.3 · · · 4000 22 a S/N per resolution element near 5180 Afor all instruments but CYCLOPS, for which the S/N per resolution element near 5220 Ais presented. b For high-precision RV observations included in the orbit determination this is the zero-point RV from the best-fit orbit. For other instruments it is the mean value. We do not provide this quantity for the lower resolution WiFeS observations which were only used to measure stellar atmospheric parameters. c For high-precision RV observations included in the orbit determination this is the scatter in the RV residuals from the best-fit orbit (which may include astrophysical jitter), for other instruments this is either an estimate of the precision (not including jitter), or the measured standard deviation. We do not provide this quantity for low-resolution observations from the ANU 2.3 m/WiFeS. As part of a systematic program of obtaining high spatial resolution imaging for HATSouth candidates, "lucky" imaging observations were obtained for HATS- 26, HATS-27 and HATS-30 using the Astralux Sur cam- era (Hippler et al. 2009) mounted on the New Technol- ogy Telescope (NTT) at La Silla Observatory, in Chile on December 23 and 28, 2015. Both the HATS-26 and HATS-30 datasets, obtained on December 23, were obtained using the SDSS z′ fil- ter, while the HATS-27 dataset, obtained on December 28, was obtained using the SDSS i′ filter. A drizzle al- gorithm (Fruchter & Hook 2002) was used to combine the images, selecting the best of them from the set of ∼ 104 exposures taken for each target (104 images with an exposure time of 40 ms each for HATS-26, 2 × 104 im- ages with an exposure time of 15 ms each for HATS-27 and 2 × 104 images with an exposure time of 15 ms each for HATS-30). Figure 4 shows the resulting images for HATS-26 and HATS-30 and Figure 5 shows the resulting image for HATS-27, all of which are the combination of the best 10% of the images acquired for each target. The resulting images show an asymmetric extended profile for HATS-26 (a purely instrumental effect as confirmed by taking images of other targets on different nights), whereas the profile is fairly symmetric for HATS-27 and HATS-30 (we note that the latter shows an instrumental artefact close to (−2, −2) arcsecs from the target star). As can be seen from our images, no obvious companions were detected out to a 5′′ radius. In order to extract quantitative information from these images, we generated 5σ contrast curves for each of our targets, which required us to model the Point Spread Functions (PSFs). We decided to model the PSFs of our targets as a weighted sum of a Moffat profile (which models the central part of the PSF) and an asymmet- ric Gaussian (to model asymmetries in the PSF wings). The full width at half maximum (FWHM) of the full model was measured numerically at 100 different an- gles by finding the points at which the model has half of the peak flux, and the median of these measure- HATS-25 HATS-26 HATS-27 HATS-25b–HATS-30b 7 HARPS 100 ) 1 - s m ( V R 50 0 -50 -100 FEROS Coralie HARPS 150 100 ) 1 - s m ( V R 50 0 -50 -100 200 150 100 ) 1 - s m ( V R 50 0 -50 -100 -150 -200 300 250 200 150 100 50 0 -50 -100 -150 -200 500 400 300 200 100 0 -100 ) 1 - s m ( - C O ) 1 - s m ( S B FEROS Coralie HARPS 0.0 0.2 FEROS Coralie 150 100 ) 1 - s m ( V R 50 0 -50 -100 0.8 1.0 0.4 0.6 0.8 1.0 Phase with respect to Tc HATS-30 ) 1 - s m ( - C O -150 50 40 30 20 10 0 -10 -20 -30 -40 60 40 20 0 -20 -40 -60 -80 -100 ) 1 - s m ( S B 0.8 1.0 0.4 0.6 Phase with respect to Tc 0.0 0.2 0.4 0.6 0.8 1.0 Phase with respect to Tc ) 1 - s m ( - C O 30 20 10 0 -10 -20 -30 -40 100 80 60 40 20 0 -20 -40 -60 -80 -100 ) 1 - s m ( S B 200 150 100 ) 1 - s m ( V R 50 0 -50 -100 -150 -200 150 100 50 0 -50 -100 -150 150 100 50 0 -50 -100 -150 -200 ) 1 - s m ( - C O ) 1 - s m ( S B 0.0 0.2 FEROS 0.0 0.2 ) 1 - s m ( - C O -150 50 40 30 20 10 0 -10 -20 -30 -40 -50 200 ) 1 - s m ( S B 150 100 50 0 -50 0.0 0.2 0.8 1.0 0.4 0.6 Phase with respect to Tc 0.4 0.6 Phase with respect to Tc HATS-28 HATS-29 150 100 Coralie HARPS CYCLOPS ) 1 - s m ( V R 50 0 -50 -100 -150 100 80 60 40 20 0 -20 -40 -60 -80 -100 150 100 50 0 -50 -100 -150 ) 1 - s m ( - C O ) 1 - s m ( S B 0.0 0.2 0.8 1.0 0.4 0.6 Phase with respect to Tc Figure 2. Phased high-precision RV measurements for the six new transiting planet systems. The instruments used are labelled in the plots. In each case we show three panels. The top panel shows the phased measurements together with our best-fit circular-orbit model (see Table 6) for each system. Zero-phase corresponds to the time of mid-transit. The center-of-mass velocity has been subtracted. The second panel shows the velocity O−C residuals from the best fit. The error bars include the jitter terms listed in Tables 6 and 7 added in quadrature to the formal errors for each instrument. The third panel shows the bisector spans (BS). Note the different vertical scales of the panels. ments (the "effective" FWHM, FWHMeff ) is taken as the resolution limit of our observations. For HATS-26, we found FWHMeff = 3.27 ± 0.35 pixels, which given the pixel scale of 23 milli-arcseconds (mas) per pixel, gives a resolution limit of 75 ± 8 mas. For HATS-27, we found FWHMeff = 3.17 ± 0.28 pixels, which implies a resolution limit of 72 ± 6 mas. Finally, for HATS-30, FWHMeff = 3.55 ±0.29 pixels, which implies a resolution limit of 81 ± 7 mas. All the effective FWHMs are close to the diffraction limit of the instrument, which is ∼ 50 mas (Hippler et al. 2009). Once modelled, we subtracted the PSF of the target stars from the images and generated the contrast curves by an "injection and recovery" approach, in which we injected signals with the same fitted PSF parameters at different positions (r,θ) in the image, where r is the distance from the target star and θ is the azimuthal angle around it. We sampled r in steps of FWHMeff , while the angles are sampled at each radius covering 2π radians with independent regions of arc-length equal to FWHMeff. The injected sources were scaled in order to simulate a wide range of contrasts, exploring from ∆z′ = 0 to ∆z′ = 10 in 0.01 steps, where ∆z′ is the magnitude contrast with respect to the target star. We considered an injected source to be detectable if five or more pixels were 5σ above the noise level, which was estimated as the standard deviation in a box of size FWHMeff × FWHMeff at each position in the residual image at which the signals were injected. Finally, the contrast at each radius was obtained by averaging the azimuthal contrasts and the standard deviation of these azimuthal contrasts was taken as the error on the contrast at each radius. The resulting contrast curves for HATS-26 (blue) and HATS-30 (orange) are shown on Figure 6, where the grey bands show the uncertainty of the contrast at each radius. The corresponding contrast curve for HATS-27 is shown in Figure 7. Code to model the PSFs of images as explained here and to generate these contrast curves can be found at https://github.com/nespinoza/luckyimg-reduction. 8 Espinoza et al. HATS-25 HATS-26 HATS-27 i-band 2015 Feb 23 (LCOGT 1m) i-band 2015 Apr 19 (LCOGT 1m) RC-band 2015 Mar 12 (PEST 0.3m) i 2015 Mar 16 (LCOGT 1m) s t e s f f o y r a r t i b r A - ) g a m ( ∆ i i 2015 May 21 (LCOGT 1m) 2015 Jun 04 (LCOGT 1m) 0 0.02 0.04 0.06 0.08 0.1 0.12 s t e s f f o y r a r t i b r A - ) g a m ( ∆ i 2015 Apr 9 (LCOGT 1m) 0 0.02 0.04 0.06 0.08 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 Time from transit center (days) HATS-28 Time from transit center (days) HATS-29 Time from transit center (days) HATS-30 s t e s f f o y r a r t i b r A - ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0 0.02 0.04 s t e s f f o y r a r t i b r A - ) g a m ( ∆ 0.06 0.08 i-band 2015 Aug 31 (LCOGT 1m) i 2015 Sep 03 (LCOGT 1m) s t e s f f o y r a r t i b r A - ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 i-band 2015 Jun 1 (LCOGT 1m) i-band 2014 Oct 19 (LCOGT 1m) -0.01 0 0.01 i 2014 Jun 24 (LCOGT 1m) 0.02 i 2014 Oct 23 (LCOGT 1m) 0.03 s t e s f f o y r a r t i b r A - ) g a m ( ∆ 0.04 0.05 0.06 -0.1 -0.05 0 0.05 0.1 Time from transit center (days) 0.07 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 Time from transit center (days) -0.15 -0.1 -0.05 0 0.05 0.1 0.15 Time from transit center (days) Figure 3. Unbinned transit light curves for the six new transiting planet systems. The light curves have been corrected for quadratic trends in time fitted simultaneously with the transit model, and for correlations with up to three parameters describing the shape of the PSF. The dates of the events, filters and instruments used are indicated. Light curves following the first are displaced vertically for clarity. Our best fit from the global modeling described in Section 3.3 is shown by the solid lines. The residuals from the best-fit model are shown below in the same order as the original light curves. The error bars represent the photon and background shot noise, plus the readout noise. Note the differing vertical and horizontal scales used for each system. For HATS-25 we do not show the LCOGT 1 m light curves from UT 2015 Mar 16 and 26 which were taken entirely out of transit. HATS-25b–HATS-30b 9 ) c e s c r a ( C E D ∆ 4 2 0 −2 −4 HATS-26 2015-12-23 (AstraLux Sur+z') −4 −2 0 2 4 ∆ R.A. (arcsec) 0 −1 −2 −3 −4 −5 R e l a t i v e fl u x ( l o g - s c a l e ) ) c e s c r a ( C E D ∆ 4 2 0 −2 −4 HATS-30 2015-12-23 (AstraLux Sur+z') −4 −2 0 2 4 ∆ R.A. (arcsec) 0 −1 −2 −3 −4 −5 R e l a t i v e fl u x ( l o g - s c a l e ) Figure 4. (Left) AstraLux Sur z′-band observations of HATS-26. Circles of 1′′ radius (approximately the mean FWHM measured for the image) and 5′′ radius are shown for reference on the images. The central lines indicate the fitted center of the star with our PSF modelling (see text). (Right) Same image but for HATS-30. Note the difference in the shape of the PSF, which is a purely instrumental effect. 10 Espinoza et al. HATS-27 2015-12-28 (AstraLux Sur+i') ) c e s c r a ( C E D ∆ 4 2 0 −2 −4 0 −1 −2 −3 −4 −5 R e l a t i v e fl u x ( l o g - s c a l e ) 0 1 2 3 4 5 6 7 ′ i ∆ HATS-27 −4 −2 0 2 4 ∆ R.A. (arcsec) Figure 5. AstraLux Sur i′-band observations of HATS-27. The circles and lines indicate the same distances and positions as the ones described in Figure 4. 0 1 2 3 4 5 Radial distance (arcsec) Figure 7. Contrast curve generated for HATS-27 using our As- traLux Sur i′-band observations. Gray bands show the uncertainty given by the scatter in the contrast in the azimuthal direction at a given radius (see text for details). 0 1 2 3 4 5 6 ′ z ∆ HATS-26 HATS-30 0 1 2 3 4 5 Radial distance (arcsec) Figure 6. Contrast curves generated for HATS-26 (blue) and HATS-30 (orange) using our AstraLux Sur z′-band observations. Gray bands show the uncertainty given by the scatter in the con- trast in the azimuthal direction at a given radius (see text for details). HATS-25b–HATS-30b 11 3. ANALYSIS 3.1. Properties of the parent stars We determine the properties of the host stars us- ing the Zonal Atmospherical Stellar Parameter Estima- tor (ZASPE, Brahm et al., in preparation) on median combined FEROS spectra for all our systems except for HATS-25, where only one FEROS spectrum was ), log-gravity used. With the effective temperature (Teff∗ (log g⋆) metallicity ([Fe/H]) and the projected stellar ro- tational velocity of the star (v sin i) calculated for each of our systems, the Yonsei-Yale (Y2, Yi et al. 2001) isochrones were used to obtain the physical parameters of the host stars. However, instead of using log g⋆ to search for the best-fit isochrone, we follow Sozzetti et al. (2007) in using the stellar density (ρ∗), which is well constrained parameter by our transit fits. Once this was done and physical parameters were found, a second ZASPE iter- ation was done for all systems except for HATS-27, for which a second iteration did not improve the results. In this second iteration, the revised value of log g⋆ was used as input in order to derive the final properties of the stars. In order to calculate the distances to these stars, we compared their measured broad-band photometry to the predicted magnitudes in each filter from the isochrones, assuming an extinction law from Cardelli et al. (1989) with RV = 3.1. The resulting parameters for HATS- 25, HATS-26 and HATS-27 are given in Table 4, and for HATS-28, HATS-29 and HATS-30 in Table 5. The loca- tions of each star on an Teff⋆–ρ⋆ diagram (similar to a Hertzsprung-Russell diagram) are shown in Figure 8. It is interesting to note that while HATS-25, HATS- 28, HATS-29 and HATS-30 are typical G dwarfs, HATS- 26 and HATS-27 stand out as slightly evolved F stars which are just after and in the turn-off points, respec- tively. Consequently, they have radii of 2.04+0.15 −0.11R⊙ and 1.74+0.17 −0.10R⊙ which (combined with their effective tem- peratures of 6071 ± 81 K and 6438 ± 64 K, respectively) implies relatively large luminosities of 5.06+0.90 −0.64L⊙ and 4.67+0.92 −0.58L⊙. Because of this, their planets receive larger insolation levels than typical hot Jupiters with the same periods. 3.2. Excluding blend scenarios In order to exclude blend scenarios we carried out an analysis following Hartman et al. (2012). We attempt to model the available photometric data (including light curves and catalog broad-band photometric measure- ments) for each object as a blend between an eclipsing binary star system and a third star along the line of sight. The physical properties of the stars are constrained us- ing the Padova isochrones (Girardi et al. 2000), while we also require that the brightest of the three stars in the blend have atmospheric parameters consistent with those measured with ZASPE. We also simulate compos- ite cross-correlation functions (CCFs) and use them to predict RVs and BSs for each blend scenario considered. Based on this analysis we rule out blended stellar eclipsing binary scenarios for all six systems. However, in general we cannot rule out the possibility that one or more of these objects may be an unresolved binary star system with one component hosting a transiting planet, although limits can be placed on those scenar- ios for HATS-26, HATS-27 and HATS-30 based on our lucky imaging observations shown on Section 2.4. The results for each object are as follows: • HATS-25: All blend models tested give higher χ2 than a model of single star with a planet. Those blend models which cannot be rejected with greater than 5σ confidence predict either RV or BS varia- tions greater than 1 km s−1, which are excluded by the observations. • HATS-26: All blend models tested can be rejected with greater than 5σ confidence based on the pho- tometry alone. In particular, the blend models pre- dict a large out-of-transit variation due to the tidal distortion of the binary star components. Such a variation is ruled out by the HATSouth photome- try. • HATS-27: Same conclusion as for HATS-25. • HATS-28: All blend models tested can be rejected with greater than 4σ confidence based on the pho- tometry alone. • HATS-29: Blend models which cannot be rejected with greater than 5σ confidence based on the pho- tometry alone generally predict large RV and BS variations exceeding 1 km s−1. There is a narrow region of parameter space where the blend models are rejected at 4σ confidence based on the photom- etry, and the simulated RVs and BSs have scatters of a few 100 m s−1, which is not much greater than the measured values. However, the simulated RVs do not phase with the photometric ephemeris. • HATS-30: All blend models tested can be rejected with greater than 4σ confidence based on the pho- tometry alone. 3.3. Global modeling of the data We modeled the HATSouth photometry, the follow- up photometry, and the high-precision RV measure- ments following P´al et al. (2008); Bakos et al. (2010); Hartman et al. (2012). We fit Mandel & Agol (2002) transit models to the light curves, allowing for a dilu- tion of the HATSouth transit depth as a result of blend- ing from neighboring stars and over-correction by the trend-filtering method. For the follow-up light curves we include a quadratic trend in time, and linear trends with up to three parameters describing the shape of the PSF, in our model for each event to correct for systematic er- rors in the photometry. We fit Keplerian orbits to the RV curves allowing the zero-point for each instrument to vary independently in the fit, and allowing for RV jitter which we we also vary as a free parameter for each instru- ment. We used a Differential Evolution Markov Chain Monte Carlo procedure to explore the fitness landscape and to determine the posterior distributions of the pa- rameters. Note that we tried fitting both fixed-circular- orbits and free-eccentricity models to the data, and for all six systems find that the data are consistent with a cir- cular orbit. We estimate the Bayesian evidence for the fixed-circular and free-eccentricity models for each sys- tem, and find that in all six cases the fixed-circular model 12 Objecta HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 Espinoza et al. Light curve data for HATS-25–HATS-30. Table 3 BJDb (2,400,000+) 56076.42690 56090.33807 55955.86458 56113.52396 56016.14672 56062.51729 56020.78399 56006.87363 56030.05986 56076.43037 Magc 0.00302 0.00531 −0.00442 0.00956 0.00478 −0.00009 0.00065 −0.01152 −0.00319 0.00830 σMag 0.00469 0.00449 0.00406 0.00462 0.00406 0.00864 0.00445 0.00436 0.00425 0.00472 Mag(orig)d Filter Instrument · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · r r r r r r r r r r HS HS HS HS HS HS HS HS HS HS Note. - This table is available in a machine-readable form in the online journal. A portion is shown here for guidance regarding its form and content. a Either HATS-25, HATS-26, HATS-27, HATS-28, HATS-29 or HATS-30. b Barycentric Julian Date is computed directly from the UTC time without correction for leap seconds. c The out-of-transit level has been subtracted. For observations made with the HATSouth instruments (identified by "HS" in the "Instrument" column) these magnitudes have been corrected for trends using the EPD and TFA procedures applied prior to fitting the transit model. This procedure may lead to an artificial dilution in the transit depths. The blend factors for the HATSouth light curves are listed in Tables 6 and 7. For observations made with follow-up instruments (anything other than "HS" in the "Instrument" column), the magnitudes have been corrected for a quadratic trend in time, and for variations correlated with three PSF shape parameters, fit simultaneously with the transit. d Raw magnitude values without correction for the quadratic trend in time, or for trends correlated with the shape of the PSF. These are only reported for the follow-up observations. HATS-25 HATS-26 HATS-27 0.0 0.5 1.0 ] 3 m c / g [ * ρ 1.5 6200 6000 5800 5600 Effective temperature [K] 5400 HATS-28 2.0 2.5 0.0 0.5 1.0 ] 3 m c / g [ * ρ 1.5 2.0 2.5 6200 6000 5800 5600 Effective temperature [K] 5400 0.0 0.5 1.0 ] 3 m c / g [ * ρ 1.5 2.0 2.5 6200 6000 5800 5600 Effective temperature [K] 5400 0.0 0.5 1.0 ] 3 m c / g [ * ρ 1.5 2.0 2.5 6200 6000 5800 5600 Effective temperature [K] 5400 0.0 0.5 1.0 1.5 2.0 7000 6800 6600 6400 Effective temperature [K] HATS-30 6200 6000 ] 3 m c / g [ * ρ 0.0 0.5 1.0 ] 3 m c / g [ * ρ 1.5 2.0 2.5 6200 6000 5800 5600 Effective temperature [K] 5400 Figure 8. Model isochrones from Yi et al. (2001) for the measured metallicities of each of the six new transiting planet host stars. We show models for ages of 0.2 Gyr and 1.0 to 14.0 Gyr in 1.0 Gyr increments (ages increasing from left to right). The adopted values of Teff⋆ and ρ⋆ are shown together with their 1σ and 2σ confidence ellipsoids. The initial values of Teff⋆ and ρ⋆ from the first ZASPE and light curve analyses are represented with a triangle. has greater evidence. In particular, for the HATS-25, HATS-26, HATS-28, HATS-29 and HATS-30 systems, the Bayesian evidence for the fixed-circular-orbit model is 10, 8, 5, 660 and 3 times greater, respectively, than the eccentric-orbit model, favouring the former in these cases. For HATS-27, both models are indistinguishable, but the eccentricity is poorly constrained by the data at hand, giving implausibly high values for it. We therefore adopt the parameters that come from the fixed-circular- orbit models for all of the systems. The resulting pa- rameters for HATS-25b, HATS-26b and HATS-27b are listed in Table 6, while for HATS-28b, HATS-29b and HATS-30b they are listed in Table 7. As can be observed from the tables, all the presented planets can be classified as typical hot Jupiters, with short-periods, similar masses of ∼ 0.6MJ and larger- than-Jupiter radii. 4. DISCUSSION In this paper we present six new transiting planets discovered by the HAT-South survey. Figure 9 puts the discovered exoplanets in the context of all known transiting hot Jupiters (here defined as planets with 0.1MJ < M < 5MJ and periods P < 10d) discov- ered to date14 with secure masses and radii (i.e., masses 14 Data taken from exoplanets.eu on 2016/02/01. HATS-25b–HATS-30b 13 Stellar parameters for HATS-25, HATS-26 and HATS-27 Table 4 Parameter HATS-25 Value HATS-26 Value HATS-27 Value Source Astrometric properties and cross-identifications 2MASS 13513786-2346522 2MASS 09394244-2835081 2MASS 12541261-4635157 2MASS-ID . . . . . . . . . GSC-ID . . . . . . . . . . . R.A. (J2000) . . . . . . . Dec. (J2000) . . . . . . . µR.A. (mas yr−1) µDec. (mas yr−1) Spectroscopic properties Teff⋆ (K). . . . . . . . . . . [Fe/H] . . . . . . . . . . . . . v sin i (km s−1) . . . . vmac (km s−1) . . . . . vmic (km s−1) . . . . . . γRV (m s−1) . . . . . . . Photometric properties B (mag) . . . . . . . . . . . V (mag) . . . . . . . . . . . g (mag) . . . . . . . . . . . . r (mag) . . . . . . . . . . . . i (mag) . . . . . . . . . . . . J (mag) . . . . . . . . . . . H (mag) . . . . . . . . . . . Ks (mag) . . . . . . . . . . Derived properties M⋆ (M⊙) . . . . . . . . . . R⋆ (R⊙) . . . . . . . . . . . log g⋆ (cgs) . . . . . . . . ρ⋆ (g cm−3). . . . . . . . ρ⋆ (g cm−3) e . . . . . . L⋆ (L⊙) . . . . . . . . . . . MV (mag) . . . . . . . . . MK (mag,ESO). . . . Age (Gyr) . . . . . . . . . AV (mag) . . . . . . . . . Distance (pc) . . . . . . GSC 6716-01190 13h51m37.80s −23◦46′52.2′′ −20.5 ± 1.0 −11.9 ± 1.1 5715 ± 73 0.020 ± 0.050 3.88 ± 0.50 3.90 1.04 GSC 6614-01083 09h39m42.44s −28◦35′08.1′′ −1.3 ± 1.4 −6.1 ± 1.3 6071 ± 81 −0.020 ± 0.050 7.48 ± 0.50 4.44 1.29 31663.2 ± 3.6 −12515.9 ± 6.7 −3582 ± 12 13.812 ± 0.030 13.097 ± 0.030 13.380 ± 0.020 12.909 ± 0.040 12.687 ± 0.050 11.788 ± 0.022 11.487 ± 0.024 11.416 ± 0.021 0.994 ± 0.035 1.107 ± 0.069 4.347 ± 0.053 1.03 ± 0.20 1.03 ± 0.20 1.17 ± 0.17 4.67 ± 0.16 3.10 ± 0.14 7.5 ± 1.9 0.083 ± 0.061 466 ± 30 13.553 ± 0.030 12.955 ± 0.030 13.229 ± 0.010 12.822 ± 0.010 12.695 ± 0.030 11.839 ± 0.024 11.510 ± 0.024 11.435 ± 0.021 1.299+0.113 −0.056 2.04+0.15 −0.11 3.936 ± 0.046 0.219 ± 0.033 0.218 ± 0.034 5.06+0.90 −0.64 3.03 ± 0.17 1.68 ± 0.15 4.04+0.62 −0.94 0.140 ± 0.070 907+69 −49 13.239 ± 0.050 12.766 ± 0.040 12.927 ± 0.040 12.665 ± 0.040 12.515 ± 0.080 11.831 ± 0.022 11.651 ± 0.023 11.550 ± 0.023 1.415 ± 0.048 1.74+0.17 −0.10 4.107 ± 0.049 0.380 ± 0.063 0.379 ± 0.064 4.67+0.92 −0.58 3.06 ± 0.17 1.98 ± 0.15 2.30 ± 0.22 0.084 ± 0.066 840+80 −51 GSC 8245-02236 12h54m12.60s −46◦35′15.8′′ −10.2 ± 1.1 4.7 ± 1.1 2MASS 2MASS UCAC4 UCAC4 6438 ± 64 0.090 ± 0.040 9.32 ± 0.50 5.01 1.67 ZASPEa ZASPE ZASPE Assumed Assumed FEROS or HARPSb APASSc APASSc APASSc APASSc APASSc 2MASS 2MASS 2MASS YY+ρ⋆+ZASPE d YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE Light curves YY+Light curves+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE Note. - For HATS-25 and HATS-26 the fixed-circular-orbit model has a higher Bayesian evidence than the eccentric-orbit model (it is 10 and 8 times greater for these two systems respectively). We therefore assume a fixed circular orbit in generating the parameters listed for both of these systems. For HATS-27 the free-eccentricity model has an indistinguishable Bayesian evidence from the fixed-circular model, but in this case the eccentricity is poorly constrained with implausibly high values permitted by the low S/N RV measurements. For this system we also adopt the fixed-circular model parameters. a ZASPE = Zonal Atmospherical Stellar Parameter Estimator routine for the analysis of high-resolution spectra (Brahm et al. 2016, in preparation), applied to the FEROS spectra of HATS-25 and HATS-26. These parameters rely primarily on ZASPE, but have a small dependence also on the iterative analysis incorporating the isochrone search and global modeling of the data. b From FEROS for HATS-26 and from HARPS for HATS-25 and HATS-27. The error on γRV is determined from the orbital fit to the RV measurements, and does not include the systematic uncertainty in transforming the velocities to the IAU standard system. The velocities have not been corrected for gravitational redshifts. c From APASS DR6 for as listed in the UCAC 4 catalog (Zacharias et al. 2012). d YY+ρ⋆+ZASPE = Based on the YY isochrones (Yi et al. 2001), ρ⋆ as a luminosity indicator, and the ZASPE results. e In the case of ρ⋆ we list two values. The first value is determined from the global fit to the light curves and RV data, without imposing a constraint that the parameters match the stellar evolution models. The second value results from restricting the posterior distribution to combinations of ρ⋆+Teff⋆+[Fe/H] that match to a YY stellar model. and radii inconsistent with zero at 3 − σ). We can see that the discovered exoplanets all fall in a heavily pop- ulated region of the mass distribution of hot Jupiters near ∼ 0.6MJ. However, although HATS-30b, HATS- 29b, HATS-28b and HATS-25b all fall in the peak of the radius distribution, with radii of ∼ 1.2RJ , making them all moderately inflated planets, HATS-26b (1.75RJ ) and HATS-27b (1.50RJ ) fall on the high-end part of it, mak- ing them highly inflated planets. These two hot Jupiters also have the lowest densities of the group: HATS-26b has a density of only 0.153 ± 0.042 g cm−3, while HATS- 27b has a density of 0.180+0.083 −0.057 g cm−3. These densities are quite unusual not only in this group of planets, but also among the population of hot Jupiters in general: of the known systems, only ∼ 10 have densities lower than 0.2 g cm−3. The empirical relations in equation (9) of Enoch et al. (2012) predict the radii of these six new exoplanets to within the uncertainties. Therefore, these exoplanets ap- pear to follow the trends followed by other close-in ex- 14 Espinoza et al. Stellar parameters for HATS-28, HATS-29 and HATS-30 Table 5 Parameter HATS-28 Value HATS-29 Value HATS-30 Value Source Astrometric properties and cross-identifications 2MASS-ID . . . . . . . . . GSC-ID . . . . . . . . . . . R.A. (J2000) . . . . . . . Dec. (J2000) . . . . . . . µR.A. (mas yr−1) µDec. (mas yr−1) Spectroscopic properties Teff⋆ (K). . . . . . . . . . . [Fe/H] . . . . . . . . . . . . . v sin i (km s−1) . . . . vmac (km s−1) . . . . . vmic (km s−1) . . . . . . γRV (m s−1) . . . . . . . Photometric properties B (mag) . . . . . . . . . . . V (mag) . . . . . . . . . . . g (mag) . . . . . . . . . . . . r (mag) . . . . . . . . . . . . i (mag) . . . . . . . . . . . . J (mag) . . . . . . . . . . . H (mag) . . . . . . . . . . . Ks (mag) . . . . . . . . . . Derived properties M⋆ (M⊙) . . . . . . . . . . R⋆ (R⊙) . . . . . . . . . . . log g⋆ (cgs) . . . . . . . . ρ⋆ (g cm−3). . . . . . . . ρ⋆ (g cm−3) e . . . . . . L⋆ (L⊙) . . . . . . . . . . . MV (mag) . . . . . . . . . MK (mag,ESO). . . . Age (Gyr) . . . . . . . . . AV (mag) . . . . . . . . . Distance (pc) . . . . . . 2MASS 18573592-4908184 2MASS 19002314-5453354 2MASS 00222848-5956331 GSC 8382-00661 18h57m36.00s −49◦08′18.5′′ 10.3 ± 1.6 −2.4 ± 1.4 5498 ± 84 0.010 ± 0.060 2.6 ± 1.0 3.56 0.93 GSC 8763-00475 GSC 8471-00231 19h00m23.04s −54◦53′35.5′′ 2.8 ± 1.3 −37.1 ± 3.7 5670 ± 110 0.160 ± 0.080 2.35 ± 0.80 3.83 1.02 00h22m28.49s −59◦56′33.2′′ −25.3 ± 1.0 −8.2 ± 1.0 5943 ± 70 0.060 ± 0.050 4.11 ± 0.50 4.25 1.19 −8650.5 ± 9.1 −19719.3 ± 6.9 −78.6 ± 4.2 14.697 ± 0.020 13.934 ± 0.080 14.274 ± 0.030 13.717 ± 0.010 13.615 ± 0.010 12.522 ± 0.026 12.188 ± 0.025 12.086 ± 0.029 0.929 ± 0.036 0.922 ± 0.040 4.476 ± 0.039 1.68 ± 0.27 1.67 ± 0.22 0.696 ± 0.084 5.28 ± 0.14 3.53 ± 0.10 6.2 ± 2.8 0.055+0.124 −0.055 521 ± 25 13.361 ± 0.010 12.612 ± 0.010 12.950 ± 0.010 12.430 ± 0.010 12.154 ± 0.010 11.286 ± 0.026 10.933 ± 0.021 10.877 ± 0.019 1.032 ± 0.049 1.073 ± 0.038 4.389 ± 0.027 1.17 ± 0.11 1.17 ± 0.11 1.07 ± 0.13 4.77 ± 0.15 3.166 ± 0.088 5.5+2.6 −1.7 0.111 ± 0.082 351 ± 15 12.790 ± 0.010 12.192 ± 0.010 12.439 ± 0.010 12.046 ± 0.010 11.935 ± 0.010 11.129 ± 0.024 10.826 ± 0.024 10.793 ± 0.019 1.093 ± 0.031 1.061 ± 0.039 4.425 ± 0.030 1.34 ± 0.19 1.29 ± 0.14 1.25 ± 0.12 4.57 ± 0.12 3.123 ± 0.088 2.3 ± 1.2 0.0000 ± 0.0066 339 ± 16 2MASS 2MASS UCAC4 UCAC4 ZASPEa ZASPE ZASPE Assumed Assumed FEROS or HARPSb APASSc APASSc APASSc APASSc APASSc 2MASS 2MASS 2MASS YY+ρ⋆+ZASPE d YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE Light curves YY+Light Curves+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE YY+ρ⋆+ZASPE Note. - For all three systems the fixed-circular-orbit model has a higher Bayesian evidence than the eccentric-orbit model (it is 5, 660, and 3 times greater for HATS-28, HATS-29 and HATS-30, respectively). We therefore assume a fixed circular orbit in generating the parameters listed for these systems. a ZASPE = Zonal Atmospherical Stellar Parameter Estimator routine for the analysis of high-resolution spectra (Brahm et al. 2016, in preparation), applied to the FEROS spectra of HATS-28 and HATS-26. These parameters rely primarily on ZASPE, but have a small dependence also on the iterative analysis incorporating the isochrone search and global modeling of the data. b From FEROS for HATS-28 and HATS-30, and from HARPS for HATS-29. The error on γRV is determined from the orbital fit to the RV measurements, and does not include the systematic uncertainty in transforming the velocities to the IAU standard system. The velocities have not been corrected for gravitational redshifts. c From APASS DR6 for as listed in the UCAC 4 catalog (Zacharias et al. 2012). d YY+ρ⋆+ZASPE = Based on the YY isochrones (Yi et al. 2001), ρ⋆ as a luminosity indicator, and the ZASPE results. e In the case of ρ⋆ we list two values. The first value is determined from the global fit to the light curves and RV data, without imposing a constraint that the parameters match the stellar evolution models. The second value results from restricting the posterior distribution to combinations of ρ⋆+Teff⋆+[Fe/H] that match to a YY stellar model. oplanets, namely, that both increasing their semi-major axes and the effective temperatures leads to an increase in planetary radii. To further illustrate this, the right panel of Figure 9 shows the equilibrium temperature- radius diagram for the same exoplanets as on the left plot. We can clearly see that the correlation followed by most of the discovered transiting hot Jupiters to date is also followed by our newly discovered exoplanets. In terms of future characterization, all the presented planets (except HATS-27b) have expected transmission signals between ∼ 700 − 900 ppm and all (except HATS- 28) have magnitudes between V ∼ 12 − 13, making them interesting targets for future atmospheric studies. Fig- ure 10 illustrates V band magnitude versus the expected transmission signals for our newly discovered planets along with planets discovered to date, where the formula used to calculate the signal assumes an atmosphere that is five scale-heights thick, and is given by δtranspec = 10RpH R2 ∗ , where Rp is the planetary radius, R∗ is the stellar radius and H = kBTp/mgp is the planetary scale-height, cal- culated using Boltzmann's constant, kB, the planetary equilibrium temperature, Tp, the mean mass of the con- HATS-25b–HATS-30b 15 Orbital and planetary parameters for HATS-25b, HATS-26b and HATS-27b Table 6 Parameter Light curve parameters P (days) . . . . . . . . . . . . . . . . . . . . . Tc (BJD) a . . . . . . . . . . . . . . . . . . . T14 (days) a . . . . . . . . . . . . . . . . . . T12 = T34 (days) a . . . . . . . . . . . a/R⋆ . . . . . . . . . . . . . . . . . . . . . . . . . b . . . . . . . . . . . . . . . . . . . . . . . ζ/R⋆ Rp/R⋆ . . . . . . . . . . . . . . . . . . . . . . . b2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . b ≡ a cos i/R⋆ . . . . . . . . . . . . . . . . i (deg) . . . . . . . . . . . . . . . . . . . . . . . HATSouth blend factors c Blend factor . . . . . . . . . . . . . . . . . . Limb-darkening coefficients d c1, R . . . . . . . . . . . . . . . . . . . . . . . . . c2, R . . . . . . . . . . . . . . . . . . . . . . . . . c1, r . . . . . . . . . . . . . . . . . . . . . . . . . . c2, r . . . . . . . . . . . . . . . . . . . . . . . . . . c1, i . . . . . . . . . . . . . . . . . . . . . . . . . . c2, i . . . . . . . . . . . . . . . . . . . . . . . . . . RV parameters K (m s−1) . . . . . . . . . . . . . . . . . . . . e e . . . . . . . . . . . . . . . . . . . . . . . . . . . RV jitter FEROS (m s−1) f . . . RV jitter HARPS (m s−1) . . . . RV jitter Coralie (m s−1) . . . . . Planetary parameters Mp (MJ) . . . . . . . . . . . . . . . . . . . . . Rp (RJ) . . . . . . . . . . . . . . . . . . . . . . C(Mp, Rp) g . . . . . . . . . . . . . . . . . ρp (g cm−3) . . . . . . . . . . . . . . . . . . log gp (cgs) . . . . . . . . . . . . . . . . . . . a (AU) . . . . . . . . . . . . . . . . . . . . . . . Teq (K) . . . . . . . . . . . . . . . . . . . . . . Θ h . . . . . . . . . . . . . . . . . . . . . . . . . . log10hF i (cgs) i . . . . . . . . . . . . . . . HATS-25b Value HATS-26b Value HATS-27b Value 4.2986432 ± 0.0000045 2456870.36872 ± 0.00051 3.3023881 ± 0.0000076 2456867.4232 ± 0.0012 4.637038 ± 0.000014 2457029.3374 ± 0.0011 0.1335 ± 0.0025 0.0190 ± 0.0027 10.03 ± 0.62 17.39 ± 0.17 0.1171 ± 0.0026 0.290+0.087 −0.102 0.538+0.075 −0.105 86.93 ± 0.71 0.2173 ± 0.0041 0.0196 ± 0.0035 5.01 ± 0.27 10.14 ± 0.13 0.0879 ± 0.0055 0.109+0.098 −0.084 0.33+0.12 −0.17 86.2 ± 1.9 0.2013 ± 0.0033 0.0186 ± 0.0030 0.0895 ± 0.0043 7.55+0.42 −0.59 10.98+0.11 −0.14 0.13+0.13 −0.10 0.36+0.15 −0.19 87.3 ± 1.3 0.959 ± 0.046 0.775 ± 0.077 0.778 ± 0.086 · · · · · · 0.3674 0.3192 0.2774 0.3246 76.8 ± 5.0 < 0.176 · · · < 0.12 · · · 0.613 ± 0.042 1.26 ± 0.10 0.01 0.38 ± 0.10 2.976 ± 0.075 0.05163 ± 0.00060 1277 ± 42 0.0500 ± 0.0054 8.778 ± 0.057 · · · · · · 0.2947 0.3611 0.2145 0.3580 73.3 ± 8.0 < 0.245 < 28 < 9.1 < 8.1 0.650 ± 0.076 1.75 ± 0.21 0.27 0.153 ± 0.042 2.724+0.074 −0.103 0.04735+0.00133 −0.00068 1918 ± 61 0.0264 ± 0.0038 9.485 ± 0.054 0.2295 0.3855 0.2511 0.3857 0.1754 0.3788 51 ± 13 < 0.581 72 ± 17 < 38.0 < 142.2 0.53 ± 0.13 1.50+0.20 −0.11 −0.00 0.180+0.083 −0.057 2.75 ± 0.15 0.06110 ± 0.00068 1659+66 −46 0.0292 ± 0.0081 9.232+0.067 −0.050 Note. - For HATS-25 and HATS-26 the fixed-circular-orbit model has a higher Bayesian evidence than the eccentric-orbit model (it is 10 and 8 times greater for these two systems respectively). We therefore assume a fixed circular orbit in generating the parameters listed for both of these systems. For HATS-27 the free-eccentricity model has an indistinguishable Bayesian evidence from the fixed-circular model, but in this case the eccentricity is poorly constrained with implausibly high values permitted by the low S/N RV measurements. For this system we also adopt the fixed-circular model parameters. a Times are in Barycentric Julian Date calculated directly from UTC without correction for leap seconds. Tc: Reference epoch of mid transit that minimizes the correlation with the orbital period. T14: total transit duration, time between first to last contact; T12 = T34: ingress/egress time, time between first and second, or third and fourth contact. b Reciprocal of the half duration of the transit used as a jump parameter in our MCMC analysis in place of a/R⋆. It is related to a/R⋆ by the expression ζ/R⋆ = a/R⋆(2π(1 + e sin ω))/(P√1 − b2√1 − e2) (Bakos et al. 2010). c Scaling factor applied to the model transit that is fit to the HATSouth light curves. This factor accounts for dilution of the transit due to blending from neighboring stars and over-filtering of the light curve. These factors are varied in the fit, and we allow independent factors for observations obtained with different HATSouth camera and field combinations. d Values for a quadratic law, adopted from the tabulations by Claret (2004) according to the spectroscopic (ZASPE) parameters listed in Table 4. e For fixed circular orbit models we list the 95% confidence upper limit on the eccentricity determined when √e cos ω and √e sin ω are allowed to vary in the fit. f Term added in quadrature to the formal RV uncertainties for each instrument. This is treated as a free parameter in the fitting routine. In cases where the jitter is consistent with zero we list the 95% confidence upper limit. g Correlation coefficient between the planetary mass Mp and radius Rp estimated from the posterior parameter distribution. h The Safronov number is given by Θ = 1 i 2 (Vesc/Vorb)2 = (a/Rp)(Mp/M⋆) (see Hansen & Barman 2007). Incoming flux per unit surface area, averaged over the orbit. 16 Espinoza et al. Orbital and planetary parameters for HATS-28b, HATS-29b and HATS-30b Table 7 Parameter Light curve parameters P (days) . . . . . . . . . . . . . . . . . . . . . Tc (BJD) a . . . . . . . . . . . . . . . . . . . T14 (days) a . . . . . . . . . . . . . . . . . . T12 = T34 (days) a . . . . . . . . . . . a/R⋆ . . . . . . . . . . . . . . . . . . . . . . . . . b . . . . . . . . . . . . . . . . . . . . . . . ζ/R⋆ Rp/R⋆ . . . . . . . . . . . . . . . . . . . . . . . b2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . b ≡ a cos i/R⋆ . . . . . . . . . . . . . . . . i (deg) . . . . . . . . . . . . . . . . . . . . . . . HATSouth blend factors c Blend factor 1 . . . . . . . . . . . . . . . . Blend factor 2 . . . . . . . . . . . . . . . . Blend factor 3 . . . . . . . . . . . . . . . . Limb-darkening coefficients d c1, r . . . . . . . . . . . . . . . . . . . . . . . . . . c2, r . . . . . . . . . . . . . . . . . . . . . . . . . . c1, i . . . . . . . . . . . . . . . . . . . . . . . . . . c2, i . . . . . . . . . . . . . . . . . . . . . . . . . . RV parameters K (m s−1) . . . . . . . . . . . . . . . . . . . . e e . . . . . . . . . . . . . . . . . . . . . . . . . . . RV jitter FEROS (m s−1) f . . . RV jitter HARPS (m s−1) . . . . RV jitter Coralie (m s−1) . . . . . RV jitter CYCLOPS (m s−1) . Planetary parameters Mp (MJ) . . . . . . . . . . . . . . . . . . . . . Rp (RJ) . . . . . . . . . . . . . . . . . . . . . . C(Mp, Rp) g . . . . . . . . . . . . . . . . . ρp (g cm−3) . . . . . . . . . . . . . . . . . . log gp (cgs) . . . . . . . . . . . . . . . . . . . a (AU) . . . . . . . . . . . . . . . . . . . . . . . Teq (K) . . . . . . . . . . . . . . . . . . . . . . Θ h . . . . . . . . . . . . . . . . . . . . . . . . . . log10hF i (cgs) i . . . . . . . . . . . . . . . HATS-28b Value HATS-29b Value HATS-30b Value 3.1810781 ± 0.0000039 2457034.28300 ± 0.00046 4.6058749 ± 0.0000063 2457031.95618 ± 0.00038 3.1743516 ± 0.0000026 2456629.76156 ± 0.00036 0.0981 ± 0.0018 0.0185 ± 0.0020 9.63 ± 0.42 24.84 ± 0.28 0.1331 ± 0.0029 0.414+0.052 −0.059 0.643+0.039 −0.047 86.17 ± 0.42 0.1338 ± 0.0014 0.0186 ± 0.0014 10.96 ± 0.34 17.30 ± 0.12 0.1201 ± 0.0027 0.252+0.049 −0.040 0.502+0.046 −0.041 87.37 ± 0.34 0.893 ± 0.042 0.859 ± 0.039 · · · · · · 0.4137 0.2900 0.3148 0.3030 97 ± 12 < 0.202 32.9 ± 8.7 · · · · · · · · · · · · · · · 0.3875 0.3096 0.2914 0.3213 78.4 ± 7.1 < 0.158 · · · < 4.5 < 0.68 36 ± 11 0.672 ± 0.087 1.194 ± 0.070 0.01 0.48 ± 0.11 3.065 ± 0.076 0.04131 ± 0.00053 1253 ± 35 0.0498 ± 0.0070 8.746 ± 0.048 0.653 ± 0.063 1.251 ± 0.061 0.31 0.411 ± 0.060 3.010 ± 0.049 0.05475 ± 0.00088 1212 ± 30 0.0557 ± 0.0051 8.687 ± 0.044 0.1146 ± 0.0012 0.0150 ± 0.0012 8.82 ± 0.31 19.99 ± 0.13 0.1137 ± 0.0017 0.237+0.053 −0.056 0.487+0.052 −0.061 86.84 ± 0.48 0.963 ± 0.027 0.823 ± 0.031 0.967 ± 0.027 0.3275 0.3438 0.2450 0.3430 91.8 ± 4.7 < 0.096 < 4.7 · · · < 25 · · · 0.706 ± 0.039 1.175 ± 0.052 0.10 0.543 ± 0.076 3.105 ± 0.044 0.04354 ± 0.00042 1414 ± 32 0.0478 ± 0.0033 8.955 ± 0.039 Note. - For all three systems the fixed-circular-orbit model has a higher Bayesian evidence than the eccentric-orbit model (it is 5, 660, and 3 times greater for HATS-28, HATS-29 and HATS-30, respectively). We therefore assume a fixed circular orbit in generating the parameters listed for these systems. a Times are in Barycentric Julian Date calculated directly from UTC without correction for leap seconds. Tc: Reference epoch of mid transit that minimizes the correlation with the orbital period. T14: total transit duration, time between first to last contact; T12 = T34: ingress/egress time, time between first and second, or third and fourth contact. b Reciprocal of the half duration of the transit used as a jump parameter in our MCMC analysis in place of a/R⋆. It is related to a/R⋆ by the expression ζ/R⋆ = a/R⋆(2π(1 + e sin ω))/(P√1 − b2√1 − e2) (Bakos et al. 2010). c Scaling factor applied to the model transit that is fit to the HATSouth light curves. This factor accounts for dilution of the transit due to blending from neighboring stars and over-filtering of the light curve. These factors are varied in the fit, and we allow independent factors for observations obtained with different HATSouth camera and field combinations. For HATS-30 blend factors 1 through 3 are used for the G754.3, G754.4 and G755.1 observations, respectively. d Values for a quadratic law, adopted from the tabulations by Claret (2004) according to the spectroscopic (ZASPE) parameters listed in Table 4. e For fixed circular orbit models we list the 95% confidence upper limit on the eccentricity determined when √e cos ω and √e sin ω are allowed to vary in the fit. f Term added in quadrature to the formal RV uncertainties for each instrument. This is treated as a free parameter in the fitting routine. In cases where the jitter is consistent with zero we list the 95% confidence upper limit. g Correlation coefficient between the planetary mass Mp and radius Rp estimated from the posterior parameter distribution. h The Safronov number is given by Θ = 1 i 2 (Vesc/Vorb)2 = (a/Rp)(Mp/M⋆) (see Hansen & Barman 2007). Incoming flux per unit surface area, averaged over the orbit. HATS-25b–HATS-30b 17 ) J R ( i s u d a R t e n a l P 2.25 2 1.75 1.5 1.25 1 0.75 0.5 HS-26b HS-27b HS-25b HS-28b HS-29b HS-30b 2.25 2 1.75 1.5 1.25 1 0.75 ) J R ( i s u d a R t e n a l P HS-26b HS-27b HS-29b HS-25b HS-28b HS-30b 0.5 1 2 5 Planet Mass (MJ ) 0.5 500 1000 1500 2000 2500 3000 Equilibrium Temperature (K) Figure 9. (Left) Mass-radius diagram for all the transiting hot Jupiters discovered to date (grey points). Red points indicate the discovered exoplanets presented in this work. The black lines show the mass-radius relations of 4.5 Gyr old planets at 0.045 AU from the Sun obtained from Fortney et al. (2007) for core-free giant planets (solid line) and for planets with 100M⊕ cores (dashed line), which are appropriate for the insolation levels received by HATS-25b, HATS-28b, HATS-29b and HATS-30b. The blue lines show the same relations but for planets at 0.02 AU, more (but not exactly) appropriate for the insolation levels received by HATS-26b and HATS-27b. We note, however, that these relations imply insolation levels around 2500 times the solar insolation level at Earth, while the actual insolation levels for HATS-26b and HATS-27b are closer to 2250 and 1250 times the solar flux at Earth, respectively. (Right) Equilibrium temperature-radius diagram for all the transiting hot Jupiters discovered to date along with the discovered exoplanets presented in this work with the same colors as in the left plot. stituents that make up the atmosphere of the planet (as- sumed to be H2), m, and the acceleration due to gravity on the planetary surface, gp. Systems already character- ized by transmission spectroscopy are indicated in blue. As can be seen, the discovered exoplanets add to the increasing fraction of planets that have expected trans- mission signals on the same order as those already char- acterized. The most interesting systems in this respect are HATS-26b (V = 12.9), which has an expected trans- mission signal of ∼ 900 ppm and a long transit duration of 5.2 hours, and HATS-29b (V = 12.6), which has an ex- pected transmission signal of ∼ 700 ppm, a transit depth two times that of HATS-26b and a transit duration of 3.2 hours. Although not a good target for transmission, HATS- 27b (V = 12.8) is an attractive system if one is inter- ested in estimating the projected spin-orbit alignment of the system: despite its modest planet-to-star ratio of (Rp/R∗ = 0.0895 ± 0.0041), the host star rotates at a moderately high rate (v sin(i) of 9.32 ± 0.5 km/s) which, coupled with the long transit duration of 4.8 hours, makes this inflated hot Jupiter a good target for follow- up Rossiter-McLaughlin (RM) observations. In partic- ular, using equation (6) of Gaudi & Winn (2007), the amplitude of the RM effect, KR, should be ≈ 75 m/s. We obtained a precision of ∼ 30 m/s in 10 minute expo- sures with HARPS for this star, making the RM effect readily detectable. In addition, given that the tempera- ture of the host star is 6428 ± 64 K, the system lies in a very interesting regime at which it has been claimed that planetary orbits of hot Jupiters shift from aligned to misaligned (Albrecht et al. 2012; Addison et al. 2016). Development of the HATSouth project was funded by NSF MRI grant NSF/AST-0723074, operations have been supported by NASA grants NNX09AB29G and NNX12AH91H, and follow-up observations receive par- ) m p p ( l a n g i S n o i s s i m s n a r T 2000 1500 1000 500 0 0 ppt 10 ppt 21 ppt 31 ppt HS-26b HS-28b HS-29b HS-25b HS-30b HS-27b 8 9 10 V Magnitude 11 12 13 14 15 Figure 10. Visual magnitude versus expected transmission signal for all the hot Jupiters discovered to date (grey points). Blue points indicate systems that have already been characterised via transmission spectroscopy, while red points indicate the exoplanets presented in this work. The size of the points indicate the transit depth with larger points indicating larger transit depths; the legend in the upper left corner indicates the corresponding depths in parts per thousand (ppt). tial support from grant NSF/AST-1108686. N.E. is supported by CONICYT-PCHA/Doctorado Nacional. A.J. acknowledges support from FONDECYT project 1130857, BASAL CATA PFB-06, and project IC120009 "Millennium Institute of Astrophysics (MAS)" of the Millenium Science Initiative, Chilean Ministry of Econ- omy. R.B. and N.E. acknowledge support from project IC120009 "Millenium Institute of Astrophysics (MAS)" of the Millennium Science Initiative, Chilean Ministry of Economy. V.S. acknowledges support form BASAL CATA PFB-06. This work is based on observations made with ESO Telescopes at the La Silla Observatory. This paper also uses observations obtained with facilities of the Las Cumbres Observatory Global Telescope. Work at 18 Espinoza et al. the Australian National University is supported by ARC Laureate Fellowship Grant FL0992131. Work at UNSW is supported by ARC Discovery Project DP130102695. We acknowledge the use of the AAVSO Photometric All-Sky Survey (APASS), funded by the Robert Martin Ayers Sciences Fund, and the SIMBAD database, op- erated at CDS, Strasbourg, France. Operations at the MPG 2.2 m Telescope are jointly performed by the Max Planck Gesellschaft and the European Southern Obser- vatory. The imaging system GROND has been built by the high-energy group of MPE in collaboration with the LSW Tautenburg and ESO. We thank the MPG 2.2 m telescope support team for their technical assistance dur- ing observations. We thank Helmut Steinle and Jochen Greiner for supporting the GROND observations pre- sented in this manuscript. Observing time were ob- tained through proposals CN2013-B55, CN2014A-104, CN2014B-57, CN2015A-51 and ESO 096.C-0544. We are grateful to P.Sackett for her help in the early phase of the HATSouth project. HATS-25b–HATS-30b 19 REFERENCES Hippler, S., Bergfors, C., Brandner Wolfgang, et al. 2009, The Addison, B. C., Tinney, C. G., Wright, D. J., & Bayliss, D. 2016, ArXiv e-prints, 1603.05754 Addison, B. C., Tinney, C. G., Wright, D. J., et al. 2013, ApJ, 774, L9 Albrecht, S., Winn, J. N., Johnson, J. A., et al. 2012, ApJ, 757, 18 Bakos, G. ´A., Torres, G., P´al, A., et al. 2010, ApJ, 710, 1724 Bakos, G. ´A., Csubry, Z., Penev, K., et al. 2013, PASP, 125, 154 Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, A&A, 402, 701 Batygin, K., & Stevenson, D. J. 2010, ApJ, 714, L238 Batygin, K., Stevenson, D. J., & Bodenheimer, P. H. 2011, ApJ, 738, 1 Bayliss, D., Zhou, G., Penev, K., et al. 2013, AJ, 146, 113 Bayliss, D., Hartman, J. D., Bakos, G. ´A., et al. 2015, AJ, 150, 49 Benneke, B. 2015, ArXiv e-prints, 1504.07655 Brown, T. M., Baliber, N., Bianco, F. B., et al. 2013, PASP, 125, 1031 Burrows, A., Hubeny, I., Budaj, J., & Hubbard, W. B. 2007, ApJ, 661, 502 Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Charbonneau, D., Brown, T. M., Latham, D. W., & Mayor, M. 2000, ApJ, 529, L45 Claret, A. 2004, A&A, 428, 1001 Crossfield, I. J. M. 2015, PASP, 127, 941 Dopita, M., Hart, J., McGregor, P., et al. 2007, Ap&SS, 310, 255 Enoch, B., Collier Cameron, A., & Horne, K. 2012, A&A, 540, A99 Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298 Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661 Fruchter, A. S., & Hook, R. N. 2002, PASP, 114, 144 Gaudi, B. S., & Winn, J. N. 2007, ApJ, 655, 550 Girardi, L., Bressan, A., Bertelli, G., & Chiosi, C. 2000, A&AS, 141, 371 Goldreich, P., & Tremaine, S. 1980, ApJ, 241, 425 Hansen, B. M. S., & Barman, T. 2007, ApJ, 671, 861 Hartman, J. D., Bakos, G. ´A., B´eky, B., et al. 2012, AJ, 144, 139 Hartman, J. D., Bayliss, D., Brahm, R., et al. 2015, AJ, 149, 166 Henry, G. W., Marcy, G. W., Butler, R. P., & Vogt, S. S. 2000, ApJ, 529, L41 Messenger, 137, 14 Huang, X., & Cumming, A. 2012, ApJ, 757, 47 Jord´an, A., Brahm, R., Bakos, G. ´A., et al. 2014, AJ, 148, 29 Kataria, T., Sing, D. K., Lewis, N. K., et al. 2016, ArXiv e-prints, 1602.06733 Kaufer, A., & Pasquini, L. 1998, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 3355, Optical Astronomical Instrumentation, ed. S. D'Odorico, 844–854 Kislyakova, K. G., Holmstrom, M., Lammer, H., Odert, P., & Khodachenko, M. L. 2014, Science, 346, 981 Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369 Lissauer, J. J., & Stevenson, D. J. 2007, Protostars and Planets V, 591 Louden, T., & Wheatley, P. J. 2015, ApJ, 814, L24 Madhusudhan, N., Amin, M. A., & Kennedy, G. M. 2014, ApJ, 794, L12 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Messenger, 114, 20 Ohta, Y., Taruya, A., & Suto, Y. 2005, ApJ, 622, 1118 P´al, A., Bakos, G. ´A., Torres, G., et al. 2008, ApJ, 680, 1450 Perna, R., Menou, K., & Rauscher, E. 2010, ApJ, 724, 313 Petrovich, C. 2015, ApJ, 805, 75 Queloz, D., Eggenberger, A., Mayor, M., et al. 2000, A&A, 359, L13 Queloz, D., Mayor, M., Udry, S., et al. 2001, The Messenger, 105, 1 Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954 Sing, D. K., Fortney, J. J., Nikolov, N., et al. 2016, Nature, 529, 59 Sozzetti, A., Torres, G., Charbonneau, D., et al. 2007, ApJ, 664, 1190 Spiegel, D. S., & Burrows, A. 2013, ApJ, 772, 76 Winn, J. N. 2007, in Astronomical Society of the Pacific Conference Series, Vol. 366, Transiting Extrapolar Planets Workshop, ed. C. Afonso, D. Weldrake, & T. Henning, 170 Wu, Y., & Lithwick, Y. 2011, ApJ, 735, 109 -. 2013, ApJ, 763, 13 Yi, S., Demarque, P., Kim, Y.-C., et al. 2001, ApJS, 136, 417 Zacharias, N., Finch, C. T., Girard, T. M., et al. 2012, VizieR Online Data Catalog, 1322, 0 Zechmeister, M., & Kurster, M. 2009, A&A, 496, 577 Zhou, G., Bayliss, D., Penev, K., et al. 2014, ArXiv e-prints, 1401.1582 20 Espinoza et al. Table 8 Relative radial velocities and bisector spans for HATS-25–HATS-30. Star BJD (2,450,000+) RVa (m s−1) b σRV (m s−1) BS σBS Phase Instrument (m s−1) (m s−1) HATS-25 HATS-25 HATS-25 HATS-25 HATS-25 HATS-25 HATS-25 HATS-25 7067.85231 7067.87380 7068.86477 7070.85785 7071.87936 7072.88746 7118.73269 7120.73472 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-26 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-27 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 HATS-28 6828.49681 6829.52934 7031.72967 7035.82606 7037.84686 7049.79153 7050.84666 7053.88112 7054.81498 7056.81639 7067.70058 7069.77078 7070.73942 7072.71165 6828.57385 6828.62287 6829.58090 6841.56122 6842.51976 6845.58436 6846.47434 6847.47811 6850.59743 6851.54418 6852.48123 6852.58575 6854.49043 6855.47772 6856.49742 7067.80560 7068.84310 7069.87176 7070.83755 7071.86649 7072.87384 7118.60461 7119.69411 7119.76445 7120.70644 7121.56784 7466.59338 7467.57122 7468.57169 7181.60951 7182.79396 7183.58139 7184.65002 7187.76095 7188.73064 7189.88679 7191.66177 7192.67821 7193.76250 7194.62609 7196.85775 7218.71803 7220.77528 7223.54993 7224.55108 7227.52090 34.25 4.25 −70.75 57.25 54.25 −48.75 80.25 −77.75 −65.10 2.90 58.90 34.90 22.90 −56.10 −1.10 −3.10 97.90 −65.10 68.91 −92.09 41.91 −72.09 29.59 154.59 16.59 50.37 22.37 −47.63 30.37 28.37 54.37 −33.63 −16.63 63.37 −92.63 −37.63 29.37 −38.64 −24.64 74.36 −18.64 −100.64 −1.64 −45.64 224.37 68.37 105.36 33.37 −6.64 25.36 62.36 −138.64 124.36 54.36 −21.64 −58.64 58.36 86.36 −53.64 88.36 −67.64 −48.64 −96.64 17.36 55.36 −89.64 73.36 114.36 HATS-25 10.00 11.00 7.00 9.00 21.00 9.00 16.00 11.00 HATS-26 28.00 30.00 15.00 16.00 15.00 16.00 15.00 17.00 14.00 17.00 18.00 20.00 18.00 16.00 HATS-27 40.00 41.00 37.00 21.00 17.00 17.00 26.00 14.00 23.00 18.00 18.00 22.00 20.00 15.00 13.00 25.00 15.00 24.00 20.00 40.00 25.00 23.00 26.00 16.00 27.00 15.00 33.00 27.00 27.00 −19.0 −2.0 −38.0 −27.0 15.0 −23.0 −36.0 −14.0 150.0 80.0 30.0 68.0 34.0 71.0 64.0 59.0 32.0 69.0 67.0 8.0 92.0 28.0 24.0 107.0 79.0 14.0 71.0 48.0 −69.0 24.0 8.0 11.0 74.0 56.0 48.0 99.0 32.0 73.0 18.0 −6.0 30.0 58.0 −29.0 247.0 468.0 147.0 55.0 100.0 27.0 37.0 47.0 HATS-28 18.00 −149.0 22.00 25.0 22.00 −35.0 22.00 72.0 −36.0 15.00 −94.0 14.00 13.00 −5.0 12.00 −33.0 2.0 19.00 1.0 11.00 45.0 12.00 −38.0 15.00 −13.0 25.00 20.00 4.0 −25.0 24.00 14.00 −46.0 −46.0 10.00 40.0 40.0 28.0 32.0 70.0 36.0 50.0 40.0 24.0 26.0 12.0 13.0 12.0 12.0 12.0 13.0 11.0 13.0 38.0 42.0 38.0 34.0 29.0 29.0 27.0 14.0 12.0 12.0 15.0 10.0 14.0 12.0 12.0 14.0 13.0 11.0 10.0 48.0 30.0 48.0 38.0 72.0 48.0 44.0 16.0 11.0 48.0 11.0 48.0 38.0 38.0 25.0 30.0 30.0 30.0 20.0 18.0 18.0 16.0 25.0 15.0 16.0 20.0 34.0 26.0 32.0 19.0 14.0 0.941 HARPS 0.946 HARPS 0.176 HARPS 0.640 HARPS 0.878 HARPS 0.112 HARPS 0.777 HARPS 0.243 HARPS 0.213 Coralie 0.526 Coralie 0.754 FEROS 0.995 FEROS 0.607 FEROS 0.224 FEROS 0.543 FEROS 0.462 FEROS 0.745 FEROS 0.351 FEROS 0.647 HARPS 0.273 HARPS 0.567 HARPS 0.164 HARPS 0.704 Coralie 0.715 Coralie 0.922 Coralie 0.505 FEROS 0.712 FEROS 0.373 FEROS 0.565 FEROS 0.781 FEROS 0.454 FEROS 0.658 FEROS 0.860 FEROS 0.883 FEROS 0.293 FEROS 0.506 FEROS 0.726 FEROS 0.296 HARPS 0.520 HARPS 0.741 HARPS 0.950 HARPS 0.172 HARPS 0.389 HARPS 0.251 HARPS 0.486 FEROS 0.501 FEROS 0.704 HARPS 0.890 FEROS 0.296 HARPS 0.507 HARPS 0.723 HARPS 0.313 FEROS 0.686 FEROS 0.933 FEROS 0.269 FEROS 0.247 FEROS 0.552 FEROS 0.915 FEROS 0.473 FEROS 0.793 FEROS 0.134 FEROS 0.405 FEROS 0.107 FEROS 0.979 FEROS 0.625 FEROS 0.498 FEROS 0.812 FEROS 0.746 FEROS HATS-25b–HATS-30b Table 8 Relative radial velocities and bisector spans for HATS-25–HATS-30. 21 HATS-28 7230.76298 79.36 11.00 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-29 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 HATS-30 7118.83135 7119.83810 7120.82811 7149.24168 7149.25763 7149.27360 7150.27180 7150.28776 7150.30372 7152.13921 7152.15453 7152.16986 7179.75812 7180.75107 7181.76069 7182.73909 6932.62878 6939.66985 6940.55468 6941.71844 6968.73018 6970.67182 6972.60936 6982.70613 6984.64881 6985.58892 6997.56102 6998.61894 6999.66282 44.32 −26.68 −75.68 −72.08 −22.98 −11.78 96.82 11.02 127.52 −1.98 −36.58 −15.58 −28.10 −88.10 13.90 87.90 −59.90 86.61 31.61 −64.39 70.61 −57.39 −27.39 −83.90 89.10 −49.90 67.10 −97.90 31.10 19.00 12.00 7.00 16.80 17.70 7.20 16.30 17.20 12.80 9.70 8.00 11.00 15.00 16.00 13.00 14.00 HATS-30 10.00 15.00 14.00 15.00 12.00 15.00 14.00 10.00 10.00 10.00 11.00 10.00 10.00 −35.0 −66.0 −42.0 −41.0 · · · · · · · · · · · · · · · · · · · · · · · · · · · 12.0 92.0 −15.0 −73.0 4.0 −10.0 −5.0 23.0 −43.0 −77.0 −1.0 19.0 34.0 29.0 −12.0 13.0 −8.0 16.0 0.765 FEROS 60.0 44.0 28.0 · · · · · · · · · · · · · · · · · · · · · · · · · · · 26.0 29.0 24.0 26.0 11.0 19.0 18.0 21.0 15.0 19.0 19.0 11.0 11.0 11.0 13.0 10.0 11.0 0.862 HARPS 0.080 HARPS 0.295 HARPS 0.464 CYCLOPS 0.468 CYCLOPS 0.471 CYCLOPS 0.688 CYCLOPS 0.691 CYCLOPS 0.695 CYCLOPS 0.093 CYCLOPS 0.097 CYCLOPS 0.100 CYCLOPS 0.090 Coralie 0.305 Coralie 0.525 Coralie 0.737 Coralie 0.411 FEROS 0.629 Coralie 0.908 Coralie 0.274 Coralie 0.784 Coralie 0.395 Coralie 0.006 Coralie 0.186 FEROS 0.798 FEROS 0.095 FEROS 0.866 FEROS 0.199 FEROS 0.528 FEROS a The zero-point of these velocities is arbitrary. An overall offset γrel fitted independently to the velocities from each instrument has been subtracted. b Internal errors excluding the component of astrophysical jitter considered in Section 3.3. ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) g a m ∆ g a m ∆ -0.03 -0.02 -0.01 0 0.01 0.02 0.03 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 HATS-27 -0.4 -0.2 0 0.2 0.4 Orbital phase -0.06 -0.04 -0.02 0 0.02 0.04 0.06 Orbital phase HATS-27 0.0 0.5 1.0 1.5 ] 3 m c / g [ * ρ 2.0 7000 6800 6600 6400 6200 6000 Effective temperature [K] HATS-27 RC-band 2015 Mar 12 (PEST 0.3m) i 2015 Apr 9 (LCOGT 1m) s t e s f f o y r a r t i b r A - ) g a m ( ∆ 0 0.02 0.04 0.06 0.08 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 Time from transit center (days) ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) HATS-27 200 150 100 FEROS Coralie HARPS ) 1 - s m ( V R 50 0 -50 -100 -150 -200 300 250 200 150 100 50 0 -50 -100 -150 500 400 300 200 100 0 -100 ) 1 - s m ( - C O ) 1 - s m ( S B 0.0 0.2 0.4 Phase with respect to Tc 0.6 0.8 1.0 ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days) ) g a m ( ∆ -0.01 0 0.01 0.02 0.03 0.04 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 Time from transit center (days)
1802.09602
1
1802
2018-02-26T20:56:42
Exoplanet Classification and Yield Estimates for Direct Imaging Missions
[ "astro-ph.EP" ]
Future NASA concept missions that are currently under study, like Habitable Exoplanet Imaging Mission (HabEx) & Large Ultra-Violet Optical Infra Red (LUVOIR) Surveyor, would discover a large diversity of exoplanets. We propose here a classification scheme that distinguishes exoplanets into different categories based on their size and incident stellar flux, for the purpose of providing the expected number of exoplanets observed (yield) with direct imaging missions. The boundaries of this classification can be computed using the known chemical behavior of gases and condensates at different pressures and temperatures in a planetary atmosphere. In this study, we initially focus on condensation curves for sphalerite ZnS, H2O, CO2 and CH4. The order in which these species condense in a planetary atmosphere define the boundaries between different classes of planets. Broadly, the planets are divided into rocky (0.5 - 1.0RE), super-Earths (1.0- 1.75RE), sub-Neptunes (1.75-3.5RE), sub-Jovians (3.5 - 6.0RE) and Jovians (6-14.3RE) based on their planet sizes, and 'hot', 'warm' and 'cold' based on the incident stellar flux. We then calculate planet occurrence rates within these boundaries for different kinds of exoplanets, \eta_{planet}, using the community co-ordinated results of NASA's Exoplanet Program Analysis Group's Science Analysis Group-13 (SAG-13). These occurrence rate estimates are in turn used to estimate the expected exoplanet yields for direct imaging missions of different telescope diameter.
astro-ph.EP
astro-ph
Exoplanet Classification and Yield Estimates for Direct Imaging Missions Ravi kumar Kopparapu1,2,3,4, Eric H´ebrard1,5, Rus Belikov6, Natalie M. Batalha6, Gijs D. Mulders7,8, Chris Stark9, Dillon Teal1, Shawn Domagal-Goldman1,3, Avi Mandell1 ABSTRACT Future NASA concept missions that are currently under study, like Habitable Exoplanet Imaging Mission (HabEx) & Large Ultra-Violet Optical Infra Red (LUVOIR) Surveyor, would discover a large diversity of exoplanets. We propose here a classification scheme that distinguishes exoplanets into different categories based on their size and incident stellar flux, for the purpose of providing the expected number of exoplanets observed (yield) with direct imaging missions. The boundaries of this classification can be computed using the known chemical behavior of gases and condensates at different pressures and temperatures in a planetary atmosphere. In this study, we initially focus on condensation curves for sphalerite ZnS, H2O, CO2 and CH4. The order in which these species condense in a planetary atmosphere define the boundaries between different classes of planets. Broadly, the planets are divided into rocky (0.5 − 1.0R⊕), super-Earths (1.0 − 1.75R⊕), sub-Neptunes (1.75 − 3.5R⊕), sub-Jovians (3.5 − 6.0R⊕) and Jovians (6 − 14.3R⊕) based on their planet sizes, and 'hot', 'warm' and 'cold' based on the incident stellar flux. We then calculate planet occurrence rates within these boundaries for different kinds of exoplanets, ηplanet, using the community 1NASA Goddard Space Flight Center, 8800 Greenbelt Road, Mail Stop 699.0, Building 34, Greenbelt, MD 20771 2Department of Astronomy, University of Maryland College Park, College Park, MD 3NASA Astrobiology Institute's Virtual Planetary Laboratory, P.O. Box 351580, Seattle, WA 98195, USA 4Blue Marble Space Institute of Science, 1001 4th Ave, Suite 3201, Seattle, Washington 98154, USA 5School of Physics and Astronomy, University of Exeter, EX4 4QL, Exeter, UK 6NASA Ames Research Center 7Lunar and Planetary Laboratory, University of Arizona, Tucson, AZ 85721, USA 8Earths in Other Solar Systems Team, NASA Nexus for Exoplanet System Science, USA 9Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218 – 2 – co-ordinated results of NASA's Exoplanet Program Analysis Group's Science Analysis Group-13 (SAG-13). These occurrence rate estimates are in turn used to estimate the expected exoplanet yields for direct imaging missions of different telescope diameter. Subject headings: planets and satellites: atmospheres 1. Introduction The discoveries of exoplanets over the last two decades has revealed planetary bodies of various sizes and masses around other stars (Rowe et al. 2014; Anglada-Escud´e et al. 2016; Coughlin et al. 2016; Kane et al. 2016; Morton et al. 2016; Gillon et al. 2017; Dittman et al. 2017). More specifically, the location of these exoplanets around their host star has a signif- icant influence not only on the prospects of their detectability, but also on the atmospheric chemical composition and the capability of characterizing these atmospheres. Furthermore, the relatively large sample of nearly 3400 confirmed planets and another ∼ 4700 planet candidates to-date1 has enabled us to calculate exoplanet occurrence rates in our galaxy (Catanzarite & Shao 2011; Traub 2012; Howard et al. 2012; Bonfils et al. 2013; Dressing & Charbonneau 2013; Petigura et al. 2013; Kopparapu 2013; Gaidos 2013; Fressin et al. 2013; Dong & Zhu 2013; Foreman-Mackey et al. 2014; Morton & Swift 2014; Silburt et al. 2015; Dressing & Charbonneau 2015; Burke et al. 2015; Mulders et al. 2015). These initial esti- mates are dominated by close-in planets due to the sensitivity of the detection techniques and search pipelines. Nevertheless, these studies made a crucial and a significant leap in understanding planet diversity, and paved a way for comparative planetology of exoplanets. Several of the above mentioned studies have also focused on obtaining an estimate of the fraction of stars that have at least one terrestrial mass/size planet in the habitable zone (HZ), or η⊕. Estimates of η⊕ for Sun-like stars have been calculated by the data collected from the Kepler mission. Earlier estimates ranged from 0.02 (Foreman-Mackey et al. 2014) to 0.22 (Petigura et al. 2013) for GK dwarfs, but more recent analyses (Burke et al. 2015) imply that systematic errors dominate. For M-dwarfs, η⊕ is estimated to be ∼ 20% on an average (Dressing & Charbonneau 2015). Apart from the general curiosity of finding how common are Earth-like planets in our galaxy, the focus on η⊕ has a more practical application: It can be used in the design of direct imaging missions, like the concept studies under consideration 1http://exoplanetarchive.ipac.caltech.edu/ – 3 – HabEx2, the Habitable Exoplanet Explorer; and LUVOIR3, the Large UV-Optical-InfraRed surveyor with the goal of detecting biosignatures, and also in calculating 'exo-Earth candidate yield', the number of potentially habitable extrasolar planets (exoEarth candidates) that can be detected and spectroscopically characterized (e.g., Stark et al. 2014, 2015). Crucial to these estimates are the location of the main-sequence HZ, which have been studied by both 1-D and 3-D climate models (Kasting et al. 1993; Selsis et al. 2007b; Abe et al. 2011; Pierrehumbert & Gaidos 2011; Kopparapu et al. 2013; Leconte et al. 2013a; Yang et al. 2013; Zsom et al. 2013; Kopparapu et al. 2014; Wolf & Toon 2014; Yang et al. 2014a; Wolf & Toon 2015; Way et al. 2015; Godolt et al. 2015; Leconte et al. 2015; Kopparapu et al. 2016; Haqq-Misra et al. 2016; Ramirez & Kaltenegger 2017; Kopparapu et al. 2017). With some exceptions for certain types of planets (Kane et al. 2014), there has not been an overarching way to classify planets beyond the HZ. The lack of a systematic way to classify exoplanets in general, combined with the allure of planets within it, has led to direct imag- ing mission yield analyses that focus on HZ planets to the exclusion of everything else (e.g., Stark et al. 2014, 2015). While some mission studies have attempted to classify the non-HZ planets into hot, warm and cold planets that a mission would discover4, the boundaries for the classification are arbitrarily fixed without giving consideration to chemical behavior of gases and condensates in a planetary atmosphere. Classifying different size planets based on the transition/condensation of different species (Burrows et al. 2004; Burrows 2005) pro- vides a physical motivation in estimating exoplanet mission yields, separate from exo-Earth candidate yields. In the search for exo-Earth candidates, we will undoubtedly detect a multitude of brighter planets. According to Stark et al. (2014), for an 8m size telescope, the number of exo-Earth candidates detected is ∼ 20 (see Fig. 4 in Stark et al. 2014), although this is strongly dependent on the value of η⊕. At the same time, the number of stars observed to detect these exo-Earth candidates is ∼ 500. If we assume that, on an average, every star has a planet of some size (Cassan et al. 2012; Suzuki et al. 2016), then there are ∼ 500 exoplanets of all sizes that can be observed. Not considering the ∼ 20 exo-Earth candidates, the bulk of the exo-planets will fall into 'non-Earth' classification, without any distinguishing features between them. This provides an additional motivation to devise a scheme based on planetary size and corresponding atmospheric characteristics of exoplanets. Some work has been done at the theoretical level to derive the radiative response of 2http://www.jpl.nasa.gov/habex/ 3https://asd.gsfc.nasa.gov/luvoir/ 4https://exoplanets.nasa.gov/exep/studies/probe-scale-stdt/ – 4 – an irradiated atmosphere (Robinson & Catling 2012; Parmentier & Guillot 2014; Robinson & Catling 2014; Parmentier et al. 2015), but these analytical tools have not been either used to derive any general boundaries nor tied to planet occurence estimates, and were not designed with that intent in mind. This highlights the need for a theory-based system to classify planets beyond the habitable zone for the purposes of understanding the diversity of worlds future missions could explore. Such a system should be based on the properties we can measure today, primarily size and orbital information, and the boundaries in the scheme should divide planets with major differences in the properties that would be observable with current and future missions. Fortunately, much of the theory needed for such a classification scheme already exists. There has been significant progress in understanding how the size of a planet is a major control on composition, and therefore on future observables (Rogers & Seager 2010a,b). This includes work on the relationship between size, density, and bulk composition (Fortney et al. 2007; Weiss & Marcy 2014; Rogers 2015; Wolfgang et al. 2016; Chen & Kipping 2017). While the exact values for the boundaries of mass-radius change based on the specific analysis or theoretical technique, there is growing evidence of structure in the occurrence rate distribution that suggests compositional aggregates: 1) small, rocky worlds whose bulk composition and behavior is dominated by Fe, Mg and Si species; 2) planets with Fe/Mg/Si cores but significant gas envelopes consisting of H/He, CH4, NH3 ices; and 3) gas giants whose bulk composition and behavior is dictated almost exclusively by its volatiles. Similarly, there have been significant advances in our understanding of how the orbital separation of non-HZ planets could affect the chemical composition of the atmosphere (Ca- hoy et al. 2010). A constant theme across these studies is the influence of clouds. As a planet moves further from its host star, its atmosphere will cool and lead to condensation of progressively less volatile chemicals in the atmosphere. This condensation would create a cold trap and an associated cloud deck. The result of this is a significant change in spectral properties: as the condensing species would be trapped at or below the cloud deck, the cloud deck itself would absorb and scatter light, causing preferential sampling of the layers at or above the cloud deck. Multiple "onion-like" cloud decks can form as sequentially less volatile species condense at higher altitudes for planets with greater star-planet separation distances and correspondingly lower levels of incoming stellar flux. This process can be ob- served in detail in the gas giants of our own Solar System (Evans & Hubbard 1972), and has been postulated to be a driver for the atmospheric structure and observable properties of exoplanet atmospheres (Burrows & Sharp 1999; Sudarsky et al. 2003; Burrows et al. 2004; Burrows 2005; Fortney 2005; Marley et al. 2007; Morley et al. 2013; Wakeford & Singh 2015; Wakeford et al. 2017). We also note that the habitable zone itself has been defined in a manner consistent with this, as the instellations (stellar incident flux) at which liquid water – 5 – clouds form but carbon dioxide clouds do not (Abe et al. 2011). This prior work enables the construction of an overarching scheme for identifying classes of planets. This scheme could apply to all worlds, regardless of whether they are rocky or gaseous (or something in between). And it would be both based on the current observable properties, and prediction of major transitions in future observables. In short, this represents a comprehensive means of predicting the diversity yields of future planet characterization missions. Below we discuss in more detail how we simulate the processes underlying this clas- sification scheme. These simulations define the boundaries between different planet classes, for which we calculate occurrence rates based on prior exoplanet detection missions. The occurrence rates allow us to simulate exoplanet yields - not just for HZ planets but for a diversity of different kinds of worlds. Finally, we close with a discussion of the caveats of this approach, and the implications of this scheme for future missions. 2. A New Classification Scheme A planet size - and the relationship between its size and mass - appears to be primarily driven by volatile inventory. For example, the atmospheric composition of larger planets is predominantly H2/He, while smaller planets can have a mixture of CH4, CO2, H2O and NH3. High-temperature atmospheres, such as hot Jupiters, should have their chemistry - and therefore their spectral features - determined primarily by equilibrium chemistry. Low- temperature atmospheres will have chemistry dictated by photochemistry, but this will be secondary to determining what species are condensing in their atmospheres. The exception to this - which we will discuss later - is for photochemical aerosols, which could have a major impact in the same manner that clouds do. The chemical behavior of gases and condensates in a planetary atmosphere can be de- termined as a function of pressure, temperature, and metallicity. Using results from Lodders and Fegley (2002) and Visscher et al. (2006) adapted with solar abundances taken from Lodders (2010), we have computed the condensation curves for sphalerite ZnS, H2O, CO2 and CH4 as a function of pressure and temperature for systems with a solar metallicity. Pressure-temperature profiles of planetary atmospheres are tightly related to the incom- ing stellar flux. We define the boundaries between our different selected planetary cases as the stellar fluxes for which these four species condense out. For instance, ZnS clouds have been considered as possible condensates in hot exoplanet atmospheres (Morley et al. 2012; Charnay et al. 2015), so the location (or the stellar flux) at which ZnS clouds form in a planetary atmosphere denotes the first boundary for our planet classification. Moving – 6 – further away from the star, at relatively lower stellar fluxes, H2O starts condensing in the atmosphere. This, then, becomes the next boundary for grouping different planets. Between these two boundaries is where one would expect to find ZnS mineral clouds and H2O in a gaseous state. Continuing to lower fluxes, CO2 and CH4 condensates bracket final bound- aries. The results are independent of any particular model atmospheres, and in principle, any pressure-temperature profile may be superimposed on these condensation curves to find the equilibrium composition along the profile. In particular, the intersection of a particular pressure-temperature profile with one of these condensation curves indicates the pressure and temperature at which the respective species condenses out in the planetary atmosphere considered. We have investigated the different incident stellar fluxes, or instellations, for which condensates could form within the regions of an exoplanet system that will be probed by future direct imaging missions. Specifically, we simulate the star-planet separations for which ZnS, H2O, CO2 and CH4 would condense out in planetary atmospheres. Other metal- lic clouds can condense at distances closer to the star than the ZnS condensation line, and other volatiles (e.g., NH3) can condense at orbiting distances beyond the CH4 condensation line, but such worlds are likely undetectable by future direct imaging missions so are not simulated here. We have considered six different planetary size boundaries: 0.5 R⊕, 1.0 R⊕, 1.75 R⊕, 3.5 R⊕, 6.0 R⊕, and 14.3 R⊕ in our grid. These boundaries represent, respectively: the radius (0.5 R⊕) at which planets in the habitable zone appear to not have a sufficient gravity well to retain atmospheres (Zahnle & Catling 2017); the "super-Earths" (1− 1.75 R⊕) and "sub- Neptunes" (1.75 − 3.5 R⊕), as defined by Fulton et al. (2017) (see section 4.4) based on the observed gap in the radius distribution of small planets with orbital periods shorter than 100 days; the assumed upper limit on Neptune-size planets (6R⊕) based on the small peak in the radius distribution from Fulton et al. (2017); and the radius past which planets transition to brown dwarf stars (Chen & Kipping 2017). We have computed the corresponding pressure- temperature atmospheric profiles using (1) the non-grey analytical model of Parmentier and Guillot (2014) with the coefficients from Parmentier et al. (2015) and the Rosseland opacity functional form of Valencia et al. (2013), and (2) the grey analytical model of Robinson and Catling (2012, 2014), both modified to take the planetary size and instellation as unique input parameters. We have used the Robinson and Catling (2012, 2014) model for planets with radius smaller than 3.5 R⊕ at low instellations, and the Parmentier and Guillot (2014) model for planets with a radius of 14.3 R⊕ at high instellations. We have assumed an internal temperature Tint = 0 and mass-radius relations taken from Weiss and Marcy (2014) for planets with radius smaller than 3.5 R⊕. We have assumed an internal temperature Tint = 100 and the density calculated from Mass-radius relation from Chen & Kipping (2017) for planets with radius equal to 14.3 R⊕. Assuming a planetary mass – 7 – of 0.414MJ (131M⊕), the density is ∼ 0.25 g.cm−3. We have considered L = 0.95L(cid:12) and M = 0.965M(cid:12) for the parent star5. As an illustrative example, we show in Fig. 1 the condensation curves for the four different species we considered (solid black), along with temperature profiles for two different size planets: 0.5R⊕ and 14.3R⊕, at different incident stellar fluxes ('instellation'). The figure shows that ZnS would condense out at ∼ 10mb in the atmosphere of a highly irradiated (220 times Earth flux) 14.3 Earth radius planet, a typical hot-Jupiter, whereas CH4 would condense out in the atmosphere of a 0.5 Earth radius planet receiving only ∼ 1/280th the flux Earth is receiving. Following this procedure, we have derived the radius and stellar flux where other gaseous species condense in the atmosphere. Table 1 provides the corresponding data of the planetary radius and stellar flux boundaries that can be used for classifying planets into different regimes. These boundaries are parameterized in Eq.(1) and are also available as an online calculator at: http://www3.geosc.psu.edu/~ruk15/extrasolar/. Table 1: Planetary radii and stellar flux values at which the given species condense in a planetary atmosphere, for a star with L = 0.95L(cid:12) and M = 0.965M(cid:12). These limits form the boundaries for classifying planets into different categories to calculate exoplanet yield estimates. See Fig. 2 and Section 2.2. Stellar Flux (Earth flux) Radius (R⊕) ZnS 182 187 188 220 220 220 0.5 1.0 1.75 3.5 6.0 14.3 H2O 1.0 1.12 1.15 1.65 1.65 1.7 CO2 0.28 0.30 0.32 0.45 0.40 0.45 CH4 0.0035 0.0030 0.0030 0.0030 0.0025 0.0025 5The values of the stellar luminosity and mass are obtained as follows: We downloaded the confirmed and candidate Kepler catalog from NEXSCI, found the median values of stellar luminosity and mass for each data set, and then took the average value of luminosity and mass from these median values. Median Luminosity for candidate planet list: 1.057; Median Stellar mass for candidate planet list: 0.97; Median Luminosity for confirmed planet list: 0.86; Median Stellar mass for confirmed planet list: 0.96; Average luminosity of confirmed & candidate: (1.057 + 0.86)/2 = 0.95; Average stellar mass of confirmed & candidate: (0.97+0.96)/2 = 0.965 – 8 – F (Rp)i = aix5 + bix4 + cix3 + dix2 + eix + fi (1) where F (Rp)i is the stellar flux, normalized to the current Earth flux (1360 Wm−2), at which species i = (ZnS, H2O, CO2, CH4) condenses on a planet with radius Rp, and x = Rp/R⊕ given in Table 1. The coefficients for the four condensing species are given in Table 2. Table 2: Coefficients to be used in Eq.(1). a b c d e f ZnS 0.0010338041 H2O 0.0017802416 0.0002947546 CO2 CH4 -0.0000033096 - 0.0255230451 - 0.0443704003 - 0.0070583509 0.0000889715 0.1858822989 0.3302966853 0.0483928147 -0.0007644190 - 0.4990171468 - 0.9299682851 - 0.1198359100 0.0026719183 0.5690844110 1.1366785108 0.1477297602 -0.0038305535 1.6385396777 0.6255832476 0.2304769313 0.0048373922 2.1. Application of the Classification Scheme to Obtain Planet Occurrence Rates Extending the insights obtained from Fig. 1 and Table 1, it is then possible to define various zones as a function of stellar flux and planetary radius. This should qualitatively ap- ply across all stellar types and the entire field of planets, even if the quantitative positions of the boundaries change due to, for example, the age of the system and the amount of internal heat released from planets, or the stellar energy distribution (SED) and its relationship to planetary albedo (Segura et al. 2003). For a full consideration of such caveats, see section 3.1. This particular framing of the parameter space will allow us to calculate the occurrence of different kinds of planets within each zone based on their condensation conditions, as both radius and flux are measurable quantities. We have performed such a calculation of ηplanet, the fraction of stars that have a planet within one of the zones defined by a condensing species. As an illustrative example of how these boundaries can be used to calculate the occurrence of planets, we apply these criteria to the preliminary parametric model introduced by one of NASA's Exoplanet Program Analysis Group (ExoPAG) science analysis studies (SAG13). A detailed discussion of the SAG13 model is outside the scope of this paper, but we will summarize the most critical points. The SAG13 model is based on a simple meta-analysis of planet occurrence rates from – 9 – Fig. 1.- Dependence of ZnS, H2O, CO2 and CH4 condensation with pressure and tempera- ture in any planetary atmosphere (solid black) along with the pressure-temperature profiles for two different sizes of planets, 0.5 R⊕ and 14.3 R⊕ and two different instellations, 0.004 I⊕ and 220 I⊕, respectively (solid red). The intersections of the two sets of curves indicate that CH4 and ZnS are condensing out in each of considered planetary atmospheres. CH4,cond → CH4CO2,cond → CO2ZnS → ZnRobinson and Catling, 2012Robinson and Catling, 2014Parmentier and Guillot, 2014Parmentier et al., 2015r = 0.5 rEarth / I = 0.0035r = 14.3 rEarth / I = 220H2Ocond → H2O – 10 – many different individual publications and groups. Specifically, the SAG13 group collected tables of occurrence rates calculated over a standard grid of planet radius, period, and stellar type. A full description of the grid is as follows. The ith bin in the planet radius is defined as the interval: Ri = [1.5i−2, 1.5i−1)R⊕ This implies the following bin edges: [0.67, 1.0, 1.5, 2.3, 3.4, 5.1, 7.6, 11, 17,...] R⊕. The jth bin in the planet period is defined as: Pj = 10 . [2j−1, 2j)days (2) (3) This implies the following bin edges: [10, 20, 40, 80, 160, 320, 640,...] days Data and models from peer-reviewed publications (Petigura et al. 2013, ForemanMackey et al. 2014, Burke et al. 2015, Traub 2015, Dressing & Charbonneau 2013) were integrated over the standard grid, and supplemented by several unpublished tables from the 2015 Kepler "hack week" which were based on Q1-Q17 DR24 catalog, Kepler completeness curves, and data products at the time. However, for our current work, we did not use the SAG13 standard grid mentioned above because the SAG13 grid does not represent the condensation sequences described in the previous section. Instead, we took the stellar fluxes from Table 1 where species condensation happens, and calculated the corresponding orbital periods based on the stellar mass (0.965) M(cid:12) and luminosity (0.95) L(cid:12) described in footnote 5. It should be noted that the SAG-13 grids are available for different spectral types. Herein, our work focuses on the G dwarf population and employs the corresponding grids. The SAG13 submissions were then processed as follows. First, within each spectral type, the sample geometric mean (µi,j) and variance (σ2 i,j) was computed in each (i, j)-th bin of the period-radius grid, across the different submissions. The mean values µi,j formed a "baseline" table of occurrence rates. "Optimistic" and "pessimistic" tables were also defined by using the µi,j ± 1σi,j values for each (i, j)-th bin. SAG13 then fit a piecewise power law to the "pessimistic", "baseline", and "optimistic" combined tables. The power law had the following form: ∂2N (R, P ) ∂lnR ∂lnP = ΓiRαiP βi (4) – 11 – For optimistic case, Γi = [1.06, 0.78], αi = [−0.68,−0.82], βi = [0.32, 0.67]; for pes- simistic case Γi = [0.138, 0.72], αi = [0.277,−1.56], βi = [0.204, 0.51]; for baseline, Γi = [0.38, 0.73], αi = [−0.19,−1.18], βi = [0.26, 0.59] The break between two pieces of the power law was set at 3.4 R⊕ (following Burke et al. 2015), hence the two values for the coefficients, and a least squares fit was performed separately to each of the pieces. Similarly to the mean and variance above, logarithms of occurrence rates were used when performing the least squares of log occurrence rates, rather than actual occurrence rates, in order to properly balance the effects of small and large occurrence rates. This resulted in "pessimistic", "baseline", and "optimistic" parametric models. These models were then integrated across the planet parameter boundaries described in this paper. It should be stressed that community-sourced data do not represent independent mea- surements or estimates of scientific quantities, so that the SAG13 sample mean and variances should not be interpreted as a formal mean and uncertainty of exoplanet occurrence rates. Rather, they simply represent one possible way to measure the state of knowledge as well as the disagreement on the rates within the occurrence rate community. In other words, the SAG13 "pessimistic", "baseline", and "optimistic" cases refer to the typical pessimistic, av- erage, and optimistic submissions within the SAG13 community survey, rather than formal scientific results. Alternative ways of combining SAG13 results are also possible, such as: including only peer-reviewed submissions, including submissions based only on different catalogs, removing outliers, etc. A detailed analysis of this is beyond the scope of this paper, but as a general rule, combinations tend to fall somewhere between the occurrence rates published in Petigura et al.(2013) and Burke et al. (2015) for G-dwarfs, which represent a range of about 4 in the warm rocky planet regime, with a tendency to be closer to Burke et al. (2015). For example, the geometric mean combination which we use in this paper is about 25% lower than Burke et al. (2015) for warm rocky planets, though it is significantly higher than Petigura et al. (2013). It should also be stressed that extrapolation is implied when integrating fitted power laws into cold planets or very small planet sizes, so the numbers in those regions remain very unreliable. To judge the robustness of the SAG13 occurrence rate estimates, independent occurrence rates were calculated using the inverse detection efficiency method based on the data from the DR25 catalog. The occurrence rate per bin, η, is given by η = 1 n(cid:63) Σnp i 1 compi (5) – 12 – Where compi is the survey completeness evaluated at the radius and orbital period of each planet in the bin, n(cid:63) is the number of stars surveyed, np is the number of planets in each bin. The planet list is taken from Thompson et al. (2017), using a disposition score cut of 0.9. The stellar properties are taken from Mathur et al. (2017), removing giant stars with logg < 4.2 and culling the G dwarfs by selecting stars with 5300 K < Tef f < 6000 K. The completeness of each star is calculated with KeplerPORT (Burke & Catanzarite 2017), and the survey completeness is calculated by averaging over all stars that were successfully searched for planets according to the timeoutsumry flag. The above equation assumes that the vetting completeness (the fraction of planet tran- sit signals (TCEs, Threshold Crossing Events) properly classified as planet candidates) and reliability (the fraction of transiting candidates that are not caused by instrumental arti- facts or statistical false alarms) are 100%. The vetting completeness and reliability are very important for small planets, especially at long orbital periods. The vetting completeness decreases substantially when one employs a high score cut on the DR25 catalog while the reliability approaches 100%. The net effect is that occurrence rates are likely to be under- estimated by ignoring both corrections. It should be noted that all SAG13 calculations also ignored vetting completeness and reliability. We note that more work needs to be done to do a reliable occurrence rate calculation. For regions with no planet detections (low instellation, long period orbits), occurrence rates were estimated with a parametric function that is a broken power-law in period and ra- dius. Free parameters were constrained using the Exoplanet Population Observation Simula- tor (EPOS, Mulders et al. in prep). EPOS generates planet populations from this parametrized description using a Monte Carlo simulation, and conducts synthetic observations using the survey completeness from the DR25 catalog. The synthetic observable populations are com- pared with the observed planet distribution from Kepler in the range P = [2, 400] days and Rp = [0.5, 8] R⊕ , and the posterior parameters are estimated using emcee (Foreman-Mackey et al. 2013). Binned occurrence rates are calculated by marginalizing the posterior paramet- ric distribution over the bin area, and taking the 50% and 16% and 84% percentiles for the mean and 1-sigma error, respectively. Table 4 provides the occurrence rates calculated from Eq.(5) for the same bins as in Tables 1 & 2. The values are more or less consistent within the uncertainties of SAG13 ηbasl from Table 2. However, the extrapolations into the cold planet regimes (low instel- lation fluxes) results in a disagreement between SAG13 values and from Eq. (5). This is expected, considering that (1) the cold regimes do not have any planet detections and any extrapolations are expected to wildly deviate (even between the methodologies) from the true distribution, and (2) the SAG13 rates are a combination of several individual methodoligies. – 13 – Table 3: Occurrence rates of planets in different boundaries, defined in Ta- ble 1 classification scheme. The ηplanet values are estimated using SAG-13 occurrence rates. The (cid:63) values are based on extrapolation and therefore are very uncertain. Planet Type (Stellar Flux range) Radius (R⊕) Hot rocky (182-1.0) Warm rocky (1.0-0.28) Cold rocky (0.28-0.0035) Hot super-Earths (187-1.12) Warm super-Earths (1.12-0.30) Cold super-Earths (0.30-0.0030) Hot sub-Neptunes (188-1.15) Warm sub-Neptunes (1.15-0.32) Cold sub-Neptunes (0.32-0.0.0030) Hot sub-Jovians (220-1.65) Warm sub-Jovians (1.65-0.45) Cold sub-Jovians (0.45-0.0030) Hot Jovians (220-1.65) Warm Jovians (1.65-0.40) Cold Jovians (0.40-0.0025) 0.5-1.0 0.5-1.0 0.5-1.0 1.0-1.75 1.0-1.75 1.0-1.75 1.75-3.5 1.75-3.5 1.75-3.5 3.5-6.0 3.5-6.0 3.5-6.0 6.0-14.3 6.0-14.3 6.0-14.3 ηpess 0.22 0.09(cid:63) 0.50(cid:63) 0.21 0.087 0.50(cid:63) 0.29 0.12 0.77(cid:63) 0.05 0.04 0.58(cid:63) 0.028 0.023 0.34(cid:63) ηbasl 0.67 0.30(cid:63) 1.92(cid:63) 0.47 0.21 1.42(cid:63) 0.48 0.22 1.63(cid:63) 0.07 0.07 1.35(cid:63) 0.056 0.053 1.01(cid:63) ηopt 2.04 1.04(cid:63) 7.61(cid:63) 1.04 0.54 4.14(cid:63) 0.79 0.41 3.52 (cid:63) 0.12 0.13 3.19(cid:63) 0.11 0.12 3.07(cid:63) – 14 – Table 4: Occurrence rates calculated from Eq.(5). Comparing with the ηbasl values from Table 3 from SAG-13 occurrence rates, the SAG13 values are more or less consistent with the values given below within the uncertainties. We use the ηbasl values from Table 3 to calculate the exoplanet yield estimates in section 2.2. As with Table 3, the (cid:63) values are extrapolated. Planet Type (Stellar Flux range) Radius (R⊕) Hot rocky (182-1.0) 0.5-1.0 η 0.552+0.195 −0.150 Warm rocky (1.0-0.28) 0.5-1.0 0.215+0.148 −0.099 (cid:63) Cold rocky (0.28-0.0035) 0.5-1.0 1.09+1.48−0.755 (cid:63) Hot super-Earths (187-1.12) 1.0-1.75 Warm super-Earths (1.12-0.30) 1.0-1.75 Cold super-Earths (0.30-0.0030) 1.0-1.75 Hot sub-Neptunes (188-1.15) 1.75-3.5 Warm sub-Neptunes (1.15-0.32) 1.75-3.5 Cold sub-Neptunes (0.32-0.0.0030) 1.75-3.5 Hot sub-Jovians (220-1.65) 3.5-6.0 Warm sub-Jovians (1.65-0.45) 3.5-6.0 Cold sub-Jovians (0.45-0.0030) 3.5-6.0 0.374+0.068 −0.056 0.145+0.071 −0.061 0.78+0.86−0.52 (cid:63) 0.356+0.049 −0.047 0.147+0.058 −0.057 0.85+0.88−0.57 (cid:63) 0.113+0.019 −0.018 0.051+0.021 −0.020 0.279+0.31−0.18 (cid:63) Hot Jovians (220-1.65) 6.0-14.3 0.004+0.011 −0.004 Warm Jovians (1.65-0.40) 6.0-14.3 Cold Jovians (0.40-0.0025) 6.0-14.3 0.002+0.004 −0.001 0.008+0.031 −0.007 (cid:63) – 15 – Integrating the SAG13 parametric model across the bin boundaries defined in Table 1 gives the occurrence rates in Fig 2, where we have assumed L = 0.95L(cid:12) and M = 0.965M(cid:12). Each class of planet has an occurrence that is a mixture of astrophysical effect and an observational bias. For example, even though it is easier to detect giant planets in close orbits, their occurrence rate is comparatively smaller (0.05) than close-in sub-Neptune or terrestrial-size planets (∼ 0.48, respectively). The implication is that hot, giant planets are likely less in number. The trend in Fig. 2 indicates that the occurrence rate generally increases, from larger planets to the smaller ones in any particular bin. We can define the inner edge of the habitable zone (HZ) as the boundary where H2O starts condensing in a terrestrial planets atmosphere, and the outer edge of the habitable zone as CO2 condensation boundary (Abe et al. 2011). Within this zone, it appears that the terrestrial size planets have higher occurrence rates (0.2 − 0.3) than compared to either Jovians (0.053) or Neptunes (0.07) planets. However, it should be noted that the occurrence rate of terrestrial planets in this regime is severely restricted by low number statistics. 2.2. Mission yield estimates With the planet categorization scheme and associated occurrence rates described above, we estimated the exoplanet yields for each type of planet using the yield optimization code of Stark et al. (2015). Briefly, this code works by simulating the detection of extrasolar planets around nearby stars over the lifetime of a mission. To do so, it distributes a large number of synthetic planets around each nearby star, sampling all possible orbits and phases consistent with the planet definition, illuminates them with starlight, calculates an exposure time for each planet given a set of assumptions about the instrument and telescope, and determines the fraction of planets that are detectable within a given exposure time (i.e., the "completeness"). The code optimizes the exposure time of each observation, as well as which stars are observed, the number of observations to each star, and the delay time between observations, to maximize the yield for a given type of planet. We first ran the yield code to define the set of observations that maximized the yield of exoEarth candidates. We adopted the same baseline mission parameters defined in Table 3 of Stark et al. (2015) with exception to the OWA, which increased from 15 λ/D to 30 λ/D, the contrast for spectral characterization which improved from 5 × 10−10 to 1 × 10−10, the spectral resolution which increased from R=50 to R=70, and the SNR required for spectral characterization which increased from 5 to 10 per spectral channel. We also adopted a new definition for exoEarth candidates. We distributed exoEarth candidates across the – 16 – Kopparapu et al. (2014) conservative HZ, which ranges from 0.95 - 1.67 AU for a solar twin. The semi-major axis distribution followed the analytic SAG13 occurrence rate fits. ExoEarth candidates ranged from 0.5 1.4 Earth radii, with the lower radius limit set by 0.8 ∗ (a ∗ (L(cid:63))1/2)−0.5, a reasonable limit following the work of Zahnle & Catling (2017). In this paper, the All exoEarth candidates were assigned a flat geometric albedo of 0.2. exoEarths are used solely to optimize an observation plan. We do not report on the yield of these exoEarth candidates, focusing instead on the yields of the classes of planets defined in section 2 when following such an observation plan. We then locked this set of observations in place and re-ran the yield code to calculate the yield of each planet type discussed above, simply by changing the planets input parameters each time. For each planet type, we distributed planet radii and orbital period according to the SAG13 distribution. We assumed Lambertian phase functions for all planets. To calculate the brightness of a given planet, we must also know the planets albedo. The actual distribution of exoplanet albedos is unknown. So for this study, we simply assigned each planet type a single reasonable albedo. We adopted a wavelength-independent geometric albedo of 0.2 for rocky planets and 0.5 for all other planets. To calculate an expected yield, we must also assign each planet type an occurrence rate. Table 3 lists the occurrence rates obtained from §2.1 in each bin of planet size and planet type (hot/warm/cold). The histogram plot in Fig.3 visualizes the total scientific impact of the habitable planet candidate survey. The y-axis gives the expected total numbers of exoplanets observed (yields), which are also given by the numbers above the bars. By "expected", we mean the most probable yield after many trials of an identically executed survey. Three sizes of exoplanets are shown, consistent with Table 1. For each planet size, three incident stellar flux classes are shown: hot (red), warm (dark blue), and cold (ice blue). The boundaries between the classes correspond to the temperatures where metals, water vapor, and carbon dioxide condense in a planet's atmosphere. The warm bin is not the same as the habitable planet candidate bin, as it is likely too generous. We also calculated each planet's yield when deviating from the baseline mission param- eters. Figs. 4 & 5 show the sensitivity of each planet's yield to changes in a single mission parameter. Each yield curve has been normalized to unity at the value of the baseline mis- sion. As expected, the yield of hot planets is more sensitive to the IWA than cold planets, and the yield of cold planets is more sensitive to OWA than hot planets. However, surpris- ingly, the yield of cold Jupiters is quite sensitive to IWA, suggesting that an observation plan optimized for the detection of exoEarths will typically detect cold Jupiters in gibbous phase near the IWA. We note that in general, larger apertures are less sensitive to changes in mission parameters than smaller apertures. – 17 – Fig. 2.- Planet occurrence rate estimates from SAG13 baseline analysis (see Table 3) as a function of incident flux and planetary radius, assuming a star with L = 0.95L(cid:12) and M = 0.965M(cid:12). The boundaries of the boxes represent the regions where different chemical species are condensing in the atmosphere of that particular size planet at that stellar flux, according to equilibrium chemistry calculations. The radius division is from Fulton et al. (2017) for super-Earths and sub-Neptunes, and from Chen & Kipping (2017) for the upper limit on Jovians. The '(cid:63)' values are based on extrapolation and therefore are very uncertain. See §2.1 for more details. 0.00350.281182Stellar Flux (Earth units)0.511.753.5614.3Planet Radius (RE)Hot JoviansHot Neptunes/sub-Jovians0.0560.070.480.47Hot Sub-NeptunesNeptunes0.07SubNeptunes0.220.21Hot Super-EarthsHot EarthsWarmEarths0.670.30SuperEarthsWarmJovians0.053ColdJoviansCold NeptunesCold sub-NeptunesCold Super-EarthsCold Earths1.01*1.35*1.63*1.42*7.61*ZnSH2OCO2CH4 – 18 – (a) (b) (c) Fig. 3.- Expected number of exoplanets observed (y-axis) for the baseline occurrence rates in each planet category (rocky, super-Earths, sub-Neptunes, sub-Jovians and Jovians) for hot (red), warm (blue) and cold (ice-blue) incident stellar fluxes shown in Table 1 and Fig.2. The telescope sizes are (a) 4m, (b) 8m and (c) 16m. The occurrence rates, as well as yield estimates, ignore multiplicity and the planet categories were all treated as effectively independent. – 19 – 3. Discussion In this section, we discuss caveats to our classification scheme, and its relevance to future missions that plan to detect and/or characterize extrasolar planets. 3.1. Caveats to the classification scheme The boundaries discussed in earlier sections are made out of the necessity for creating a single classification scheme that applies to all planets, and that can translate current planet obervations into predictions of future planet yields. No single scheme will be able to properly capture the complex interactions between myriad planetary processes in exoplanet atmospheres. An analogy with the habitable zone here is useful: The habitable zone is not a good means of determining the habitability of a single planet. Instead, it is best used to understand how many potentially habitable worlds a given mission may be able to probe for evidence of habitability. Similarly, the classification scheme discussed in this paper is probably not the best means for determining whether a single world has a certain combination of clouds decks or upper atmospheric composition. Ultimately, that should be determined by specific observations of such worlds. What this scheme is useful for is to help understand the diversity, and the number, of worlds that can be expected from future missions and ground-based observatories. Even given the above overall caveat, more specific caveats exist. Much of the work upon which this classification scheme is based is relatively new, and represents an area of rapidly evolving thought. This means that the resulting classification scheme will also have to evolve as these theories are better elucidated, tested, and refined. Ultimately, when future exoplanet characterization missions and ground-based observatories become operational, they will pro- vide the tests of the various hypotheses contained in the drawing of the boundaries shown in Fig. 2. This again draws similarities to the habitable zone, which has evolved over the years, and will ultimately be determined by missions. But until such observations are made, this scheme (and the HZ) will enable predictions of how many such worlds we will be able to eventually epxlore in detail. An example of this rapid evolution is the work by myriad groups on the mass-radius relationship of planets (Weiss & Marcy 2014; Rogers 2015; Wolfgang et al. 2016; Chen & Kipping 2017) Clearly, this area of research is rapidly evolving - and further data on this relationship is anticipated with continues ground-based observations of planets around M dwarfs combines with TESS observations of the same target. Thus, we anticipate changes to the boundaries selected in this study as these new data and models are incorporated into – 20 – this line of research. We also anticipate that these data will allow for the study of how these dividing lines are a function of instellation received by the planet. If that is the case, the "horizontal" lines in our classification scheme (Figure 2) would instead be diagonal. We used horizontal lines here as a first-order determination of this scheme and because deriving the slope of those lines empirically or theoretically is well beyond the scope of this work. This is something that may be true more generally - the lines drawn in our scheme may all be diagonal, as we think instellation could impact planet size categories and that planet size could impact the instellations at which various cloud decks form. A second category of caveat is the lack of consideration of many planetary processes and planet/star properties in this study. Kinetic chemistry, atmospheric circulation, internal heat generation, different star types, and differing bulk planet composition are not explicitly considered in this study. Each of these processes could affect the boundaries considered here. Our boundaries have been determined at equilibrium. However, the distribution of chemical constituents in planetary atmospheres can be strongly affected by the so-called chemical quenching, in which as the material moves, the temperature and pressure within the gas mixture change and chemical reactions may become slower, potentially reaching a certain state for which the chemical abundances are 'frozen' (Madhusudhan et al. 2016). Also, photochemistry could also impact the ability of a condensate to form. This is most evident in the potential for secondary aerosols that form when photochemical byproducts cause supersaturation and condensation in an atmosphere. This process occurs on many solar system worlds (Courtin 2005; Waite et al. 2007; Gladstone et al. 2016), is thought to have occurred on Archean Earth (Arney et al. 2016), and likely occurs on exoplanets (Moses et al. 2011, 2014). However, these processes themselves will likely represent a signif- icant overprint over the transitions that are proposed here, where within one of the planet categories we propose, planets that are closer to their host star will have a gradual and increasingly important contribution from non-equilibrium chemistry. But others may also exist, including the potential for other photochemical byproducts to form in an atmosphere and be evidenced in the planetary spectrum. However, we expect these will be secondary in importance compared to the optical depth and cold-trapping effects of discrete cloud layers. The specific properties of the planet and star are also not considered here. Atmospheric circulation, internal heat generation, and the stellar energy distriburion (SED) are also ignored here. These could all impact the specific position of the boundaries we propose. Changing the SED can lead to changes in the planetary albedo, which would affect the position of these boundaries just as they affect the habitable zone (Segura et al. 2003; Shields et al. 2013, 2014; Yang et al. 2013, 2014a; Kopparapu et al. 2014, 2016; Wolf et al. 2017). Greater internal heat generation, or decreased convection could lead to a suppression of – 21 – cloud formation. These are not independent variables, as increased heat generation should lead to increased convection and the SED could also impact circulation as it impacts the altitude-dependent deposition of energy in the atmosphere. Although these properties and processes are not included here, they could be in future papers, just as the consideration of the habitable zone now includes consideration of internal heat sources (Barnes et al. 2013; Haqq-Misra & Kopparapu 2015), atmospheric circulation and convection. Finally, we do not consider the impact of the bulk chemical composition of the planet. Clearly, a planet that is extremely carbon poor will be less likely to condense CO2 or CH4. Lastly, we note that the simulations of the scheme in this paper considers a limited parameter space, due to its focus on a near-term problem faced by the exoplanet commu- nity. Given the ongoing studies of flagship missions that propose to directly image and spectroscopically characterize extrasolar planets - HabEx and LUVOIR - we require an abil- ity to discuss and compare the exoplanet yields from such missions. This will also help us understand how these missions will complement the discoveries that will be made by already-planned ground-based and space-based observatories. The focused utility shown in this manuscript comes at the expense of other applications of the proposed classification scheme. It could also be used to simulate the yields from future transit spectroscopy mis- sions such as JWST or CHEOPS. It also helps us understand how detection missions such as TESS, PLATO, and WFIRST will provide complementary discoveries to past exoplanet detections, and increase the total expected diversity of known planets. These are all worthy applications of the scheme we propose, but we save these applications for future manuscripts. 3.2. Application of classification scheme to future space-based direct imaging missions These estimates of the abundances of different planet types allow for projections of the yields of future exoplanet missions. As mentioned in the introduction, two such missions are under study in advance of the next astrophysics Decadal Survey: HabEx and LUVOIR. Most of the discussion of these missions - and past direct imaging studies - has focused on their ability to find and characterize rocky planets in the habitable zones of nearby stars. However, the observations enabled by such missions would also bring an ability to observe other kinds of worlds. All planets between the inner working angle and outer working angle of the starlight suppression at the time of observation can be observed. And many of these worlds will be brighter than rocky planets in the habitable zone. Therefore, even observation strategies that are optimized for maximizing the yield of rocky planets in the habitable zone will also yield observations of a considerable diversity of worlds. – 22 – None of these yield estimates should be taken as simulations of the yields of the ar- chitectures the HabEx and LUVOIR teams are studying. For one, these simulations hold everything except telescope diameter constant, as a way to demonstrate how planet diversity scales with telescope size and as a way to show end-member yields across the range of mis- sion sizes being explored (For a more extensive options of varying quantities other than the telescope diameter, see here: http://jt-astro.science:5100/multiplanet_vis). More importantly, the two studies will explore four different architectures which may have differ- ent instrument properties, assume different levels of technological development, and different starlight suppression techniques. None of those nuances are considered here. We refer the reader to the eventual reports from these teams for better estimates of the yields from these missions. What we present here should instead be considered an example of how these planet categories can be useful to such yield simulations. Our simulations predict that a 4m-class mission would observe a great diversity of worlds (Figure 3(a)). At that scale, a mission whose observations are designed to maximize the yield of potential Earths will also yield the detection and characterization of all of the planet types discussed here, with the exception of hot Jupiters. Hot Jupiters are not observed by a 4m- class mission because the tight inner working angle of the 4m mission, and because of the low abundances of hot Jupiters (see Table 2) . But a total of up to 90 planets are observed, including up to 12 rocky/super-Earths planets in the habitable zone. The diversity of these simulations will allow tests - albeit on a small sample size within each bin - of the planet bins proposed here. An 8m-class mission yields even more planets (Figure 3(b)). We predict that class of mission would observe over 300 planets, including at least a few planets in each of the planet classes proposed in this manuscript. Although neither LUVOIR nor HabEx is explicitly considering an 8m mission, this size sits at the dividing line between the two studies and represents an approximation of what a large HabEx or a small LUVOIR could enable. The larger versions of LUVOIR (Fig. 3(c)) bring the ability to not only observe planets, but to test the occurrence of different features within each of the planet bins. It would observe dozens of each planet type, providing larger sample sizes which enables to study each planet type as a population. As shown by Stark et al. 2014, a sample of 30 planets of a given type would allow us to be sensitive to any feature that has at least a 10% chance of occurring. For this class of mission, there would be a total of up to 1000 worlds. This includes several hundred Neptune-size planets with orbits that likely cause water clouds - but not carbon dioxide clouds - to form. The least populated bins are for hot Jupiters, which only contains ∼ 15 detections, and for "cold rocky" which would have CO2 clouds, which contains only few detections. But all other bins have at least 50 detections. As mentioned – 23 – above, 30 planets of a given type would allow us to be sensitive to any feature that has a ∼ 10% chance of occurrence within that planet type (Stark et al. 2014). It is this large number of observation that would allow the myriad hypotheses contained in this manuscript to be tested. For example, the presence/absence of various cloud types could be plotted as a function of the energy incident at the top of exoplanet atmospheres. And the absence of absorption features associated with cold-trapping could be measured in each of these bins. This logic has been applied to habitable planets before (Stark et al. 2014); here we demonstrate that it also applies to planets beyond the habitable zone. 4. Conclusions NASA concept mission studies that are currently underway, like LUVOIR and HabEX, are expected to discover a large diversity of exoplanets. These larger missions would provide large enough sample sizes that we could study each planet type in the context of its "rela- tives" in the Solar system or currently known exoplanets. We present here a classification scheme for exoplanets in planetary radius and stellar flux bins, based on chemical species' condensation sequences in planetary atmospheres. This chemical behavior of gases is depen- dent on pressure, temperature and metllicity, and we primarily focus on condensation curves of sphalerite ZnS, H2O, CO2 and CH4. The order of condensation of these species repre- sent the order in which the boundaries of our classification scheme are defined. We then calculated the occurrence rates of different classes of exoplanets within these boundaries, and estimated the direct imaging mission yields for various telescope sizes. While the main focus of future flagship exoplanet direct imaging missions is to characterize a habitable world for bio-signatures, the missions will also have the ability to observe other kinds of planets within the system. Therefore, distinguishing features that separate planet categories based on current observables (planet radius and incident stellar flux), and a scheme to illustrate these categories, is essential in calculating the expected direct imaging mission yields, and correspondingly choosing an optimal observational strategy. The authors would like to thank an anonymous reviewer for their constructive comments. We also thank Aki Roberge, Mark Marley and Eric Lopez for providing comments on the work and the manuscript. The authors would like to gratefully acknowldege the following colleagues who have contributed to SAG-13 community occurrence rate project: Chris Burke, Joe Catanzarite, Courtney Dressing, Will Farr, B. J. Fulton, Daniel Foreman-Mackey, Danley Hsu, Erik Petigura, Wes Traub. We would also like to thank the NASA's Kepler mission for organizing the "hack week" – 24 – (a) (b) Fig. 4.- Yield for each planet type when deviating from the baseline mission by varying one parameter at a time. The top panel is for a 4m mirror size and bottom is for 8 meter. Solid, dashed, and dotted lines correspond to rocky, sub-Neptune, and Jovian planet types. Color scheme is the same as Figure 3. – 25 – from October 13 - 15 2015, that provided a forum for project scientists and community researchers to interact, and led to significant input into this manuscript. R. K and S. D. G. gratefully acknowledge funding from NASA Astrobiology Institute's Virtual Planetary Laboratory lead team, supported by NASA under cooperative agreement NNH05ZDA001C. E.H. was supported by an appointment to the NASA Postdoctoral Program at NASA God- dard Space Flight Center, administered by Universities Space Research Association through a contract with NASA. REFERENCES Abe, Y., Abe-Ouchi, A., Sleep, N. H. et al. 2011. Astrobiology, 11, 443 Anglada-Escud´e, G., Amado, P. J., Barnes, J. et al. 2016, Nature, 536, 7617, 437 Arney, G., Domagal-Goldman, S. D., Meadows, V. S. et al. 2016. Astrobiology, 16, 11 Barnes, R., Mullins, K., Goldblatt, C. et al. 2013. Astrobiology, 13, 225 Bonfils, X., Delfosse, X., Udry, S., et al. 2013. A&A, 549, A109 Burke, C. J., et al. 2015. ApJ, 809, 8 Burke, C. J., & Catanzarite, J. Kepler Science document, KSCI-19111-002, Edited by Michael R. Haas and Natalie M. Batalha Burrows, A., & Sharp, C. M. 1999, ApJ, 512, 2 Burrows, A., Sudarsky, D., & Hubeny, I. 2004. ApJ, 609, 1 Burrows, A. 2005. Nature, 433, 261 Cahoy, K. L., Marley, M. S., & Fortney, J. J. 2010, ApJ, 724, 1 Cassan, A. et al. 2012. Nature, 481, 7380 Catanzarite, J., & Shao, M. 2011. ApJ, 738, 151 Charnay, B., et al. 2015, ApJ, 813, L1 Chen, J., & Kipping, D. 2016. ApJ, 834, 17 Coughlin, J. L. et al. 2016. ApJS, 224, 12 Courtin, R. 2005. Space Science Reviews, 116, 1-2 – 26 – Dittman, J. A., Irwin, J. M., Charbonneau, D. et al. 2017. Nature, 544, 333 Dong, S., & Zhu, Z. 2013. ApJ, 778, 53 Dressing, C., & Charbonneau, D. 2013. ApJ, 767, 95 Dressing, C., & Charbonneau, D. 2015. ApJ, 807, 1 Evans, D., & Hubbard, W. 1972. Nature, 240, 162 Foreman-Mackey, D., Hogg, D. W., Morton, T. D. 2014. ApJ, 795, 1 Fortney, J. J. 2005. MNRAS, 364, 469 Fortney, J. J., Marley, M. S., Barnes, J. W. 2007. ApJ, 659, 2 Fressin, F., Torres, G., Charbonneau, D., et al. 2013. ApJ, 766, 81 Fulton, B., J., Petigura, E. A., Howard, A. W., et al. 2017, AJ, 154, 3 Gaidos, E. 2013, ApJ, 770, 90 Gillon, M., Triaud, A. H. M. J., Demory, B. O. et al. 2017. Nature, 542, 456 Gladstone, G. R., Stern, S. A., Ennico, K. et al. 2016. Science, 351, 6279 Godolt, M., Grenfell, J. L., Hamann-Reinus, A. et al. 2015. Planetary and Space Science, 111, 62 Haqq-Misra, J., & Kopparapu, R. K. 2015. MNRAS, 446, 428 Haqq-Misra, J., Kopparapu, R. K., Batalha, N. 2016. ApJ, 827, 2 Howard, A. W. et al. 2012, ApJS, 201, 15 Kane, S. R., Kopparapu, R. K., Domagal-Goldman, S. 2014. ApJ, 794, L5 Kane, S. R., Hill, M. L., Kasting, J. F. et al. 2016. ApJ, 830, 1 Kasting, J., F., Whitmire, D., P., & Reynolds. R. T. 1993, Icarus, 101, 108 Kopparapu, R. K., Ramirez, R., Kasting, J. F., Eymet, V., Robinson, T. D., Mahadevan, S., Terrien, R. C., Domagal-Goldman, S. D., Meadows, V., & Deshpande, R. 2013, ApJ, 765, 131 Kopparapu, R. K. 2013, ApJ, 767, L8 – 27 – Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J. et al. 2014. ApJ, 787, L29 Kopparapu, R. K., Wolf, E. T., Haqq-Misra, J. et al. 2016. ApJ, 819, 1 Kopparapu, R. K., Wolf, E. T., Arney, G. 2017, ApJ, 845, 5 Leconte, J., Forget, F., Charnay, B., Wordsworth, R., & Pottier, A. 2013a. Nature, 504, 268 Leconte, J., Wu, H., Menou, K., & Murray, N. 2015. Science, 347, 632 Lopez, E. D., & Fortney, J. J. 2014. ApJ, 792, 1 Madhusudhan, N., Ag´undez, M., Moses, J. I., Hu, Y. 2016. Space Science Reviews, 205, 1 Marley, M. S., Fortney, J., Seager, S., & Barman, T. 2007. Protostars and Planets V, 733 Mathur, S., Huber, D., Batalha, N. M., et al. 2017. ApJS, 229, 2 Morley, C. V., Fortney, J. J., Marley, M. S. et al. 2012. ApJ, 756, 2 Morley, C. V., Fortney, J. J., Kempton, E. M.-R. et al. 2013. ApJ, 775, 33 Morton, T. D., & Swift, J. 2014, 791, 10 Morton, T. D. et al. 2016. ApJ, 822, 86 Moses, J. I., Visscher, C., Fortney, J. J. et al. 2011. ApJ, 737, 15 Moses, J. I. 2014, Nature, 505, 7481 Mulders, G. D., Pascucci, I., Apai, D. 2015. ApJ, 798, 2 Parmentier, V., & Guillot, T. 2014. A&A, 562, A133 Parmentier, V. et al. 2015. A&A, 574, A35 Pierrehumbert, R. T., & Gaidos, E. 2011. ApJ, 734, L13 Petigura, E. A., Howard, A. W., Marcy, G. W. 2013, PNAS, 110, 48 Pollack, J. et al. 1996. Icarus, 124, 1 Ramirez, R. M., & Kaltenegger, L. 2017. ApJ, 837. L4 Robinson, T. D., & Catling, D. C. 2012. ApJ, 757, 104 Robinson, T. D., & Catling, D. C. 2014. Nature geoscience, 7, 1 – 28 – Rogers, L., & Seager, S. 2010a, ApJ, 712, 2 Rogers, L., & Seager, S. 2010b, ApJ, 716, 2 Rogers, L. 2015. ApJ, 801, 41 Rowe, J. F. et al. 2014. ApJ, 716, 45 Silburt, A., Gaidos, E., Wu, Y. 2015, ApJ, 799, 2 Segura, A. et al. 2003. Astrobiology, 3, 4 Selsis, F. et al. 2007b. A&A, 476, 137 Shields, A. L., Meadows, V. S., Bitz, C. M., et al. 2013. Astrobiology, 13, 8 Shields, A. L., Bitz, C. M., Meadows, V. S. et al. 2014. ApJ, 785, L9 Stark, C. C., Roberge, A., Mandell, A., Robinson, T. D. 2014. ApJ, 795, 122 Stark, C. C. et al.. 2015. ApJ, 808, 149 Sudarsky, D., Burrows, A., Hubeny, I. 2003. ApJ, 588, 1121 Suzuki, D., Bennett, D. P., Sumi, T. 2016. ApJ, 833, 145 Thompson, S. E., Coughlin, J. L., Kelsey, H. et al. 2017. submitted to ApJS, arXiv:1710.06758 Traub, W. A. 2012. ApJ, 745, 20 Waite, J. H., Young, D. T., Cravens, T. E. et al. 2007. Science, 316, 5826 Wakefor, H. R., & Singh, D. K. 2015. A&A, 573, A122 Wakeford, H. R., Visscher, C., Lewis, N. K. et al. 2017. MNRAS, 464, 4247 Way, M. J., Del Genio, A. D., Kelley, M., Aleinov, I., Clune, T. 2015. arXiv:1511.07283 Weiss, L., & Marcy, G. 2014. ApJ, 783, L6 Wolf, E., & Toon, E. 2014. Geophysical Research Letters. 41, doi:10.1002/2013GL058376 Wolf, E., & Toon, E. 2015. JGRA, 120, 5775 Wolf, E. T., Shields, A. L., Kopparapu, R. K., et al. 2017. ApJ, 837, 107 – 29 – Wolfgang, A., Rogers, L., Ford, E. B. 2016. ApJ, 825, 19 Wordsworth, R., Forget, F., & Eyment, V. 2010, Icarus, 210, 2, 992 Yang, J., Cowan, N. B., & Abbot, D. S. 2013. ApJ, 771, L45 Yang, J., Gwenael, B., Fabrycky, D., & Abbot, D. 2014. ApJ, 787, L2 Zahnle, K. J., & Catling, D. C. 2017. ApJ, 843, 122 Zsom, A., Seager, S., de Wit, J. et al. 2013, ApJ, 778, 2 This preprint was prepared with the AAS LATEX macros v5.2. – 30 – Fig. 5.- Same as Fig. 4, but for a 16m mirror size.
1608.00963
2
1608
2016-11-04T14:20:27
Precise radial velocities of giant stars IX. HD 59686 Ab: a massive circumstellar planet orbiting a giant star in a ~13.6 au eccentric binary system
[ "astro-ph.EP" ]
Context: For over 12 yr, we have carried out a precise radial velocity survey of a sample of 373 G and K giant stars using the Hamilton \'Echelle Spectrograph at Lick Observatory. There are, among others, a number of multiple planetary systems in our sample as well as several planetary candidates in stellar binaries. Aims: We aim at detecting and characterizing substellar+stellar companions to the giant star HD 59686 A (HR 2877, HIP 36616). Methods: We obtained high precision radial velocity (RV) measurements of the star HD 59686 A. By fitting a Keplerian model to the periodic changes in the RVs, we can assess the nature of companions in the system. In order to discriminate between RV variations due to non-radial pulsation or stellar spots we used infrared RVs taken with the CRIRES spectrograph at the Very Large Telescope. Additionally, to further characterize the system, we obtain high-resolution images with LMIRCam at the Large Binocular Telescope. Results: We report the likely discovery of a giant planet with a mass of $m_{p}~\sin i=6.92_{-0.24}^{+0.18}~M_{Jup}$ orbiting at $a_{p}=1.0860_{-0.0007}^{+0.0006}$ au from the giant star HD 59686 A. Besides the planetary signal, we discover an eccentric ($e_{B}=0.729_{-0.003}^{+0.004}$) binary companion with a mass of $m_{B}~\sin i=0.5296_{-0.0008}^{+0.0011}~M_{Sun}$ orbiting at a semi-major axis of just $a_{B}=13.56_{-0.14}^{+0.18}$ au. Conclusions: The existence of the planet HD 59686 Ab in a tight eccentric binary system severely challenges standard giant planet formation theories and requires substantial improvements to such theories in tight binaries. Otherwise, alternative planet formation scenarios such as second generation planets or dynamical interactions in an early phase of the system's lifetime should be seriously considered in order to better understand the origin of this enigmatic planet.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. HD59686_rev September 11, 2018 c(cid:13)ESO 2018 Precise radial velocities of giant stars IX. HD 59686 Ab: a massive circumstellar planet orbiting a giant star in a ∼13.6 au eccentric binary system(cid:63),(cid:63)(cid:63),(cid:63)(cid:63)(cid:63) Mauricio Ortiz1, Sabine Reffert1, Trifon Trifonov1, 2, Andreas Quirrenbach1, David S. Mitchell3, Grzegorz Nowak10, 11, Esther Buenzli4, Neil Zimmerman7, 8, Mickaël Bonnefoy9, Andy Skemer5, Denis Defrère5, Man Hoi Lee2, 12, Debra A. Fischer6, and Philip M. Hinz5 1 Landessternwarte, Zentrum für Astronomie der Universität Heidelberg, Königstuhl 12, 69117 Heidelberg, Germany e-mail: [email protected] 2 Department of Earth Sciences, The University of Hong Kong, Pokfulam Road, Hong Kong 3 Physics Department, California Polytechnic State University, San Luis Obispo, CA, 93407, USA 4 Institute for Astronomy, ETH Zurich, Wolfgang-Pauli-Strasse 27, 8093 Zurich, Switzerland 5 Steward Observatory, Department of Astronomy, University of Arizona, 933 N. Cherry Ave, Tucson, AZ 85721, USA 6 Department of Astronomy, Yale University, New Haven, CT, 06511, USA 7 Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidelberg, Germany 8 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 9 Univ. Grenoble Alpes, IPAG, F-38000 Grenoble, France. CNRS, IPAG, F-38000 Grenoble, France 10 Instituto de Astrofísica de Canarias, 38205 La Laguna, Tenerife, Spain 11 Departamento de Astrofísica, Universidad de La Laguna, 38206 La Laguna, Tenerife, Spain 12 Department of Physics, The University of Hong Kong, Pokfulam Road, Hong Kong Received 26 April 2016 / Accepted 30 July 2016 ABSTRACT Context. For over 12 yr, we have carried out a precise radial velocity (RV) survey of a sample of 373 G- and K-giant stars using the Hamilton Échelle Spectrograph at the Lick Observatory. There are, among others, a number of multiple planetary systems in our sample as well as several planetary candidates in stellar binaries. Aims. We aim at detecting and characterizing substellar and stellar companions to the giant star HD 59686 A (HR 2877, HIP 36616). Methods. We obtained high-precision RV measurements of the star HD 59686 A. By fitting a Keplerian model to the periodic changes in the RVs, we can assess the nature of companions in the system. To distinguish between RV variations that are due to non-radial pulsation or stellar spots, we used infrared RVs taken with the CRIRES spectrograph at the Very Large Telescope. Additionally, to characterize the system in more detail, we obtained high-resolution images with LMIRCam at the Large Binocular Telescope. Results. We report the probable discovery of a giant planet with a mass of mp sin i = 6.92+0.18−0.24 MJup orbiting at ap = 1.0860+0.0006 from the giant star HD 59686 A. In addition to the planetary signal, we discovered an eccentric (eB = 0.729+0.004 with a mass of mB sin i = 0.5296+0.0011 13.56+0.18−0.14 au. Conclusions. The existence of the planet HD 59686 Ab in a tight eccentric binary system severely challenges standard giant planet formation theories and requires substantial improvements to such theories in tight binaries. Otherwise, alternative planet formation scenarios such as second-generation planets or dynamical interactions in an early phase of the system's lifetime need to be seriously considered to better understand the origin of this enigmatic planet. Key words. Techniques: radial velocities -- Planets and satellites: detection -- Individual: (HD 59686 HIP 36616 HR 2877) -- Giant planets -- Binaries: spectroscopic −0.0007 au −0.003) binary companion −0.0008 M(cid:12) orbiting at a close separation from the giant primary with a semi-major axis of aB = 6 1 0 2 v o N 4 . ] P E h p - o r t s a [ 2 v 3 6 9 0 0 . 8 0 6 1 : v i X r a 1. Introduction Since the discovery of the first extrasolar planet around a solar- like star 20 yr ago by Mayor & Queloz (1995), more than 3000 (cid:63) Based on observations collected at the Lick Observatory, University of California. (cid:63)(cid:63) Based on observations collected at the Large Binocular Telescope, on Mount Graham, Arizona. (cid:63)(cid:63)(cid:63) Table 3 is only available in electronic form at the CDS via anony- mous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsarc.u- strasbg.fr/viz-bin/qcat?J/A+A/595/A55 exoplanets have been confirmed (Schneider et al. 2011)1. Of these planets, about 75% have been discovered by the transit method, which has greatly benefited from the data obtained with the Kepler Space Telescope (Borucki et al. 2010). Moreover, in 2014 alone, 715 new planets in 305 systems were detected by Kepler, almost doubling the number of exoplanets known at that time (Lissauer et al. 2014; Rowe et al. 2014), and more recently, Morton et al. (2016) have confirmed nearly 1280 new transiting Kepler planets based on probabilistic validation methods. Most 1 http://www.exoplanet.eu Article number, page 1 of 14 A&A proofs: manuscript no. HD59686_rev of the remaining 25% of planets have been found using the radial velocity (RV) technique. The spectral characteristics of solar- type main-sequence (MS) stars are favorable for RV measure- ments, which has made these stars the targets of the majority of the RV planet searches. However, a growing number of research groups are successfully searching for planets around evolved subgiant and giant stars (e.g., Döllinger et al. 2009; Johnson et al. 2010, 2011; Sato et al. 2013; Jones et al. 2015a,b; Niedzielski et al. 2015; Reffert et al. 2015; Wittenmyer et al. 2016). There are currently 95 known planetary companions orbit- ing around giant stars, of which 46% have been published during the past three years2. Giant stars allow us to access more massive stars than those typically observed on the MS. Early MS stars are normally avoided in RV planet searches as they rotate faster and have too few absorption lines for reliable high-precision RV de- terminations. On the other hand, evolved stars, such as K giants, have suitable and less broadened absorption lines for RV mea- surements, low rotational velocities, and much higher masses than late-type MS stars. Additionally, K-giant RV surveys also allow investigating how planetary systems evolve after the host star leaves the MS (Villaver & Livio 2009; Kunitomo et al. 2011; Villaver et al. 2014). Of the known extrasolar planets, about 7% orbit in multi- ple star systems3, although this number suffers from an observa- tional bias as most of the exoplanet surveys systematically avoid binary stars in their samples. For K-giant stars specifically, only four out of 72 known stars harboring planets are members of stellar multiple systems: 11 Com (Liu et al. 2008), γ1 Leo (Han et al. 2010), 91 Aqr (Mitchell et al. 2013), and 8 UMi (Lee et al. 2015). Finding planets in multiple star systems allows us to learn more about the processes of planetary formation and evo- lution. This is particularly important, since ∼ 50% of the MS stars in our solar neighborhood are members of binaries or mul- tiple systems (Duquennoy & Mayor 1991; Raghavan et al. 2010). The frequency of these planets may have a strong influence on the overall global frequency of extrasolar planets, allowing us to study the efficiency of planet formation mechanisms. Moreover, if there is any difference in the properties of planets in binary systems with respect to planets orbiting single stars, this may un- veil effects caused by additional companions in stellar systems (Desidera & Barbieri 2010; Roell et al. 2012). The majority of known planets in binary systems are in S- type orbits (circumstellar planets), meaning that the planet orbits around one member of the binary pair (e.g., Howard et al. 2010; Buchhave et al. 2011; Anderson et al. 2014), as opposed to P- type configurations (circumbinary planets), where the planet or- bits both stars beyond the binary orbit (e.g., Doyle et al. 2011; Orosz et al. 2012a,b; Welsh et al. 2012; Bailey et al. 2014). Most of the known S-type planets reside in wide-separation bi- naries (aB (cid:38) 100 au) where the influence of the stellar compan- ion on the formation process of the inner planet can probably be neglected. However, there are some interesting systems de- tected in close-separation binaries in which the stellar compan- ion is located at roughly 20 au: Gliese 86 (Queloz et al. 2000), γ Cep (Hatzes et al. 2003), HD 41004 (Zucker et al. 2004), and HD 196885 (Correia et al. 2008). The existence of planets in tight binary systems (aB (cid:46) 20 au) presents a serious challenge to current planet formation theories (Hatzes & Wuchterl 2005; Rafikov 2005). Moreover, supporting the theoretical expectation, Wang et al. (2014) found evidence that planet formation is effec- 2 http://www.lsw.uni-heidelberg.de/users/sreffert/giantplanets.html 3 http://www.univie.ac.at/adg/schwarz/multiple.html Article number, page 2 of 14 tively suppressed in binary systems with separations of less than 20 au. In this work, we report the discovery of a planet orbiting the giant star HD 59686, which we refer to as HD 59686 A, and which is part of a close-separation binary system with aB = 13.56 au. The paper is organized as follows: In Sect. 2 we de- scribe our sample selection and observations. Section 3 presents the stellar properties of the star and the Keplerian fit to the RV data from the Lick observatory. In Sect. 4 we validate the plan- etary hypothesis for the RV variations in HD 59686 A using in- frared RVs taken with CRIRES, spectral activity indicators, and the available photometry. In Sect. 5 we describe the high-contrast imaging observations of HD 59686 A obtained with LMIRCam at the Large Binocular Telescope (LBT) to image the stellar com- panion, including reduction of the data and constraints on the stellar companion to the giant star. In Sect. 6 we discuss the properties of the HD 59686 system, focusing on the nature of the stellar companion and the implications for the formation of planets in tight binaries. Finally, in Sect. 7 we present our con- clusions. 2. Observations We have continuously monitored the RVs of 373 G- and K-giant stars for more than a decade, resulting in several published planet detections (Frink et al. 2002; Reffert et al. 2006; Quirrenbach et al. 2011; Mitchell et al. 2013; Trifonov et al. 2014). Typical masses in our sample are between ∼1 -- 3 M(cid:12) and we reached RV precisions of ∼5 -- 8 m s−1. Among other results, we have found the first planet around a giant star (Frink et al. 2002) and showed that red giants with masses greater than ∼2.7 M(cid:12) host very few giant planets with an occurrence rate lower than 1.6% (Reffert et al. 2015). The original selection criteria aimed at observing 86 bright K-giant stars (V (cid:54) 6 mag) that were not variable or part of multi- ple stellar systems. Later in the survey, 93 new stars were added to the sample by imposing less stringent constraints regarding the photometric stability. Finally, in 2004, we added 194 G and K giants with bluer colors (0.8 (cid:54) B − V (cid:54) 1.2) with the aim of reducing the intrinsic RV jitter (e.g., Frink et al. 2001; Hekker et al. 2006). The inclusion of these stars allowed us to probe higher masses to test whether more massive stars host more massive planetary companions. More details on the selection criteria and on the giant star sample can be found in Frink et al. (2001) and Reffert et al. (2015). The RV observations of HD 59686 A were carried out us- ing the Hamilton Échelle Spectrograph (Vogt 1987) fed by the 0.6 m Coudé Auxiliary Telescope (CAT) of the Lick Observatory (California, USA). The Hamilton spectrograph covers the wave- length range 3755 -- 9590 Å and has a resolution of R∼ 60 000. The data were acquired and reduced using the iodine cell ap- proach as described by Butler et al. (1996). We currently have 11 -- 12 yr of data for our original set of K-giant stars, of which HD 59686 A is a member. In total, we have 88 RV measurements for HD 59686 A throughout this period of time. The Lick RVs along with their formal uncertainties are listed in Table 3. Typ- ical exposure times were 20 min, and the signal-to-noise ratios for these observations are typically about 120 -- 150. The result- ing RV measurements have a median precision of ∼5.6 m s−1. This value is below the RV jitter of 16.4±2.9 m s−1 expected for HD 59686 A based on scaling relations (see Chaplin et al. 2009; Kjeldsen & Bedding 2011). Additionally, using our K-giant sam- ple, we have obtained an empirical relation for the expected RV Table 1. Stellar parameters of HD 59686 A. Table 2. Orbital parameters of the HD 59686 system. Mauricio Ortiz et al.: Precise radial velocities of giant stars Value 5.45 0.52 ± 0.06 2.92 ± 0.30 1.126 ± 0.006 4658 ± 24 2.49 ± 0.05 0.15 ± 0.1 1.9 ± 0.2 13.2 ± 0.3 10.33 ± 0.28 96.8 ± 2.7 1.73 ± 0.47 2.6 K2 III Parameter Apparent magnitude mv (mag)a Absolute magnitude Mv (mag) Near-infrared magnitude K (mag)b Color index B − V (mag)a Effective temperature Teff (K)c Surface gravity log g (cm s−2)c Metallicity [Fe/H] (dex)d Stellar mass M(cid:63) (M(cid:12))c Stellar radius R(cid:63) (R(cid:12))c Parallax (mas)a Distance (pc) Age (Gyr)c Spectral type Notes. (a) Data from Hipparcos: van Leeuwen (2007) (b) Data from 2MASS: Skrutskie et al. (2006) (c) Reffert et al. (2015) (d) Hekker & Meléndez (2007) Parameter P (days) M0 (deg)a e ω (deg) K ( m s−1) m sin i (MJup) a (au) HD 59686 Ab HD 59686 B 11680+234−173 299.36+0.26−0.31 259+3−1 0.729+0.004 −0.003 149.4+0.2−0.2 4014+10−8 554.9+1.2−0.9 13.56+0.18−0.14 301+26−85 0.05+0.03−0.02 121+28−24 136.9+3.8−4.6 6.92+0.18−0.24 1.0860+0.0006 −0.0007 Notes. (a) This parameter is the value of the mean anomaly at the first observa- tional epoch t0 = 2451482.024 JD. 0.011 mas, deriving a radius of 11.62 ± 0.34 R(cid:12). Our estimate of the stellar radius for HD 59686 A of R(cid:63) = 13.2 ± 0.3 R(cid:12) agrees well with the value derived by Merand et al. (2004) and is slightly higher than the one obtained by Baines et al. (2008a). jitter as a function of color (see Frink et al. 2001; Trifonov et al. 2014) given by 3.2. Keplerian orbits log(RV jitter [m/s]) = (1.3 ± 0.1)(B − V) + (−0.04 ± 0.1), (1) where (B − V) is the color index. Using this relation, we expect an intrinsic RV jitter of 26.5 ± 9.2 m s−1 for HD 59686 A. This value is consistent at the 1.1σ level with the result derived from scaling relations. 3. Results 3.1. Stellar properties The stellar properties of the giant star HD 59686 A are given in Table 1. HD 59686 A is a slightly metal-rich star with [Fe/H]=0.15 ± 0.1 dex (Hekker & Meléndez 2007). To derive the stellar mass, we interpolated between the evolutionary tracks (Girardi et al. 2000), stellar isochrones, and metallicities using a trilinear interpolation scheme. This approach commonly allows two possible solutions depending on the evolutionary status of the star, namely red giant branch (RGB) or horizontal branch (HB). By taking the evolutionary timescale, that is, the speed with which the star moves through that portion of the evolu- tionary track, as well as the initial mass function into account, probabilities were assigned to each solution. The derived mass of HD 59686 A is M(cid:63) = 1.9 ± 0.2 M(cid:12) and the star was found to have a 89% probability of being on the HB. If it were instead on the RGB, then it would have a mass of 2.0 ± 0.2 M(cid:12), thus the mass is not affected within the uncertainties. More details on the method, including the stellar parameters of all K-giant stars in our Doppler survey, can be found in Reffert et al. (2015). The angular diameter of HD 59686 A was first calculated by Merand et al. (2004), using absolute spectro-photometric cal- ibration from IRAS and 2MASS observations. They derived a diameter of 1.29 ± 0.02 mas, which at the Hipparcos distance of 96.8+2.7−2.6 pc gives a value for the radius of 13.4 ± 0.4 R(cid:12). Later, Baines et al. (2008a) used the Center for High Angular Reso- lution Astronomy (CHARA) interferometric array (ten Brum- melaar et al. 2005) to measure an angular diamater of 1.106 ± We fitted Keplerian orbits to the RV data of HD 59686 A. The uncertainties were derived through bootstrapping (using 5 000 bootstrap replicates) by drawing synthetic samples from the original RV dataset with data replacement (see Press et al. 1992). We fitted for two companions in the system, to which we refer as HD 59686 Ab and HD 59686 B. In total, the Keplerian fit for HD 59686 Ab and HD 59686 B has 11 free parameters: the orbital period P, argument of perias- tron ω, RV semi-amplitude K, mean anomaly M0, and eccentric- ity e for each of the companions, and an arbitrary zero-point off- set. The RVs of HD 59686 A are shown in Fig. 1 along with the best Keplerian fit to the data. We also plot the individual signals of the planet HD 59686 Ab and the stellar object HD 59686 B. Error bars are included in all the plots. The best-fit orbital parameters for the planetary and stellar companions are given in Table 2. It is worth mentioning that K-giant stars exhibit intrinsic RV variability, known as stellar jitter. Therefore we decided to add in quadrature a jitter of 19.83 m s−1 , coming from the rms of the residuals around the fit, to our formal RV errors, which scaled down the χ2 red to a value of 1 (without jitter, χ2 =11.7). The rms of the residual RVs, after red subtracting the best Keplerian fit that includes the jitter, is 19.49 m s−1. This result is consistent with the intrinsic scatter expected from K-giant stars (Eq. 1, see also Hekker et al. 2008) and is within 1.1σ from the value derived using scaling relations. Figure 2 shows a generalized Lomb-Scargle (GLS) peri- odogram (Zechmeister & Kürster 2009) of HD 59686 A RVs. The top panel shows the results for the RV data, while the mid- dle and bottom panels show the periodogram for the residu- als after subtracting the stellar and stellar+planetary signals, re- spectively. The false-alarm probabilities (FAPs) were calculated by replacing the original RVs with randomly scrambled data through bootstraping. We computed the GLS periodogram 1 000 times for this new dataset and calculated how often a certain power level is exceeded just by chance. The estimated FAPs of 0.1%, 1%, and 5% are shown in the plot as the horizontal dotted, dashed, and dash-dotted blue lines, respectively. Article number, page 3 of 14 A&A proofs: manuscript no. HD59686_rev Fig. 1. Radial velocity measurements of the HD 59686 system. Note that a jitter of 19.83 m s−1 was added in quadrature to the formal RV uncertainties, and this is reflected in the plot. Upper left: Lick RVs together with the best Keplerian fit to the data. Upper right: RV residuals from the fit. Lower left: Phased RV variations and Keplerian fit for the ∼ 7 MJup planet HD 59686 Ab after subtracting the signal of the stellar companion. Lower right: RV data and Keplerian fit for the ∼ 0.5 M(cid:12) stellar object HD 59686 B after subtracting the planetary signal. The top panel shows one significant peak in the GLS pe- riodogram at ∼5000 days, which is approximately the length of time over which HD 59686 A has been observed. This wide peak represents the long period of HD 59686 B (P = 11680+234−173 days), for which one complete orbit has not been observed yet. How- ever, we are able to set tight constraints on the binary period as the eccentricty of the orbit is very high. The second strongest peak is at ∼340 days, very roughly matching the best Keple- rian fit for the planetary companion (P = 299.36+0.26−0.31 days). The third largest peak at ∼400 days is an alias period that disappears when the signal of the stellar companion is removed from the data. This is shown in the middle panel where the strength of the signal due to the planet increases significantly, and another alias period also appears at around 1700 days. After the planetary companion is subtracted, this peak disappears and no significant periodicities are observed in the signal of the RV residuals. By adopting a stellar mass of M(cid:63) = 1.9±0.2 M(cid:12), we derived a minimum mass of 6.92+0.18−0.24 MJup for HD 59686 Ab and a value of 554.9+1.2−0.9 MJup for HD 59686 B. The mass for HD 59686 B is equivalent to ∼ 0.53 M(cid:12), which immediately places this com- panion in the stellar regime; it cannot be a massive planet or a brown dwarf. The planet orbits the giant star at a distance of Article number, page 4 of 14 ap = 1.0860+0.0006 −0.0007 au, while the semi-major axis of the stel- lar companion is aB = 13.56+0.18−0.14 au. Furthermore, the orbit of HD 59686 B is very eccentric (e = 0.729+0.004 −0.003), which may have played an important role in the formation and/or evolution of the inner planet, as we discuss in Sect. 6. 4. Validating the planetary signal 4.1. Rotational modulation Stellar surface phenomena such as star spots, plages, or convec- tion modulated by stellar rotation may generate low-amplitude RV variations that can mimic planetary signatures. To investi- gate such false-positive scenarios, we determined the stellar ro- tation of HD 59686 A. Hekker & Meléndez (2007) estimated the projected rotational velocity of HD 59686 A to be v sin i = 4.28 ± 1.15 km s−1. Using our estimate of the stellar radius (R(cid:63) = 13.2 ± 0.3 R(cid:12)), we determine an upper limit for the rotation period of HD 59686 A of Prot(sin i(cid:63))−1 = 156.03 ± 39.35 days. This means that any low-amplitude RV variations generated by surface phenomena that are modulated by stellar rotation cannot have periods longer than ∼195 days. Therefore, it is unlikely that 20003000400050006000JD−2450000[days]−6000−4000−200002000RV[m/s]HD59686A0.00.20.40.60.81.0Phase−200−1000100200RV[m/s]HD59686Ab20003000400050006000JD−2450000[days]−100−50050100O−C[m/s]20003000400050006000JD−2450000[days]−6000−4000−200002000RV[m/s]HD59686B Mauricio Ortiz et al.: Precise radial velocities of giant stars Fig. 2. Top: GLS periodogram of the RV data of HD 59686 A. The significant peaks at ∼5000 and ∼340 days represent the orbits of the stellar and planetary companions, respectively. The 5000-day period is the time frame of our observations, therefore it is much shorter than the actual stellar period. The dotted, dashed, and dash-dotted lines show FAPs of 0.1%, 1%, and 5%, respectively. Middle: Periodogram of the residual RVs after the signal due to the stellar companion is removed from the data. Now the peak due to the planetary companion becomes much stronger and the alias period at ∼400 days disappears. Addition- aly, another alias period appears at ∼1700 days. Bottom: Periodogram of the residual RVs after subtracting the orbital fit (stellar+planetary companions); it shows no significant peaks. the periodic signal (P = 299.36+0.26−0.31 days) is generated by stellar rotation. Massarotti et al. (2008) estimated a slightly lower value for the projected rotational velocity of 3.8 km s−1 for HD 59686 A, which implies Prot(sin i(cid:63))−1 = 175.74 ± 46.42 days (assuming 1 km s−1 error in v sin i), consistent with the results of Hekker & Meléndez (2007) and with the above statement. On the other hand, Carlberg et al. (2012) calculated a value of v sin i = 0.93 ± 0.41 km s−1 for the K-giant, implying Prot(sin i(cid:63))−1 = 718 ± 495 days. However, this result has large uncertainties and, as dis- cussed by the authors, their estimates of v sin i show significant systematic differences when compared to values derived in the literature. Specifically, their estimates of v sin i are systematically lower than those reported in other studies (see Fig. 6 of Carlberg et al. 2012), which can be accounted for by an overestimation of the macroturbulence velocity, particularly in the slow-rotation regime. Regardless of the above considerations, to test the hypothesis that the 299.36+0.26−0.31 day period may be caused by stellar activ- ity, like long-period pulsations for example, we checked infrared RVs, all available photometry, and spectral activity indicators as described in the following sections. 4.2. Infrared radial velocities It is recognized that intrinsic stellar activity, such as cool spots, can create RV variations in giant stars that can mimic the pres- ence of companions (e.g., Hatzes & Cochran 2000; Hatzes et al. 2004). This poses an additional challenge for validating the in- terpretation of a periodic RV change as a bona fide planet, when compared to inactive MS stars. Moreover, some giant stars are known to be pulsating stars, which show several modes of pul- Fig. 3. Upper panel: CRIRES infrared RV measurements of HD 59686 A. The black solid line shows the best Keplerian fit obtained from the Lick data alone. Bottom panel: Residuals of the CRIRES RVs from the optical fit. The value of the rms is ∼59 m s−1, which is consis- tent with the large infrared RV errors. The mean error of the CRIRES data, with a jitter of 19.83 m s−1 added in quadrature, is ∼45 m s−1. sation with varying amplitudes and frequencies (De Ridder et al. 2009; Huber et al. 2010; Christensen-Dalsgaard et al. 2012; Stello et al. 2013). In the case of radial pulsations, the stellar surface moves away and toward the observer, which induces pe- riodic RV variations. The pulsation frequencies of a star are closely related to its density and temperature, as these control the speed at which sound waves can propagate. Using the scaling relation of Kjeld- sen & Bedding (1995), we calculated the period of the pulsation with maximum amplitude using our derived values of the radius, mass, and effective temperature, which yielded a value of 0.31 days for HD 59686 A. Although this calculation is not ideal for giant stars, it should give a reasonable estimate of the pulsation period with the largest amplitude. This value is orders of magni- tudes below the RV oscillations seen in our data. It is possible, though unlikely, that some pulsation exists in HD 59686 A with a much lower frequency, but large enough am- plitude to be detectable in our data, which could be the source of the RV oscillations we observe. Non-radial pulsations are much more complicated to model, and they can display an arbitrary number of amplitudes and periods for different modes. However, it is not expected that the RV amplitude of the pulsations in the visible waveband match the amplitude in the infrared, since the photometric variations of pulsating stars are wavelength depen- dent (e.g., Percy et al. 2001, 2008). On the other hand, if the RV oscillations are due to a companion, then the infrared and visible RV variations should be consistent with each other. In 2012 and 2013, Trifonov et al. (2015) obtained infrared RVs of HD 59686 A using the CRyogenic high-resolution In- fraRed Echelle Spectrograph (CRIRES; Käufl et al. 2004) at the Very Large Telescope (VLT), in Chile. Their CRIRES spectra have a resolution of R∼ 100 000 and cover the wavelength range 1.57 -- 1.61 µm. Details of the CRIRES observations and the re- duction process, including the measured RVs for HD 59686 A, can be found in Trifonov et al. (2015). We obtained the RV offset between the CRIRES and Lick velocities for HD 59686 A by fitting the CRIRES and Lick RVs keeping all the orbital parameters fixed. Figure 3 shows the CRIRES RV data (with the RV offset applied) together with the Article number, page 5 of 14 0.00.20.40.60.81.0Power0.00.20.40.60.81.0Power100101102103104Period[days]0.00.10.20.30.4Power−2000−1800−1600−1400−1200−1000RV[m/s]CRIRES58005900600061006200630064006500JD−2450000[days]−100−50050100O−C[m/s] A&A proofs: manuscript no. HD59686_rev HD 59686 A collected over seven years between December 13, 2002 (HJD = 2452621.84) and November 24, 2009 (HJD = 2455159.78). Unfortunately, HD 59686 A is a very bright tar- get (V=5.45) and exceeds the ASAS-3 V-band saturation limit with the used exposure times (180 seconds). The high dis- persion of the ASAS-3 V-band photometric measurements of HD 59686 A (peak-to-peak amplitude of 0.784 mag, mean value ¯V = 5.74 ± 0.19 mag) and the mean value of the errors (38.5 mmag) ensure that HD 59686 A saturates the ASAS-3 detector. The only unsaturated photometry for HD 59686 A was ac- quired by the Hipparcos mission (ESA 1997) between March 16, 1990 and March 10, 1993 (2447966.9 -- 2449057.2 JD), more than six years before first RV observations of HD 59686 A. The Hipparcos data set consists of 96 measurements with 5.6 mmag mean error, 5.6 mag mean value, and a standard deviation of 5.5 mmag, similar to the mean error of the measurements. As shown in the bottom panel of Fig. 4, no significant periodic signal was found in the photometry of these data. Additionally, we can use the Hipparcos data to investigate whether a hypothetical spot might have produced a noticeable photometric variation. We de- rived the spot filling factor that would be required to generate the observed RV amplitude of ∼137 m s−1 using the relation found by Hatzes (2002) for cool spots on sun-like stars. We obtained a spot filling factor of f = 0.1, meaning that 10% of the stellar surface must have been covered by spots to produce the large RV variation seen in the data. Using this value for the filling fac- tor, the expected photometric variability is ∆m = 0.078 mag for a temperature difference of ∆T = 1200 K between the spot and the stellar photosphere. This level of variation is one order of mag- nitude above the observed dispersion seen in the Hipparcos data. The same is true for a wide range of temperature differences of typical star spots ranging from ∆T = 200− 1200 K (e.g., Biazzo et al. 2006; O'Neal 2006). Nevertheless, hypothetic surface structure phenomena might mimic the presence of an exoplanet. For example, Hatzes & Cochran (2000) investigated the possible existence of a macro- turbulent spot to explain the RV variation of Polaris. Given the right conditions, this dark spot might cause a large RV oscil- lation without a significant photometric variation. However, the values of, for example, the magnetic field and the difference be- tween the velocity of the macroturbulent spot and the surround- ing surface must be exceptionally well fine-tuned to produce an RV variation of hundreds of m/s. In addition, if a macroturbu- lent spot causes the RV changes in HD 59686 A, then it must have been long-lived and maintained a constant and consistent effect during at least 12 yr. The same is true for long-lived long- period non-radial pulsations, which is not necessarily expected. Thus, although we cannot completely discard this scenario, a gi- ant planet orbiting the star HD 59686 A appears as the most plausible interpretation of our data. 4.4. Spectral activity indicators Since the RV measurements of HD 59686 A were acquired using the iodine-cell method, it is difficult to perform precise bisector measurements of spectral lines as the stellar spectra are affected by I2 lines. Instead of this, we performed an analysis of the Hα line, which is located in the central region of one of the Hamil- ton spectrograph orders and is known to be a good indicator of stellar activity. We measured the Hα index using the approach presented by Kürster et al. (2003). However, we broadened the width of the window centered on Hα, from ±15 km s−1 used by Kürster et al. (2003) for Barnard's star to ±45.68 km s−1 (±1 Å) Fig. 4. Upper panel: GLS periodogram of the Hα index measurements of HD 59686 A. Bottom panel: GLS periodogram of the Hipparcos V- band photometry of HD 59686 A. The dotted, dashed, and dash-dotted lines in both panels show FAPs of 0.1%, 1%, and 5%, respectively, ob- tained by bootstraping. No significant periodicities are found in the data. best Keplerian fit to the Lick data. The infrared RVs match the Keplerian model obtained from the optical data. This should in general not be the case if the RV variations were due to large amplitude stellar pulsations. Moreover, the scatter around the fit of ∼59 m s−1 is consistent with the relatively large uncertainties4 of the CRIRES RVs that are on the order of ∼45 m s−1. An additional test can be made by fitting only the CRIRES data to derive the RV semi-amplitude, KIR. Following Trifonov et al. (2015), we first subtracted the signal of the stellar compan- ion from the CRIRES data. As the presence of HD 59686 B is clearly detected in the system, it is fair to assume that the RV signal due to this star is consistent in the optical and infrared data sets. Then, we performed a Keplerian fit to the CRIRES RVs keeping all parameters fixed (the parameters obtained from the Lick RVs) with the exception of the RV semi-amplitude and RV zero point. We derived a value of KIR = 206.0 ± 29.1 m s−1. The RV semi-amplitude of Kopt = 136.9+3.8−4.6 m s−1 from the op- tical RVs is within 2.25σ from the IR value. If we calculate κ = KIR/Kopt , as in Trifonov et al. (2015), then we obtain a value of κ = 1.50± 0.22, but we note that the calculated error might be underestimated as the error on the fitting of the stellar compo- nent is not taken into account. This result shows that the near-IR signal is not flat or of a smaller amplitude than the optical one, which we would expect for a spot or pulsations; the amplitude of pulsations decreases with increasing wavelength in pulsating giant stars (Huber et al. 2003; Percy et al. 2008). We also have only a handful of moderately precise IR RVs and in addition, a stellar jitter of about 20 m s−1 for HD 59686 A, but we observe that the optical and near-IR phases are consistent, which is not necessarily expected for pulsations. This means that most likely the signal is real and caused by the gravitational perturbation of a companion in the system. 4.3. Photometry The ASAS-3 Photometric V-band Catalog (Pojma´nski 1997, 2001) contains 290 best-quality measurements (grade A) of 4 To be consistent with the optical fit, a jitter of 19.83 m s−1 was added to the formal CRIRES RV uncertainties. Article number, page 6 of 14 0.000.050.100.150.200.250.300.350.40Power100101102103104Period[days]0.000.050.100.150.200.250.300.350.40Power Mauricio Ortiz et al.: Precise radial velocities of giant stars available observational evidence at hand (e.g., Hα index, pho- tometry, infrared RVs) supports the planetary hypothesis, unless some exotic not-yet-observed surface macroturbulent structure or long-lived long-period non-radial pulsation was taking place in HD 59686 A for more than a decade, which we consider un- likely. Nevertheless, there is much that we do not know about long-period stellar oscillations in giant stars, and we cannot fully exclude such phenomena. 4.5. Discarding a hierarchical triple star system Another possibility that can mimic planets in binary systems are hierarchical triple systems in which the observed RV signals are caused by another star orbiting the binary companion instead of a planet orbiting the primary star. For instance, Schneider & Cabr- era (2006) and Morais & Correia (2008) studied the effects on the RV measurements of a star orbited by a pair in a close circu- lar orbit in a triple star system. They concluded that the effect of the binary is approximately weaker than ∼1 m/s in the RV semi- amplitude and can only mimic a low-mass Earth- or Saturn-like planet. Later, Morais & Correia (2011) extended their work to triple star systems on eccentric orbits, showing that the binary effect is stronger than in the cirular case. However, the magni- tude of the RV semi-amplitude is still about a few meters per second and cannot account for the large variation that we see in the RV data of HD 59686 A (K ∼137 m s−1). Furthermore, we can estimate the effect that a binary star system with a to- tal mass of ∼0.5 M(cid:12) with a period of ∼300 days can generate in the RV semi-amplitude (using Eq. 37 of Morais & Correia 2011). For reasonable values of the amplitudes of the frequency terms induced by a hypothetical third star in the system, we ob- tained a value of the RV semi-amplitude mimicking a planet ranging from ∼1−5 m s−1, that is, more than an order of magni- tude smaller than what we observe in our data. We therefore con- clude that a hidden star orbiting the stellar object HD 59686 B is not the cause of the observed RV variations in the system. 5. High-contrast images 5.1. Previous search for stellar companions in HD 59686 A HD 59686 A has been examined before for stellar companions. Roberts et al. (2011) found a visual component separated by 5.61(cid:48)(cid:48). Assuming a face-on circular orbit, this corresponds to a minimum separation of ∼519 au. If this component were a phys- ical companion, then the separation would lead to an orbital pe- riod far too large to be visible in our data. Baines et al. (2008b) have also observed HD 59686 A using the CHARA interferome- ter. They performed fits to the diameter of several stars and found that single stars were consistently fit with low values of χ2, while the presence of a stellar companion created a systematic behav- ior in the residuals, resulting in a high χ2 value. They saw no such systematic behavior in the fit of HD 59686 A and therefore ruled out a MS companion more massive than G5 V within a field of view of 230 mas (∼23 au). Baines et al. (2008b) also searched for small-separation bina- ries by looking for separated fringe packets in the data. If a sec- ond star were present in the system with a separation of around 10 to 100 mas (∼1 -- 10 au), then two sets of fringe packets would be detected. However, no separated fringe packet was observed for HD 59686 A. This approach relies on the assumption that the angular separation of the two stars is not small and that the po- sition angle is not perpendicular to the projected baseline angle. Most likely, the authors failed to detect HD 59686 B because Article number, page 7 of 14 Fig. 5. Upper panel: Hα index measurements at the time of each RV observation of HD 59686 A. Bottom panel: Hα index measurements as a function of the RVs due to the planetary companion HD 59686 Ab, that is, with the stellar component subtracted from the data. No significant correlation is seen in the data, which corroborates that a giant planet is part of the system. recently used by Hatzes et al. (2015) for Aldebaran. As refer- ence windows we used spectral regions that extend from −250 and −650 km s−1 and from 250 and 650 km s−1. The upper panel of Fig. 4 shows the GLS periodogram of the Hα index measure- ments. As for the Hipparcos photometry, no significant signal exist in the Hα index of HD 59686 A. Figure 5 presents the Hα index against the time of each RV observation of HD 59686 A and as a function of the RV variation induced by the planet HD 59686 Ab (without the contribution of the stellar compan- ion). The plot shows no correlation between these RVs and the Hα index. Moreover, we measured a Pearson correlation coeffi- cient of r = 0.06 with a p-value=0.58. This analysis corroborates that the 299.36+0.26−0.31 day period in the RV curve of HD 59686 A is most likely generated by the gravitational pull of a planetary companion. It is worth to note, however, that HD 59686 A shows some similarities to carbon-enhanced metal-poor (CEMP) stars (see Beers & Christlieb 2005; Masseron et al. 2010; Placco et al. 2014) in the sense that these are evolved giants, they reside in binary systems, and the secondary is very likely a white dwarf (provided that HD 59686 B is confirmed to be a white dwarf). Recently, Jorissen et al. (2016a) has identified low-amplitude RV variations superimposed on the binary trend in 3 CEMP stars in a sample of 13. They show periods of about one year and RV semi-amplitudes of hundreds of m/s. Jorissen et al. (2016b) dis- cussed the origin of the RV variations of one system in particular, HE 0017+005, and suggested that this may be due to pulsations in the envelope of the giant star. Unfortunately, the spectral types of the stars from Jorissen et al. are not well established. The au- thors assumed that all the stars have masses of ∼0.9 M(cid:12), and it is likely that these very metal-poor stars are in a different stage of the stellar evolution than HD 59686 A, which we expect to be on the HB with a 89% probability (see Reffert et al. 2015). In particular, the log g values of the RV-variable CEMP stars seems to be much lower than that of HD 59686 A (see Jorissen et al. 2016b), which makes pulsations much more plausible for those stars. Even if pulsations should be confirmed as the correct in- terpretation of the RV variations observed in CEMP stars, this will probably not be the case for HD 59686 A because all the 20003000400050006000JD−2450000[days]0.0500.0550.0600.0650.070Hαindex−200−150−100−50050100150200RadialVelocity[m/s]0.0500.0550.0600.0650.070Hαindex A&A proofs: manuscript no. HD59686_rev Fig. 6. High-contrast L(cid:48)-band LMIRCam images of HD 59686 A. Panels a, b, c, and d show the residual images after running the PCA with 10, 38, 77, and 81 principal components. Panel e shows the image obtained with a local, subannular PCA approach, and panel f presents the residual image after subtracting the stellar PSF using the new LLSG algorithm. No signal of the companion star HD 59686 B is detected in any of the panels. this star is expected to be much fainter than the giant primary and was probably below the contrast sensitivity of CHARA. With the aim of investigating the nature of the stellar object HD 59686 B, we acquired high-resolution images of this system as explained in the following sections. 5.2. Observations and data reduction The high-contrast imaging observations of HD 59686 A were carried out on February 9, 2014 using the L/M-band InfraRed Camera (LMIRCam; Skrutskie et al. 2010; Leisenring et al. 2012) mounted at the Large Binocular Telescope Interferome- ter (LBTI; Hinz et al. 2012) on Mt. Graham, Arizona. LMIR- Cam is a high-resolution camera designed to operate in the 3−5 µm wavelength range. The infrared detector is a 1024×1024 HgCdTe array, with a plate scale of 10.707±0.012 mas/pix (Maire et al. 2015) and a field of view of 11×11(cid:48)(cid:48). The observations were taken using only the left side of the LBT in pupil-stabilized mode, which further allows the use of angular differential imaging (ADI; Marois et al. 2006). The core of the PSF was intentionally saturated to increase the signal of the binary companion. Unsaturated exposures with a neutral den- sity filter were also taken for calibrating the photometry. The AO system was locked with 300 modes during the whole duration of our observations. We obtained 205 minutes of on-source in- tegration and ∼100◦ of field rotation. A total of 7 413 images of HD 59686 A were taken in the L(cid:48)- band filter (λc=3.70 µm, ∆λ=0.58 µm). To properly subtract the background emission and detector systematics, the star was dithered to two different positions on the detector separated by 4.5(cid:48)(cid:48). Additionally, our reduction steps Article number, page 8 of 14 included dark current subtraction, flatfielding, bad pixel correc- tion, bad image removal, image alignment, and trimming of the data. We were left with a 300×300 pixel datacube of 5487 re- duced images. However, during large parts of the observing se- quence, weather conditions were not optimal (seeing >1.5(cid:48)(cid:48)), so that we decided to discard 20% of the images based on the mea- surement of the correlation of each one of the frames with re- spect to a high-quality reference frame of the sequence. In total, we obtained a datacube of 4389 images. 5.3. PSF subtraction In addition to simple ADI processing, more sophisticated algo- rithms exist, such as the locally optimized combination of im- ages (LOCI; Lafrenière et al. 2007) and principal component analysis (PCA; Amara & Quanz 2012; Soummer et al. 2012; Brandt et al. 2013). They can be used to subtract the light profile of a star to detect possible companions around it. We decided to follow a PCA approach, as it has been shown to produce bet- ter contrast performance for small inner working angles (e.g., Meshkat et al. 2014). The expected binary separation at the time of our observations is small, so that even with the PCA technique it is challenging to detect any signal at all, considering that we do not know the orbital inclination and orientation of the orbit. To analyze our stack of images, we used the open-source Python package VIP5 (Gomez Gonzalez et al. 2016a), which provides a collection of routines for high-contrast imaging pro- cessing, including PCA and slight variations of it, such as an- nular and subannular PCA. The PCA algorithm models the star 5 https://github.com/vortex-exoplanet/VIP 0.200PCA10 PCs(a)0.200PCA38 PCs(b)0.200PCA77 PCs(c)0.200PCA81 PCs(d)0.200LOCAL PCA(e)0.200LLSG(f)3001500150300502502550301501530Counts [ADU]301501530402002040201001020Counts [ADU] Mauricio Ortiz et al.: Precise radial velocities of giant stars a Gaussian noise part. The low-rank carries most of the signal from the stellar PSF, the Gaussian noise captures the quasi-static features of the speckle noise, and the sparse component contains the signal of potential faint companions. The most important pa- rameter to set in the LLSG algorithm is the rank, which is equiv- alent to set the number of PCs in the standard PCA. We chose a rank of 51 as the mean of the optimum number of PCs derived before. We note, however, that varying the rank number does not change the obtained results significantly. The residual image af- ter the LLSG subtraction is shown in panel f of Fig. 6. Although the quality of the image seems to be much better than in pre- vious images, we did not detect any signal from the binary star HD 59686 B. The obtained results can be explained by (i) the poor weather during some part of the observations, (ii) the small expected an- gular separation of the companion, and (iii) the probability that the orbit orientation placed the star at a projected separation such that the companion is not visible from Earth at the time of obser- vation. 5.4. Contrast curve calculation Assuming that the orbital configuration is favorable at the ob- serving time and that we are only limited by the contrast of the binary pair and the quality of our images, we can set constraints on the maximum brightness that the companion star could have without being detected in our images. To do that, we injected fake companions of various magnitudes at different distances from the central star. As a fake companion star we used the median-combined PSF of the unsaturated data set and scaled it to different contrast ratios based on the photometry of the unsatu- rated image of HD 59686 A and taking into account the different exposure times between the saturated and unsaturated frames. The fake stars were then inserted in each of the reduced stack of images, accounting for the change in parallactic angle dur- ing the rotation sequence. We then processed these images with the VIP package in the exact same way as before and calculated the 5-σ detection limit in terms of S/N at the position of each fake star. We adopted the S/N defintion of Mawet et al. (2014) as we are working at distances very close to the center of the star, and the low-pixel statistics applies. We repeated this procedure at four different position angles for each radius and then took the average to minimize random speckle errors. In Fig. 7 we show the 5σ contrast curve of the LBT images as a function of the angular separation from the central star. Our data reach contrasts between ∼5 -- 10 mag for separations between ∼0.16(cid:48)(cid:48) -- 0.24(cid:48)(cid:48)(15.5 -- 23.2 au). We also show the maximum ex- pected binary separation at the observing time of amax ∼ 11.7 au (black dashed line) and the PSF saturation radius of rs ∼ 8.3 au (black dash-dotted line). The expected separation of the binary pair comes from a detailed study of the dynamical stability of the HD 59686 system that constrains the orbital inclination to the range i ∼ 50◦ -- 90◦ (Trifonov et al. 2016, in prep.). For an in- clination of 50◦, we derived the value of 11.7 au, which translates into ∼0.12(cid:48)(cid:48)of angular separation. Adopting higher values for the inclination results in lower values for the binary separation. Un- fortunately, the large saturation radius of the LMIRCam images (∼0.085(cid:48)(cid:48)) prevents us from deriving reliable values for the 5σ contrast in the region (cid:46) 0.15(cid:48)(cid:48) ((cid:46) 14.5 au), in which we expect HD 59686 B to reside. Nevertheless, we show in the plot (red solid lines) the expected contrasts for a star of 0.5 and 1 M(cid:12). A G-type star of 1 M(cid:12) or greater is excluded for separations (cid:38) 17 au. For lower masses and separations our sensitivity decreases Article number, page 9 of 14 Fig. 7. 5-σ detection limits in terms of the magnitude contrast in the L(cid:48)-band as a function of the distance from the central star. The black dashed line represents the binary separation upper limit of ∼11.7 au at the time of the observations. The dash-dotted line marks the saturation radius of ∼0.085(cid:48)(cid:48). The red solid lines mark the expected contrasts for a G2 V star of 1 M(cid:12) and a M0 V star of 0.5 M(cid:12), from top to bottom. light as a linear combination of a set of orthogonal basis func- tions or principal components (PCs) and fits for the PC coeffi- cients in each of the frames in the stack. This means that the pa- rameter that must be set is the number of PCs used to model the PSF in each frame. We started by estimating the optimal number of PCs by inserting a star in each of the images at a small sepa- ration from the center of the primary star. We varied the magni- tude difference of this fake companion with the central star from ∆m=8−11 mag in steps of 0.5 mag and determined the number of PCs that maximizes the S/N in an aperture of 1 FWHM cen- tered on the coordinates of the fake star after running the PCA. We searched in a grid ranging from 1-200 PCs and found that the highest S/N values were obtained for 10, 38, 77, and 81 PCs. The central saturated core of the PSF (eight-pixel radius) was masked and not considered in the fitting. We show in Fig. 6, panels a to d, the results after running the PCA in the stack of images of HD 59686 A using the previously derived numbers of PCs. No significant signal was found in the residual images. Additionally, we also performed a local PCA by fitting for the stellar PSF in quadrants of circular annuli of 3 FWHM width around the central star. In this case, the PCA is computed locally in each quadrant, and we applied a parallactic angle rejection of 1 FWHM to discard adjacent frames and avoid self-subtraction of the companion star. The number of PCs was decided automati- cally in each quadrant by minimizing the residuals after subtract- ing the PSF. The resulting resdiudals image is shown in panel e of Fig. 6. As in the full-frame PCA, no significant companion is seen in the plot. As an alternative to the standard PCA, we also used the new algorithm recently introduced by Gomez Gonzalez et al. (2016b) to subtract the stellar PSF of high-contrast images and enhance the signal of faint companions. The method is named by the au- thors local low-rank plus sparse plus Gaussian-noise decompo- sition (LLSG). The main idea of the algorithm is to use a robust PCA approach (see, e.g., Candès et al. 2009) to decompose the stellar image into three components; a low-rank, a sparse, and 0.090.110.130.150.170.190.210.230.25Angularseparation[arcsec]10864205σcontrastL0[mag]G2VM0V10152025Projectedseparation[AU] A&A proofs: manuscript no. HD59686_rev Fig. 8. Sky-projected orbit of the HD 59686 binary system assuming values of i = 50◦ and Ω = 45◦ for the orbital inclination and longitude of the ascending node, respectively. Labeled in the orbit are the positions of each star as a function of time. The green symbol marks the position of HD 59686 A and HD 59686 B at the time of our LBT observations. The dotted line is the line of nodes, and the letter P denotes the positions of the stars at periastron. The yellow dot marks the center of mass of the system. significantly, and we cannot exclude the presence of a star with masses between 0.5−1 M(cid:12). To illustrate the configuration of the binary system, we show in Fig. 8 the sky-projected orbit of HD 59686 AB derived from the fitted orbital parameters. The red labels mark the position of each of the stars at certain times (in years). The green sym- bols highlight the respective locations of HD 59686 A and HD 59686 B in the binary orbit at the time of the LBT observa- tions. The high eccentricity of the binary is clearly visible. Fortu- nately, both components are moving away from each other at the moment, so that it should become easier to detect HD 59686 B in the coming years. In about ∼2025, the system will be in apas- tron at a minimum separation of roughly ∼20 -- 21 au assuming an inclination of i = 90◦. For lower values of the inclination the bi- nary separation increases. Future high-resolution observations of this system are highly encouraged to better constrain the nature of the stellar object HD 59686 B. 6. Discussion 6.1. HD 59686 Ab: a planet in a close-separation binary Among the known S-type planets, HD 59686 Ab is very pecu- liar, mainly because it is part of a close-separation (aB = 13.6 au) and eccentric (eB = 0.7) binary system. Figure 9 shows the semi-major axis of the known S-type planets as a function of the binary separation. Planets exist in binaries with a wide range of separations, but it is clear that the majority of them show semi-major axes greater than aB ∼ 100 au. HD 59686 AB is, together with ν Octantis (Ramm et al. 2009) and OGLE-2013- BLG-0341LB (Gould et al. 2014), the binary with the closest separation of its stellar components known to harbor a planet. Article number, page 10 of 14 Fig. 9. Semi-major axis of planetary companions plotted against binary separation for all known planet-hosting binary systems. Shown are bi- naries with MS (blue circles) and evolved subgiant/giant (red triangles) primary stars as well as two microlensing binaries (green squares), in which the spectral type of the stars is not known. The filled symbols show binaries in which the secondary star is a white dwarf. The posi- tion of the HD 59686 system is marked with a red cross. The dashed line marks the 1:1 relation between planet semi-major axis and binary separation. Most of the discovered planets are found in binary stars with separations greater than ∼100 au. The microlensing Earth-mass planet OGLE-2013-BLG- 0341LB b is orbiting at approximately ∼0.8 au from its host star, and the microlensing models are compatible with a bi- nary separation of either ∼ 10 or 15 au. The case of ν Oct is particularly remarkable, since the separation of the binary pair is only aB ∼ 2.6 au and the conjectured planet is orbiting at ap ∼ 1.2 au; roughly at half the distance between both stars. Interestingly, similar to HD 59686 AB, the ν Oct system is com- posed of a single-lined K-giant binary, with a secondary star mass of ∼ 0.55 M(cid:12). Moreover, the ν Oct system is slighlty ec- centric: e ∼ 0.25 (Ramm 2015). As we discussed below, the existence of giant planets in both systems is very hard to explain by traditional theories. There are two additional systems (not included in the plot) with reported companions at a (cid:46) 20 au: KOI-1257 (Santerne et al. 2014) and α Cen (Dumusque et al. 2012). KOI-1257 b is a transiting giant planet with a period of P = 86.6 days that is part of a binary system with aB ∼ 5.3 au. However, the nature of the massive outer companion in the system is unconstrained at present; it could be anything, a planet, a brown dwarf or a stellar object (Santerne et al. 2014). On the other hand, in α Cen AB, the stellar nature of the binary components is well established, but the existence of a terrestial planet orbiting at ∼ 0.04 au has recently been questioned (Hatzes 2013; Rajpaul et al. 2016), im- plying that most likely there is no planet in the α Cen system. This would make HD 59686 AB, and ν Oct, unique systems in which to study the formation of giant planets in short-separation binaries. Another striking property of the HD 59686 system is the high eccentricity of the binary pair. With a value of eB = 0.729+0.004 −0.003, this is the most eccentric close-separation binary (aB (cid:46) 20 au) known to harbor a planet. This implies that, at periastron, both stars are separated by only ∼3.6 au. The formation of such a sys- tem presents a tremendous challenge to current planet formation theories as the smallest binary separation in which giant planets -200-150-100-50050100x offset [mas]-200-150-100-50050100y offset [mas]PP2010201220142016201820202022202420102012201420162018202020222024i = 50.0 Ω = 45.0 100101102103104105Binaryseparation[AU]10−210−1100101102Semi-majoraxisofplanet[AU] Mauricio Ortiz et al.: Precise radial velocities of giant stars Fig. 10. Final periods and eccentricities resulting from all simulations that led to the formation of a HB giant star with a WD companion. The initial masses of the stars were 1.9 and 2.3 M(cid:12). We also show histograms reflecting the distribution of final periods and eccentricities. The position of the HD 59686 system is marked with a red asterisk. Left panel: Results for BW = 0, meaning that the mass loss is treated with the traditional Reimers prescription. It is clear that none of the simulations can reproduce the HD 59686 system. In the majority of the cases the orbit is fully circularized due to tidal interactions during the AGB phase. Right panel: Results for an enhanced mass-loss rate (BW = 104) showing that a large fraction of initial orbital conditions lead to long-period and eccentric binaries similar to the HD 59686 system. The black solid line marks the period and eccentricity evolution of the model that agrees best with the orbital properties of HD 59686 AB. could form is thought to be ∼ 20 au (see Haghighipour 2009, and references therein). On the other hand, simulations have shown that terrestrial planets may form in close-separation binaries up to ∼ 0.2qb, where qb is the binary pericenter distance (Quintana et al. 2007). Therefore, this possibility is not directly excluded in the HD 59686 system, as terrestial planets might have formed up to a distance of ∼0.7 au from the primary star. 6.2. Nature of the stellar object HD 59686 B With the mass for HD 59686 B constrained in the range 0.53 -- 0.69 M(cid:12) derived from dynamical simulations, there are two op- tions for this stellar companion: it may be a typical dwarf star or, more interestingly, a white dwarf (WD). The latter possibility is not rare as there are currently three known circumstellar plan- ets orbiting stars with WD companions: GL 86 (Queloz et al. 2000), HD 27442 (Butler et al. 2001), and HD 146513 (Mayor et al. 2004). Interestingly, the system GL 86 AB is also a close- separation binary with a semi-major axis of aB = 18.4 au. With the currently available data we cannot assess the nature of the stellar object HD 59686 B with certainty, but nevertheless, we can investigate whether the WD scenario is plausible given the current orbital parameters and derived masses of the HD 59686 system. If HD 59686 B is indeed a WD, then its mass must origi- nally have been greater than the mass of HD 59686 A (1.9 M(cid:12)) because it evolved faster to a later stage of the stellar evolu- tion. The problem now resides in estimating the inital MS mass of HD 59686 B. The initial-final mass relationship (IFMR) for WDs has been a subject of intense research in the past (Weide- mann 1977, 1987, 1990; Jeffries 1997). More recently, Kalirai et al. (2009) calibrated a semi-empirical relation for the IFMR us- ing several WDs found in a set of globular clusters in the Milky Way. With this relation we can estimate an inital MS mass for HD 59686 B of ∼0.7−2.3 M(cid:12) for WD masses of 0.53 and 0.69 M(cid:12), respectively. The latter mass satisfies our intial constraint of MB > 1.9 M(cid:12). This means that, for the upper limit of our mass estimate, HD 59686 B may have evolved off the MS to end its life as a WD of ∼0.69 M(cid:12). To investigate whether this scenario is plausible, we used the detailed binary evolution code BSE (Hurley et al. 2002) to evolve a binary star pair with a set of initial orbital properties. The initial binary masses were set to 2.3 and 1.9 M(cid:12). We considered a range of initial periods and eccentricities of P = 5000 − 30000 days in steps of 100 days and e = 0.50 − 0.99 in steps of 0.01. The system was then evolved until the stellar types of the two stars were a WD and a HB star. The results of the simulations are shown in Fig. 10, where the final periods and eccentricities are plotted for all the different initial orbital configurations that led to a WD-HB pair with similar masses as those observed in the HD 59686 system. The left panel shows the results for a mass- loss prescription given by the traditional Reimers formula for red giants (Reimers 1975). It is clear that no set of initial conditions can reproduce the current orbital properties of the HD 59686 system, namely a period of P = 11680 days and eccentricity of e = 0.729. Orbits with periods of a few thousands days or less are fully circularized, and the small fraction of systems with a high eccentricity (e ∼ 0.6) shows very long orbital periods of ∼50000 days. The right panel of Fig. 10 shows the final periods and eccen- tricities for the same initial configurations as discussed before, but with an increased mass-loss rate controlled by the enhanced Article number, page 11 of 14 102103104105Period [days]0.00.20.40.60.81.0Eccentricity0300600900120015000100020003000400050006000102103104105Period [days]0.00.20.40.60.81.0Eccentricity050010001500200025000100200300400500600 A&A proofs: manuscript no. HD59686_rev wind parameter BW. This parameter was first introduced by Tout & Eggleton (1988) to explain the mass inversion of the RS CV binary star Z Her. In this scenario it is assumed that the mass loss is enhanced through tidal interactions with a binary companion. Tidally enhanced stellar winds have been used since then to ac- count for several phenomena related to giant stars in binary sys- tems, such as the eccentricities of barium stars (Karakas et al. 2000; Bonaci´c et al. 2008), symbiotic channel for SNe Ia pro- genitors (Chen et al. 2011), morphology of HB stars in globu- lar clusters (Lei et al. 2013), and long-period eccentric binaries with He WD (Siess et al. 2014) and SdB companions (Vos et al. 2015). The efficiency of the tidally enhanced stellar wind was set to BW = 104 by Tout & Eggleton (1988) to fit the observed pa- rameters of Z Her, but this value may vary depending on the specific system considered. The results plotted in the right panel of Fig. 10 are for a value of BW = 104, but we note that we are able to reproduce the orbital parameters of the HD 59686 system with several values of BW ranging from ∼5000−10000. A strik- ing difference with the case of a standard mass loss is that now a considerable fraction of the simulations shows eccentric orbits in the range ∼0.40−0.85 with periods of a few tens of thousand days, very similar to HD 59686 AB. The model that best repro- duces HD 59686 AB (black solid line) has a final eccentricity and period of e f = 0.724 and P f = 11555 days (with initial val- ues of ei = 0.82 and Pi = 9000 days), very close to the actual observed values of the HD 59686 system. These results show that the WD scenario for HD 59686 B is plausible, provided that its progenitor passed through an enhanced wind mass-loss phase during the AGB evolution. It is worth mentioning, however, that the previous calculations do not include the presence of a planet in the binary system. If the planet HD 59686 Ab existed before the presumed evolution of the stellar companion HD 59686 B, then the change from MS star to giant star to white dwarf could have affected the evolution of the planetary body. Regardless of the nature of the stellar object HD 59686 B, the formation of a planet with a stellar companion at 13.6 au with a periastron distance of only 3.6 au presents serious challenges to standard planet formation theories (e.g., Hatzes & Wuchterl 2005). In the core-accretion model (e.g., Mizuno 1980; Pollack 1984; Lissauer 1993), giant planets close to their host stars are expected to form beyond the snow line and then migrate inward to reach their current positions. For a mass of ∼1.9 M(cid:12) the snow line of HD 59686 A is located at ∼9.7 au (assuming the model of Ida & Lin 2005). However, with an eccentric stellar companion at 13.6 au, the protoplanetary disk around the primary star would be truncated at around 1 au or less (Pichardo et al. 2005), pre- venting the formation of a giant planet at this separation from the host star (Eggenberger et al. 2004). Similarly, a formation in situ at ∼ 1−2 au by disk instability (e.g., Kuiper 1951; Toomre 1964; Boos 2000) is highly unlikely as the required temperature for ef- ficient cooling would be too high for the protoplanetary disk to remain bound to the central star (Rafikov 2005). Additionally, giant planets are not expected to form by disk instability in bi- nary systems with separations of aB (cid:46) 20 au and eccentricities of eB (cid:38) 0.4 (Jang-Condell 2015). 6.3. Possible origin of the planet HD 59686 Ab With the increasing number of planets found in non-conventional configurations in binary systems, in both P-type and S-type or- bits, new mechanisms have been proposed to explain their ori- gin. For instance, Schleicher & Dreizler (2014) developed a model to explain circumbinary planets in the close binary NN Article number, page 12 of 14 Ser from the ejecta of common envelopes. They also extended their model to predict the masses of 12 planetary candidates around post-common envelope binaries (PCEBs) listed by Zoro- tovic & Schreiber (2013), showing a good agreement in sev- eral systems. Additionally, Perets (2010) and Tutukov & Fe- dorova (2012) have discussed the possibility of forming second- generation (SG) circumstellar planets in evolved binary systems. The main idea of SG planets is that an evolved star transfers mass to its companion, and when the binary separation is small enough, this could lead to the formation of an accretion disk around the primary star with sufficient mass to form planets. If the stellar object HD 59686 B is confirmed to be a WD, then this scenario appears as an interesting alternative to explain the origin of HD 59686 Ab. In principle, this system would satisfy several expected ob- servational characteristics from SG planets. As stated by Perets (2010), SG planets are expected to be almost exclusively found in evolved binary systems with compact objects, such as WD or neutron stars. They are also likely to be more massive than normal first-generation planets; with a mass roughly constrained between ∼7−9 MJup, HD 59686 Ab is among the most massive exoplanets detected so far. SG planets could reside in regions of orbital phase space forbidden to pre-existing planets by dy- namical arguments. HD 59686 Ab is marked as unstable or on the border of stability by some dynamical criteria (Holman & Wiegert 1999; Mardling & Aarseth 2001), although detailed N- body integrations allow stability for a certain parameter space in- cluding both prograde and retrograde orbital configurations (Tri- fonov et al. 2016, in prep.). For the prograde case, the bootstrap dynamical test yielded a small sample of long-term stable fits consistent with the bootstrap distributions at the 1 sigma confi- dence level. These prograde fits are locked in secular resonance with aligned orbital periapsis. The best dynamical fits assuming a retrograde orbit have slightly better quality (smaller χ2) and are long-term stable. It is worth noting that there is evidence sug- gesting that the planet in the ν Oct system, that is, the tight binary with aB = 2.6 au and a K-giant primary, is in a retrograde orbit (Eberle & Cuntz 2010; Gozdziewski et al. 2012; Ramm 2015). Although the SG planet scenario may seem attractive, we cannot discard the possibility that the current configuration of the HD 59686 system may be the result of past dynamical in- teractions in the native star cluster (Pfahl & Muterspaugh 2006). In this context, the planet HD 59686 Ab could have formed be- yond the snow line around its single host star, and later, through dynamical processes, another binary star may have exchanged one of its stellar members for this single star with the already formed planet. This scenario has been invoked in the past to ex- plain the origin of a giant planet in the system HD 188753 (Pfahl 2005; Portegies Zwart & McMillan 2005). However, the exis- tence of this planet was recently proved false by Eggenberger et al. (2007). Pfahl & Muterspaugh (2006) estimated that dynam- ical interaction in the parent star clusters would deposit giant planets in roughly 0.1% of binary systems with semi-major axis of a < 50 au. We note that this value was obtained under sev- eral assumptions and it is unlikely that we have detected such a system in our sample, which does not contain a large number of such binaries. Another similar, albeit slightly different possibility is that the present configuration of the HD 59686 system might have been generated in the past after the formation of the planet HD 59686 Ab was completed. In this scenario, planets can form in wide-separation binary systems that are not hostile for the planet formation process and later, through a close stellar en- counter or a perturbation induced by a former third star in the Mauricio Ortiz et al.: Precise radial velocities of giant stars system, the orbital parameters of the system may have changed to those observed today. This possibility was first suggested by Marzari & Barbieri (2007), who studied the dynamical evolu- tion of triple star systems with a primary star harboring a planet. They found that close stellar encounters or a perturbation of the original triple system may significantly change the binary orbit, leading to more eccentric and tight binaries with planets. Ad- ditionally, using numerical simulations, Martí & Beaugé (2012) studied the formation of the planet around γ Cep by stellar scat- tering and found that around ∼1−5% of fly-by encounters involv- ing planetary systems could lead to planets in close-separation binaries. Although this number is small, we cannot exclude this possibility for the formation of HD 59686 Ab. 7. Conclusions By obtaining high-precision RVs of the giant star HD 59686 A for more than 12 yr, we discovered a clear RV signature most likely caused by a massive (mp sin i = 6.92+0.18−0.24 MJup) giant planet, HD 59686 Ab, at a distance of ap = 1.0860+0.0006 −0.0007 au from its host star. Additionally, we detected the strong signal of an eccentric (eB = 0.729+0.004 −0.003) binary companion, HD 59686 B, orbiting with a semi-major axis of only aB = 13.56+0.18−0.14 au. This makes HD 59686 AB, together with ν Oct, the binary system with the closest separation of its stellar components known to harbor a giant planet. Furthermore, at periastron, the two stars are separated by just 3.6 au; a certainly hostile environment for the formation of any planet in this system. We acquired high-resolution images of HD 59686 A using LMIRCam at the LBT telescope with the aim of investigating the nature of the stellar object HD 59686 B. We could not di- rectly detect the star, mainly because the small expected angular separation ((cid:46) 0.12(cid:48)(cid:48)) from the host star poses great challenges to current PSF-subtraction techniques. It is most likely that the binary companion is a red dwarf star or a white dwarf. The bi- nary system will be at apastron in about 2025, with an expected separation of the binary pair of around ∼20 -- 21 au. With a fa- vorable orbital configuration it would be possible to detect the companion with a similar strategy as we followed in this work. Regardless of the nature of the binary companion, the ex- istence of a planet in an eccentric binary with a separation of (cid:46) 15 au is a challenge for standard planet formation theories, namely core accretion and disk instability. It is expected that massive giant planets form in massive protoplanetary disks with Md (cid:38) 10−2 M(cid:12). In the HD 59686 system, a disk would be tidally truncated at roughly ∼1 au (Pichardo et al. 2005), resulting in a disk not massive enough for the formation of giant planets (Jang-Condell 2015). Additionally, stirring by the tidal field may inhibit the growth of icy grains and planetesimals and also sta- bilize the disk against fragmentation (Nelson 2000; Thébault et al. 2004, 2006). Under these conditions, the in situ formation by disk instability is not a plausible mechanism for giant planet for- mation. However, Rafikov & Silsbee (2015) have recently shown that it is possible to form planets within ∼20 au separation bina- ries, provided that the protoplanetary disks are massive and only weakly eccentric. It would be interesting to test the validity of this model in the HD 59686 system. As a different approach to the origin of HD 59686 Ab, we discussed the possibility that this planet could have formed in a second-generation protoplanetary disk, assuming that the stel- lar object HD 59686 B is a white dwarf. We demonstrated that given the current properties of the system, this scenario is fea- sible, and discussed its implications regarding the formation of HD 59686 Ab. Altough not directly verifiable with the currently available data, the second-generation planet hypothesis is an at- tractive alternative for the origin of HD 59686 Ab as this system accounts for several observational characteristics for this type of planets (see Perets 2010). Another mechanism that may explain the origin of the planet, although unlikely and hardly verifiable, is the past exchange of stellar companions through dynamical interaction with the native star cluster. Our detailed analysis of the extensive RV data set of HD 59686 A supports the hypothesis that planets can exist in close binary systems with separations of aB (cid:46) 20 au, contrary to the theoretical expectations (Whitmire et al. 1998; Nelson 2000) and the recent observational support showing that short- separation binaries are rarely found among Kepler planet hosts (Wang et al. 2014). However, the question of how such planets may have formed remains unanswered as none of the standard theories can satisfactorily explain the origin of HD 59686 Ab. In this context, systems such as HD 59686 and ν Oct may become benchmark objects in the study of planet formation in tight bina- ries. Acknowledgements. We would like to thank the staff at the Lick Observatory for their support during the years of this project. We are also thankful to the CAT observers who assisted with this project, including Saskia Hekker, Simon Albrecht, David Bauer, Christoph Bergmann, Stanley Browne, Kelsey Clubb, Dennis Kügler, Christian Schwab, Julian Stürmer, Kirsten Vincke, and Do- minika Wylezalek. We also thank the staff of the LBT for carrying out the high-contrast image observations of this system. The LBT is an international collaboration among institutions in the United States, Italy and Germany. LBT Corporation partners are: The University of Arizona on behalf of the Arizona university system; Istituto Nazionale di Astrofisica, Italy; LBT Beteiligungs- gesellschaft, Germany, representing the Max-Planck Society, the Astrophysi- cal Institute Potsdam, and Heidelberg University; The Ohio State University, and The Research Corporation, on behalf of The University of Notre Dame, University of Minnesota and University of Virginia. This research has made use of the SIMBAD database and the VizieR catalog access tool, CDS, Stras- bourg, France. This publication made use of data products from the Two Mi- cron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Tech- nology, funded by the National Aeronautics and Space Administration and the National Science Foundation. This research made use of Astropy, a community- developed core Python package for Astronomy (Astropy Collaboration et al. 2013, http://www.astropy.org). This work used the python package astroML (Vanderplas al. 2012, https://github.com/astroML/astroML) for the calculation of the GLS periodogram. The plots in this publication were produced using Matplotlib (Hunter 2007, http://www.matplotlib.org). T. T. and M. H. L. were supported in part by the Hong Kong RGC grant HKU 17305015. M.O. thanks V. Bailey for useful discussions regarding the high-contrast images taken with LMIRCam. We are grateful to the anonymous referee for providing insightful arguments regarding alternative interpretations of RV variations. References Anderson, D. R., Collier Cameron, A., Delrez, L., et al. 2014, MNRAS, 445, 1114 Amara, A. & Quanz, S. P. 2012, MNRAS, 427, 948 Astropy Collaboration et al., 2013, A&A, 558, A33 Baines, E. K., McAlister, H. A., Brummelaar, T. A., et al. 2008a, ApJ, 680, 728 Baines, E. K., McAlister, H. A., Brummelaar, T.A., et al. 2008b, ApJ, 682, 577 Bailey, V., Meshkat, T., Reiter, M., et al. 2014, ApJ, 780, L4 Beers, T. C., & Christlieb, N. 2005, ARA&A, 43, 531 Biazzo, K., Frasca, A., Catalano, S., et al. 2006, MSAIS, 9, 220 Brandt, T. D., McElwain, M. W., Turner, E. L., et al. 2013, ApJ, 764, 183 Bonaci´c, M., Glebbeek, E., & Pols, O. R. 2008, A&A, 480, 797 Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977 Boss, A. P. 2000, ApJ, 536, 101 Buchhave, L. A., Latham, D. W., Carter, J. A., et al. 2011, ApJS, 197, 3 Butler, R. P., Marcy, G.W., Williams, E., et al. 1996, PASP, 108, 500 Butler, R. P., Tinney, C. G., Marcy, G. W., et al. 2001, ApJ, 555, 410 Candès, E. J., Li, X., Ma, Y., et al. 2009, CoRR, abs/0912.3599 Carlberg, J. K., Cunha, K, Smith, V. V., et al. 2012, ApJ, 757, 109 Chaplin, W. J., Houdek, G., Karoff, C., et al. 2009, A&A, 500, L21 Chen, X., Han, Z. & Tout, C. A. 2011, ApJL, 735, L31 Article number, page 13 of 14 A&A proofs: manuscript no. HD59686_rev Christensen-Dalsgaard, J. 2012, ASP Conf. Ser., 462, 503 Correia, A. C. M., Udry, S., Mayor, M., et al. 2008, A&A, 479, 271 De Ridder, J., Barban, C., Baudin, F., et al. 2009, Nature, 459, 398 Desidera, S. & Barbieri, M. 2010, A&A, 462, 345 Döllinger, M. P., Hatzes, A. P., Pasquini, L., et al. 2009, A&A, 505, 1311 Doyle, L. R., Carter, J. A., Fabrycky, D. C., et al. 2011, Science, 333, 1602 Dumusque, X., Pepe, F., Lovis, C., et al. 2012, Nature, 491, 207 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Eberle J. & Cuntz M. 2010, ApJ, 721, L168 Eggenberger, P., Udry, S., & Mayor, M. 2004, A&A, 417, 353 Eggenberger, P., Udry, S., Mazeh, T., et al. 2007, A&A, 466, 1179 ESA 1997, The Hipparcos and Tycho Catalogues, ESA SP, 1200 Frink, S., Quirrenbach, A., Fischer, D., et al. 2001, PASP, 113, 173 Frink, S., Mitchell, D. S., Quirrenbach, A., et al. 2002, ApJ, 576, 478 Girardi, L., Bressan, A., Bertelli, G., et al. 2000, A&A, 141, 371 Gomez Gonzalez, C. A., Wertz, O., Christiaens, V., et al. 2016a, VIP, Astro- physics Source Code Library, record ascl:1603.003 Gomez Gonzalez, C. A., Absil, O., Absil, P.-A., et al. 2016b, arXiv:1602.08381 Gould, A., Udalski, A., Shin, I.-G., et al. 2014, Science, 345, 46 Go´zdziewski, K., Słonina, M., Migaszewski, C., et al. 2012, MNRAS, 430, 533 Haghighipour, N. 2009, arXiv:0908.3328 Han, I., Lee, B. C., Kim, K. M., et al. 2010, A&A, 509, A24 Hatzes, A. P. & Cochran, W. D., 2000, AJ, 120, 979 Hatzes, A. P. 2002, Astron. Nachr., 323, 392 Hatzes, A. P., Cochran, W. D., Endl, M., et al. 2003, ApJ, 599, 1383 Hatzes, A. P., Setiawan, J., Pasquini, L., et al. 2004, ESA pub., Vol. 538 Hatzes, A. P., & Wuchterl, G. 2005, Nature, 436, 182 Hatzes A. P., 2013, ApJ, 770, 133 Hatzes, A. P., Cochran, W. D., Endl, M., et al. 2015, A&A, 580, A31 Hekker, S., Reffert, S., Quirrenbach, A., et al. 2006, A&A, 454, 943 Hekker, S. & Meléndez, J., 2007, A&A, 475, 1003 Hekker, S., Snellen, I. A. G., Aerts, C., et al. 2008, A&A, 480, 215 Hinz, P. M., Bippert-Plymate, T., Breuninger, A., et al. 2012, in SPIE Conf. Ser., Vol. 8446 Holman, M. J. & Wiegert, P. A. 1999, AJ, 117, 621 Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, ApJ, 721, 1467 Huber, J. P., Bedding, T. R. & O'Toole, S. J. 2003, in Asteroseismology across the HR Diagram, ed. M. J. Thompson, Volume 284, 421 Huber, D., Bedding, T. R., Stello, D., et al. 2010, ApJ, 723, 1607 Hunter, J. D. 2007, CSE, 9, 90 Hurley, J. R., Tout, C. A., & Pols, O. R. 2002, MNRAS, 329, 897 Ida, S., & Lin, D. N. C. 2005, ApJ, 626, 1045 Jang-Condell, H. 2015, ApJ, 799, 147 Jeffries, R. D. 1997, MNRAS, 288, 585 Johnson, J. A., Marcy, G. W., Fischer, D. A., et al. 2006, ApJ, 652, 1724 Johnson, J. A., Bowler, B., Howard, A., et al. 2010, ApJ, 721, 153 Johnson, J. A., Payne, M., Howard, A., et al. 2011, AJ, 141, 16 Jones, M. I., Jenkins, J.S., Rojo, P., et al. 2015a A&A, 573, 3 Jones, M. I., Jenkins, J.S., Rojo, P., et al. 2015b A&A, 580, 14 Jorissen, A., Van Eck, S., Van Winckel, H., et al. 2016a, A&A, 586, A158 Jorissen, A., Hansen, T., Van Eck, S., et al. 2016b, A&A, 586, A159 Kalirai, J. S., Saul Davis, D., Richer, H. B., et al. 2009, ApJ, 705, 408 Karakas, A. I., Tout, C. A. & Lattanzio, J. C. 2000, MNRAS, 316, 689 Käufl, H. U., Ballester, P., Biereichel, P., et al. 2004, SPIE, Vol. 5492 Kjeldsen, H. & Bedding, T. R., 1995, A&A, 293, 87 Kjeldsen, H. & Bedding, T. R. 2011, A&A, 529, L8 Kuiper, G. P. 1951, PNAS, 37, 1 Kunitomo, M., Ikoma, M., Sato, et al. 2011, ApJ, 737, 66 Kürster, M., Endl, M., Rouesnel, F., et al. 2003, A&A, 403, 1077 Lee, B.-C., Park, M.-G., Lee, S.-M., et al. 2015, arXiv:1509.09012 Lei, Z.-X., Chen, X.-F., Zhang, F.-H., et al. 2013, A&A, 549, 145 Leisenring, J. M., Skrutskie, M. F., Hinz, P. M., et al. 2012, in SPIE Conf. Ser, Vol. 8446 Lissauer, J. J. 1993, ARA&A, 31, 129 Lissauer, J., Marcy, G., Bryson, S., et al. 2014, ApJ, 784, 44 Lafrenière, D., Doyon, R., Marois, C., et al. 2007, ApJ, 670, 1367 Liu, Y. J., Sato, B., Zhao, G., et al. 2008, ApJ, 672, 553 Maire, A., Skemer, A., Hinz, P., et al 2015, A&A, 576, 133 Massarotti, A., Latham, D. W., Stefanik, R. P., et al. 2008, ApJ, 135, 209 Mardling R. A. & Aarseth S. J. 2001, MNRAS, 321, 398 Marois, C., Lafrenière, D., Doyon, R., et al. 2006, ApJ, 641, 556 Martí, J. G. & Beaugé, C. 2012, A&A, 544, 97 Marzari, F. & Barbieri, M. 2007, A&A, 467, 347 Masseron, T., Johnson, J. A., Plez, B., et al. 2010, A&A, 509, A93 Mawet, D., Milli, J., Wahhaj, Z., et al. 2014, ApJ, 792, 97 Mayor, M. & Queloz, D. 1995, Nature, 378, 355 Mayor, M., Udry, S., Naef, D., et al. 2004, 415, 391 Merand, A., Borde, P. & Coudé du Foresto, V. 2004, SPIE, Vol. 5491 Meshkat, T., Kenworthy, M. A., Quanz, S. P., et al. 2014, ApJ, 780, 17 Mitchell, D. S., Reffert, S., Trifonov, T., et al. 2013, A&A, 555, 87 Mizuno, H. 1980, PThPh, 64, 544 Article number, page 14 of 14 Morais, M. H. M. & Correia, A. C. M. 2008, A&A, 491, 899 Morais, M. H. M. & Correia, A. C. M. 2011, A&A, 525, 152 Morton, T. D., Bryson, S. T., Coughlin, J. L., et al. 2016, ApJ, 822, 86 Nelson, A. F. 2000, ApJ, 537, L65 Niedzielski, A., Wolszczan, A., Nowak, G., et al., 2015, ApJ, 803, 1 O'Neal, D. 2006, ApJ, 645, 659 Orosz, J. A., Welsh, W. F., Carter, J. A., et al. 2012, ApJ, 758, 87 Orosz, J. A., Welsh, W. F., Carter, J. A., et al. 2012, Science, 337, 1511 Percy, J. R., Wilson, J. B. & Henry, G. W. 2001, PASP, 113, 983 Percy, J. R., Mashintsova, M., Nasui, C. O., et al. 2008, PASP, 120, 523 Perets, H. B. 2010, arXiv:1001.0581 Pfahl, E. 2005, ApJ, 635, L89 Pfahl, E. & Muterspaugh, M. 2006, ApJ, 652, 1694 Pichardo, B., Sparke, L. S. & Aguilar, L. A. 2005, MNRAS, 359, 521 Placco, V. M., Frebel, A., Beers, T. C., et al. 2014, ApJ, 797, 21 Pojma´nski, G. 1997, Acta Astron., 47, 467 Pojma´nski, G. 2001, in Small Telescope Astronomy on Global Scales, eds. B. Paczynski, W.-P. Chen, & C. Lemme, IAU Colloq. 183, ASP Conf. Ser., 246, 53 Pollack, J. B. 1984, ARA&A, 22, 389 Portegies Zwart, S. F. & McMillan, S. L. 2005, ApJ, 633, L141 Press, W. H., Teukolsky, S. A., Vetterling, W. T., et al. 1992, Numerical recipes in FORTRAN. The art of scientific computing (Cambridge University Press) Queloz, D., Mayor, M., Weber, L., 2000, A&A, 354, 99 Quintana, E. V., Adams, F. C., Lissauer, J. J., et al. 2007, ApJ, 660, 807 Quirrenbach, A., Reffert, S. & Bergmann, C. 2011, AIPCS, Vol. 1331 Rafikov, R. R. 2005, ApJL, 621, L69 Rafikov, R. & Silsbee, K. 2015, ApJ, 798, 70 Raghavan, D., McAlister, H. A., Henry, T. J., et al. 2010, ApJS, 190, 1 Rajpaul, V., Aigrain, S., & Roberts, S. 2016, MNRAS, 456, L6 Ramm D. J., Pourbaix D., Hearnshaw J. B., et al. 2009, MNRAS, 394, 1695 Ramm D. J. 2015, MNRAS, 449, 4428 Reffert, S., Quirrenbach, A., Mitchell, D. S., et al. 2006, ApJ, 652, 661 Reffert, S., Bergmann, C., Quirrenbach, A., et al. 2015, A&A, 574, 116 Reimers, D. 1975, Mém. Soc. Roy. Sci. Liège, 8, 369 Roberts, J. L. C., Turner, N. H., ten Brummelaar, T. A., et al. 2011, AJ, 142, 175 Roell, T., Neuhüser, R., Seifahrt, A., et al. 2012, A&A, 542, A92 Rowe, J., Bryson, S., Marcy, G., et al. 2014, ApJ, 784, 45 Santerne, A., Hébrard, G., Deleui, M., et al. 2014, A&A, 571, 37 Sato, B., Omiya, M., Harakawa, H., et al. 2013, PASJ, 65, 85 Siess, L., Davis, P. J. & Jorissen, A. 2014, A&A, 565, 57 Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163 Skrutskie, M. F., Jones, T., Hinz, P., et al. 2010, in SPIE Conf. Ser., Vol. 7735 Schneider, J. & Cabrera, J. 2006, A&A, 445, 1159 Schneider, J., Dedieu, C., Le Sidaner, P., et al. 2011, A&A, 532, A79 Schleicher, D. & Dreizler, S. 2014, A&A, 563, 61 Soummer, R., Pueyo, L. & Larkin, J. 2012, ApJL, 755, L28 Stello, D., Huber, D., Bedding, T. R., et al. 2013, ApJ, 765, 41 Thébault, P., Marzari, F., Scholl, H., et al. 2004, A&A, 427, 1097 Thébault, P., Marzari, F. & Scholl, H., et al. 2006, Icarus, 183, 193 ten Brummelaar, T. A., McAlister, H. A., Ridgway, S. T., et al. 2005, ApJ, 628, Toomre A., 1964, ApJ, 139, 1217 Tout, C. A., & Eggleton, P. P. 1988, MNRAS, 231, 823 Trifonov, T., Reffert, S., Tan, X., et al. 2014, A&A, 568, 64 Trifonov, T., Reffert, S., Zechmeister, M., et al. 2015, A&A, 582, 54 Tutukov, A. V. & Fedorova, A. V. 2012, Astron. Rep., 56, 305 Vanderplas, J. T., Connolly, A. J., Ivezi´c, Ž., et al. 2012, proc. of CIDU, pp. 47-54 van Leeuwen, F. 2007, A&A, 474, 653 Villaver, E. & Livio, M., 2009, ApJ, 705, 81 Villaver, E., Livio, M., Mustill, A., 2014, ApJ, 794, 3 Vogt, S. S. 1987, PASP, 99, 1214 Vos, J., Østensen, R. H., Marchant, P., et al. 2015, A&A, 579, 49 Wang, J., Xie, J.-W., Barclay, T., & Fischer, D. 2014, ApJ, 783, 4 Weidemann, V. 1977, A&A, 59, 411 Weidemann, V. 1987, A&A, 188, 74 Weidemann, V. 1990, ARA&A, 28, 103 Welsh, W. F., Orosz, J. A., Carter, J. A., et al. 2012, Nature, 481, 475 Whitmire, D. P., Matese, J. L., Criswell, L.. et al. 1998, Icarus, 132, 196 Wittenmyer, R. A., Butler, R. P., Wang, L., et al. 2016, MNRAS, 455, 1398 Zechmeister, M. & Kürster, M. 2009, A&A, 496, 577 Zorotovic, M., & Schreiber, M. R. 2013, A&A, 549, A95 Zucker, S., Mazeh, T., Santos, N. C., et al. 2004, A&A, 426, 695 453
1512.02559
2
1512
2015-12-18T01:46:53
Planet Hunters. VIII. Characterization of 41 Long-Period Exoplanet Candidates from Kepler Archival Data
[ "astro-ph.EP" ]
The census of exoplanets is incomplete for orbital distances larger than 1 AU. Here, we present 41 long-period planet candidates in 38 systems identified by Planet Hunters based on Kepler archival data (Q0-Q17). Among them, 17 exhibit only one transit, 14 have two visible transits and 10 have more than three visible transits. For planet candidates with only one visible transit, we estimate their orbital periods based on transit duration and host star properties. The majority of the planet candidates in this work (75%) have orbital periods that correspond to distances of 1-3 AU from their host stars. We conduct follow-up imaging and spectroscopic observations to validate and characterize planet host stars. In total, we obtain adaptive optics images for 33 stars to search for possible blending sources. Six stars have stellar companions within 4". We obtain high-resolution spectra for 6 stars to determine their physical properties. Stellar properties for other stars are obtained from the NASA Exoplanet Archive and the Kepler Stellar Catalog by Huber et al. (2014). We validate 7 planet candidates that have planet confidence over 0.997 (3-{\sigma} level). These validated planets include 3 single-transit planets (KIC-3558849b, KIC-5951458b, and KIC-8540376c), 3 planets with double transits (KIC-8540376b, KIC-9663113b, and KIC-10525077b), and 1 planet with 4 transits (KIC-5437945b). This work provides assessment regarding the existence of planets at wide separations and the associated false positive rate for transiting observation (17%-33%). More than half of the long-period planets with at least three transits in this paper exhibit transit timing variations up to 41 hours, which suggest additional components that dynamically interact with the transiting planet candidates. The nature of these components can be determined by follow-up radial velocity and transit observations.
astro-ph.EP
astro-ph
Draft version December 21, 2015 Preprint typeset using LATEX style emulateapj v. 08/22/09 PLANET HUNTERS. VIII. CHARACTERIZATION OF 41 LONG-PERIOD EXOPLANET CANDIDATES FROM KEPLER ARCHIVAL DATA Ji Wang1,2, Debra A. Fischer1, Thomas Barclay3,4, Alyssa Picard1, Bo Ma5, Brendan P. Bowler2,6, Joseph R. Schmitt1, Tabetha S. Boyajian1, Kian J. Jek7, Daryll LaCourse7, Christoph Baranec8, Reed Riddle2, Nicholas M. Law9, Chris Lintott10, Kevin Schawinski11, Dean Joseph Simister7, Boscher Gr´egoire7, Sean P. Babin7, Trevor Poile7, Thomas Lee Jacobs7, Tony Jebson7, Mark R. Omohundro7, Hans Martin Schwengeler7, Johann Sejpka7, Ivan A.Terentev7, Robert Gagliano7, Jari-Pekka Paakkonen7, Hans Kristian Otnes Berge7, Troy Winarski7, Gerald R. Green7, Allan R. Schmitt7 Martti H. Kristiansen7 Abe Hoekstra7 (Received; Accepted) Draft version December 21, 2015 ABSTRACT The census of exoplanets is incomplete for orbital distances larger than 1 AU. Here, we present 41 long-period planet candidates in 38 systems identified by Planet Hunters based on Kepler archival data (Q0-Q17). Among them, 17 exhibit only one transit, 14 have two visible transits and 10 have more than three visible transits. For planet candidates with only one visible transit, we estimate their orbital periods based on transit duration and host star properties. The majority of the planet candidates in this work (75%) have orbital periods that correspond to distances of 1-3 AU from their host stars. We conduct follow-up imaging and spectroscopic observations to validate and characterize planet host stars. In total, we obtain adaptive optics images for 33 stars to search for possible blending sources. Six stars have stellar companions within 4′′. We obtain high-resolution spectra for 6 stars to determine their physical properties. Stellar properties for other stars are obtained from the NASA Exoplanet Archive and the Kepler Stellar Catalog by Huber et al. (2014). We validate 7 planet candidates that have planet confidence over 0.997 (3-σ level). These validated planets include 3 single-transit planets (KIC-3558849b, KIC-5951458b, and KIC-8540376c), 3 planets with double transits (KIC-8540376b, KIC-9663113b, and KIC- 10525077b), and 1 planet with 4 transits (KIC-5437945b). This work provides assessment regarding the existence of planets at wide separations and the associated false positive rate for transiting observation (17%-33%). More than half of the long-period planets with at least three transits in this paper exhibit transit timing variations up to 41 hours, which suggest additional components that dynamically interact with the transiting planet candidates. The nature of these components can be determined by follow-up radial velocity and transit observations. Subject headings: Planets and satellites: detection - surveys 0 This publication has been made possible by the partici- pation of more than 200,000 volunteers in the Planet Hunters project. Their contributions are individually acknowledged at http://www.planethunters.org/#/acknowledgements 1 Department of Astronomy, Yale University, New Haven, CT 06511, USA 2 California Institute of Technology, 1200 East California Boulevard, Pasadena, CA 91101, USA 3 NASA Ames Research Center, M/S 244-30, Moffett Field, CA 94035, USA 4 Bay Area Environmental Research Institute, Inc., 560 Third Street West, Sonoma, CA 95476, USA 5 Department of Astronomy, University of Florida, 211 Bryant Space Science Center, Gainesville, FL 32611-2055, USA 6 Department of Astronomy, University of Texas at Austin, 2515 Speedway, Stop C1400, Austin, TX 78712 7 Planet Hunter 8 Institute for Astronomy, University of Hawai‘i at M¯anoa, Hilo, HI 96720-2700, USA 9 Department of Physics and Astronomy, University of North Carolina at Chapel Hill, Chapel Hill, NC 27599-3255, USA 10 Oxford Astrophysics, Denys Wilkinson Building, Keble Road, Oxford OX1 3RH 11 Institute for Astronomy, Department of Physics, ETH Zurich, Wolfgang-Pauli-Strasse 27, CH-8093 Zurich, Switzer- 1. INTRODUCTION Since its launch in March of 2009, the NASA Ke- pler mission has been monitoring ∼160,000 stars in order to detect transiting extrasolar planets with high relative photometric precision (∼20 ppm in 6.5 h, Jenkins et al. 2010). In May 2013, the Kepler main mission ended with the failure of a second reaction wheel; however, the first four years of Kepler data have led to a wealth of planetary discoveries with a total of 4,706 an- nounced planet candidates13 (Borucki et al. 2010, 2011; Batalha et al. 2013; Burke et al. 2014). The confirmed and candidate exoplanets typically have orbital periods shorter than 1000 days because at least three detected transits are needed for iden- tification by the automated Transit Planet Search algorithm. Therefore, transiting exoplanets with periods longer than ∼1000 days are easily missed. land 13 http://exoplanetarchive.ipac.caltech.edu/ as of Nov 11 2015 2 Planet Hunters VIII The detection of short-period planets is further fa- vored because the transit probability decreases lin- early with increasing orbital distance. For these reasons, estimates of the statistical occurrence rate of exoplanets tend to focus on orbital periods shorter than a few hundred days (e.g., Fressin et al. 2013; Petigura et al. 2013; Dong & Zhu 2013). Ra- dial velocity (RV) techniques also favor the de- tection of shorter period orbits. While gas giant planets have been discovered with orbital periods longer than a decade, their smaller reflex veloc- ity restricts detection of sub-Neptune mass plan- ets to orbital radii less than ∼1 AU (Lovis et al. 2011). In principle, astrometric observations fa- vor longer period orbits; however, high precision needs to be maintained over the correspondingly longer time baselines. For shorter periods, the plan- ets need to be massive enough to introduce a de- tectable astrometric wobble in the star and Gaia should begin to contribute here (Perryman et al. 2001). Microlensing offers sensitivity to planets in wider orbits and has contributed to our statis- tical knowledge about occurrence rates of longer period planets (e.g., Gaudi 2010; Cassan et al. 2012) and direct imaging of planets in wide orbits is also beginning to contribute important informa- tion (Oppenheimer & Hinkley 2009). Here, we announce 41 long-period transiting ex- oplanet candidates from the Kepler mission. These planet candidates mostly have 1-3 visible tran- sits and typically have orbital periods between 100 and 2000 days, corresponding to orbital separations from their host stars of 1-3 AU. The candidate sys- tems were identified by citizen scientists taking part in the Planet Hunters project14. We obtain follow- up adaptive optics (AO) images (for 33 host stars) and spectroscopic observations (for 6 host stars) in an effort to validate the planet candidates and char- acterize their host stars. We derive their orbital and stellar parameters by fitting transiting light curves and performing spectral classification. The Planet Hunters project began in Decem- ber 2010 as part of the Zooniverse15 network of Citizen Science Projects. The project displays light curves from the Kepler mission to crowd- source the task of identifying transits (Fischer et al. 2012). This method is effective in finding poten- tial exoplanets not flagged by the Kepler data re- duction pipeline, since human classifiers can of- ten spot patterns in data that would otherwise confuse computer algorithms. The detection ef- ficiency of the volunteers is independent of the number of transits present in the light curve, i.e., they are as likely to identify a single transit as multiple transits in the same lightcurve, however the probability of identifying planets is higher if the transit is deeper. Schwamb et al. (2012b) de- scribed the weighting scheme for transit classifi- cations. Wang et al. (2013) and Schmitt et al. (2014a) described the process of vetting planet can- didates in detail as well as the available tools on the 14 http://www.planethunters.org/ 15 https://www.zooniverse.org Planet Hunters website. The paper is organized as follows. In §2, we model transiting light curves of planet candidates and de- rive stellar and orbital properties of these candidate systems. In §3, we present adaptive optics (AO) imaging for 33 systems and spectroscopic observa- tions for 6 systems. In §4, we calculate planet confi- dence for each planet candidate and discuss notable candidate systems. Finally, we conclude in §5 with a summary and discussions of future prospects. 2. PLANET CANDIDATES AND THEIR HOST STARS Planet Hunters identified 41 long-period planet candidates around 38 stars. In this section, we de- scribe the procedures with which we modeled these transit curves and estimated the stellar properties of their host stars. Since 17 planet candidates ex- hibit only one visible transit, their orbital periods can not be well-determined. We provide a method of constraining the orbital period for a single-transit event based on transit duration and host star prop- erties. 2.1. Modeling Light Curves We downloaded the Kepler light curves from the Mikulski Archive for Space Telescopes (MAST16) and detrended the quarterly segments using the autoKep software in the Transit Analysis Pack- age (TAP, Gazak et al. 2012). The light curves were then modeled using TAP which adopts an analytic form for the model described by Mandel & Agol (2002). The free parameters in the model include orbital period, eccentricity, argument of periastron, inclination, the ratio of semi-major axis and stel- lar radius a/R∗, the planet-star radius ratio Rp/R∗, mid transit time, linear and quadratic limb darken- ing parameters. We are particularly interested in Rp/R∗ and a/R∗. The former is used to determine the planet radius. The latter helps to estimate the orbital periods for planet candidates with only one visible transit. The following equation of constrain- ing orbital period is derived based on Equation 18 and 19 from Winn (2010): P 1 yr = (cid:18) T 13 hr(cid:19)3 ·(cid:18) ρ ρ⊙(cid:19) · (1 − b2)− 3 2 , (1) where P is period, T is the transit duration, i.e., the interval between the halfway points of ingress and egress, ρ is stellar density, ρ⊙ is the solar density, and b is the impact parameter. In a transit observa- tion, the transit duration, T , is an observable that can be parametrized the follow way: T = 1 π · P ·(cid:18) a R∗(cid:19)−1 ·p(1 − b2) · √1 − e2 1 + e sin ω , (2) where e is orbital eccentricity and ω is the argument of periastron. Most of planet candidates in this paper have or- bital periods between 100 and 2000 days, and some 16 http://archive.stsci.edu Long-Period Exoplanets 3 of these are likely to be in eccentric orbits. Ec- centricity affects the transit duration. For exam- ple, the transiting duration of a planet on an ec- centric orbit can be longer than that for a circular orbit if viewed from the time of apastron. Unfor- tunately, it is very difficult to know whether long transit durations are caused by long orbital periods or high eccentricities, especially if the stellar radius is uncertain. However, since 80% of known planets with orbital periods longer than 100 days have ec- centricities lower than 0.317, we adopt a simplified prior assumption of zero eccentricity in our models. This feeds into our estimates for orbital periods of those systems with only one transit, however the effect is not large. The main uncertainty for the planet period estimation comes from uncertainties in the stellar radius. For example, a typical 40% stellar radius error translates to a ∼40% a/R∗ er- ror. Given that the observable T stays the same, the 40% stellar radius error leads to 40% period estimation error according to Equation 2. In com- parison, floating the eccentricity between 0 and 0.3 typically changes P by 20%. Therefore, the effect of eccentricity is smaller than the effect of stellar ra- dius error on period estimation. Furthermore, set- ting eccentricity to zero reduces the number of free parameters by two, i.e., eccentricity and argument of periastron; this facilitates the convergence of the Markov chains in TAP analysis. This is especially useful when there are only 1-4 transits available to constrain the model. The posterior distribution of the MCMC analysis is used to contain the orbital period (§2.3) for systems that only have a single transit. We report results of light curve modeling for sys- tems with only one observed transit (Table 1, shown in Fig. 1 and Fig. 2), two transits (Table 2, shown in Fig. 3), and three transits (Table 3, shown in Fig. 4). 2.2. Stellar Mass and Radius Characterizing host stars for planetary systems helps us to better understand the transiting planets. In particular, the planet radius can be calculated only if stellar radius is estimated. Stellar density is required for estimating the orbital periods for those planets that exhibit only one transit (see Equation 1). We estimate stellar mass and radius in a similar way as Wang et al. (2014): we infer these two stellar properties using the Yale-Yonsei Isochrone interpo- lator (Demarque et al. 2004). The inputs for the interpolator are Teff, log g, [Fe/H], α element abun- dance [α/H] and stellar age. The first three param- eters can be obtained by analyzing follow-up stel- lar spectra or from the NASA Exoplanet Archive18 and the updated Kepler catalog for stellar proper- ties (Huber et al. 2014). We set [α/H] to be the solar value, zero, and allow stellar age to vary be- tween 0.08 and 15 Gyr. We ran a Monte Carlo simulation to consider measurement uncertainties of Teff , log g, [Fe/H]. For stars with spectroscopic 17 http://exoplanets.org/ 18 http://exoplanetarchive.ipac.caltech.edu follow-up observations (§3.2), the uncertainties are based on the MOOG spectroscopic analysis (Sneden 1973). For stars that are Kepler Objects of Interest (KOIs), the uncertainties are from the NASA Exo- planet Archive. We report the 1σ ranges for stellar masses and radii in Table 4 along with Teff , log g, and [Fe/H]. 2.3. Orbital Period Orbital periods are a fundamental parameter for exoplanets and are often used to understand the prospects for habitability. For systems with more than one visible transit, we determined the or- bital period by calculating the time interval between transits. The uncertainty of the orbital period is calculated by propagating the measurement error of the mid transit time of each transit. For systems with only one visible transit, we use Equation 1 to estimate the orbital period P , as a function of the transit duration T , stellar density ρ, and the im- pact parameter b. T and b can be constrained by modeling the transiting light curve. For instance, T can be measured directly from the transit observa- tion, and b can be inferred by fitting the light curve. On the other hand, ρ can be constrained by stellar evolutionary model as described in §2.2. Therefore, with knowledge of T , ρ, and b from transit obser- vation and stellar evolutionary model, we can con- strain the orbital period for planet candidates with only a single transit. We start with a test TAP run to obtain the pos- terior distribution of the transit duration T (Equa- tion 2) and impact parameter b. The distribution of stellar density can be obtained from the process as described in §2.2. We then start a Monte Carlo simulation to infer the distribution of orbital period. In the simulation, we sample from T , b and ρ dis- tributions, which result in a distribution of orbital period. We report the mode and 1-σ range of orbital period in Table 1. We investigate the error of our period estimation using systems with known orbital periods. For the 24 planet candidates with 2-4 transits in this paper, we compare the period ( ¯P ) estimated from individ- ual transit and the period (P ) based on the interval between mid-transit, which is much more precise than ¯P . If ¯P and P are in agreement within 1-σ error bars, then the method used for single-transit systems would seem to give a reasonable estimate and uncertainty for orbital period. The left panel of Fig. 5 shows the distribution of the difference between ¯P and P normalized by measurement un- certainty δP , which is calculated as half of the 1-σ range from the Monte Carlo simulation. About 69% of the comparisons are within 1-σ range, which in- dicates that ¯P and P agree for the majority of cases and δP is a reasonable estimation of measurement uncertainty. The right panel of Fig. 5 shows the fractional error (δP /P ) distribution of the orbital periods estimated from individual transit. The me- dian fractional error is 1.4 and the fractional error is smaller than 50% for 34% of cases, which suggests that the period estimated from an individual transit has a large uncertainty, i.e., hundreds of days. This 4 Planet Hunters VIII is because of the weak dependence of transit dura- tion on orbital period, i.e., T ∼ P 1/3, a large range of P would be consistent with the measured transit duration. As a result, orbital period uncertainty for systems with a single transit is much larger than systems with more than one visible transit. How- ever, the estimation of orbital period provides a time window for follow-up observations. 3. FOLLOW-UP OBSERVATIONS Follow-up observations include AO imaging and spectroscopy of host stars with planet candi- dates. AO imaging can identify additional stel- lar components in the system or in the fore- ground/background. These can be potential sources for flux contamination (e.g., Dressing et al. 2014) or false positives (e.g., Torres et al. 2011). Spec- troscopic follow-up observations are used to derive stellar properties that are more reliable than those derived with multi-band photometry. Furthermore, since follow-up observations exclude some scenarios for false positives, the likelihood of a planet candi- date being a bona-fide planet can be increased and a planet candidate can be statistically validated (e.g., Barclay et al. 2013). In this section, we describe our AO imaging and spectroscopic follow-up obser- vations. In addition, we discuss sources from which we obtain archival data and information about these planet host stars. 3.1. Adaptive Optics Observations In total, AO images were taken for 33 stars with planet candidates in this paper. We observed 30 tar- gets with the NIRC2 instrument (Wizinowich et al. 2000) at the Keck II telescope using the Natural Guiding Star mode. The observations were made on UT July 18th and August 18th in 2014, and Au- gust 27-28 in 2015 with excellent/good seeing be- tween 0.3′′ to 0.9′′. NIRC2 is a near infrared im- ager designed for the Keck AO system. We selected the narrow camera mode, which has a pixel scale of 9.952 mas pixel−1 (Yelda et al. 2010). The field of view (FOV) is thus ∼10′′×10′′ for a mosaic 1K ×1K detector. We started the observation in the Ks band for each target. The exposure time was set such that the peak flux of the target is at most 10,000 ADU for each frame, which is within the lin- ear range of the detector. We used a 3-point dither pattern with a throw of 2.5′′. We avoided the lower left quadrant in the dither pattern because it has a much higher instrumental noise than other 3 quad- rants on the detector. We continued observations of a target in J and H bands if any stellar companions were found. We observed 1 target with the PHARO instrument(Brandl et al. 1997; Hayward et al. 2001) at the Palomar 200-inch telescope. The observation was made on UT July 13rd 2014 with seeing varying between 1.0′′ and 2.5′′. PHARO is behind the Palomar-3000 AO system, which provides an on-sky Strehl of up to 86% in K band (Burruss et al. 2014). The pixel scale of PHARO is 25 mas pixel−1. With a mosaic 1K ×1K detector, the FOV is 25′′×25′′. We normally obtained the first image in the Ks band with a 5-point dither pattern, which had a throw of 2.5′′. The exposure time setting criterion is the same as the Keck observation: we ensured that the peak flux is at least 10,000 ADU for each frame. If a stellar companion was detected, we observed the target in J and H bands. We observed 11 targets between UT 2014 Aug 23rd and 30th with the Robo-AO system installed on the 60-inch telescope at Palomar Observatory (Baranec et al. 2013, 2014). Observations consisted of a sequence of rapid frame-transfer read-outs of an electron multiplying CCD camera with 0.′′043 pix- els at 8.6 frames per second with a total integra- tion time of 90 s in a long-pass filter cutting on at 600nm. The images were reduced using the pipeline described in Law et al. In short, after dark subtraction and flat-fielding using daytime cal- ibrations, the individual images were up-sampled, and then shifted and aligned by cross-correlating with a diffraction-limited PSF. The aligned images were then co-added together using the Drizzle al- gorithm (Fruchter & Hook 2002) to form a single output frame. The final “drizzled” images have a finer pixel scale of 0.′′02177/pixel. (2014). The raw data from NIRC2 and PHARO were pro- cessed using standard techniques to replace bad pix- els, flat-field, subtract thermal background, align and co-add frames. We calculated the 5-σ detection limit as follows. We defined a series of concentric annuli centering on the star. For the concentric an- nuli, we calculated the median and the standard deviation of flux for pixels within these annuli. We used the value of five times the standard deviation above the median as the 5-σ detection limit. We report the detection limit for each target in Table 5. Detected companions are reported in Table 6. 3.2. Spectroscopic Observation We obtained stellar spectra for 6 stars using the East Arm Echelle (EAE) spectrograph at the Palo- mar 200-inch telescope. The EAE spectrograph has a spectral resolution of ∼30,000 and covers the wavelength range between 3800 to 8600 A. The ob- servations were made between UT Aug 15th and 21st 2014. The exposure time per frame is typically 30 minutes. We usually obtained 2-3 frames per star and bracketed each frame with Th-Ar lamp ob- servations for wavelength calibration. Because these stars are faint with Kepler magnitudes mostly rang- ing from 13 to 15.5 mag, the signal to noise ratio (SNR) of their spectra is typically 20-50 per pixel at 5500 A. We used IDL to reduce the spectroscopic data to get wavelength calibrated, 1-d, normalized spectra. These spectra were then analyzed by the newest ver- sion of MOOG (Sneden 1973) to derive stellar prop- erties such as effective temperature (Teff ), surface gravity (log g) and metallicity [Fe/H] (Santos et al. 2004). The iron line list used here was obtained from Sousa et al. (2008) excluding all the blended lines in our spectra due to a limited spectral reso- Long-Period Exoplanets 5 lution. The measurement of the equivalent widths was done systematically by fitting a gaussian pro- file to the iron lines. The equivalent widths together with a grid of Kurucz Atlas 9 plane-parallel model atmospheres (Kurucz 1993) were used by MOOG to calculate the ion abundances. The errors of the stellar parameters are estimated using the method described by Gonzalez & Vanture (1998). The tar- gets with spectroscopic follow-up observations are indicated in Table 4. 3.3. Archival AO and Spectroscopic Data From CFOP For those targets for which we did not con- duct follow-up observations, we searched the Ke- pler Community Follow-up Observation Program19 (CFOP) for archival AO and spectroscopic data. We found that only one target had AO images from CFOP. KIC-5857656 was observed at the Large Binocular Telescope on UT Oct 3rd 2014, but the image data was not available. A total of 14 of tar- gets had spectroscopic data based on CFOP, but only 8 of them had uploaded stellar spectra. We used the spectroscopically-derived stellar properties for these stars in the subsequent analyses. 3.4. Stars without AO and Spectroscopic Data For stars without AO and spectroscopic data, we obtained their stellar properties from the NASA Exoplanet Archive if they were identified as Ke- pler Objects of Interest (KOIs). If the stars are not KOIs, then we obtained their stellar properties from the update Kepler catalog for stellar proper- ties (Huber et al. 2014). 4. PLANET CANDIDATES AND NOTABLE SYSTEMS Fig. 6 shows a scatter plot of planet radii and orbital periods for planet candidates found with the Kepler data. Most of the known KOIs (88%) have orbital periods shorter than 100 days so the planet candidates discovered by the Planet Hunters help to extend the discovery space into the long period regime. We emphasize that we have included in this paper planet candidates with one or two observed transits which would otherwise excluded by the Ke- pler pipeline. This approach enables the Planet Hunters project to be more sensitive to long-period planet candidates, allowing us to explore a larger parameter space. 4.1. Planet Confidence of Planet Candidates The follow-up observations for these long-period candidates help to exclude false positive scenarios such as background or physically-associated eclips- ing binaries. We use a method called planetary synthesis validation (PSV) to quantify the planet confidence for each planet candidate (Barclay et al. 2013). PSV has been used to validate several planet candidates such as Kepler-69c (Barclay et al. 2013), PH-2b (Wang et al. 2013), and Kepler- 102e (Wang et al. 2014). PSV makes use of tran- siting observations and follow-up observations to 19 https://cfop.ipac.caltech.edu exclude improbable regions in parameter space for false positives. For parameter space that cannot be excluded by observations, PSV adopts a Bayesian approach to calculate the probability of possible false positives and gives an estimation of planet confidence between 0 and 1 with 1 being an ab- solute bona-fide planet. We adopt a planet confi- dence threshold of 0.997 (3-σ) for planet validation. The threshold is more conservative than previous works (e.g., Rowe et al. 2014). The inputs for the PSV code are planet radius, transit depth, pixel centroid offset between in and out of transit, pixel centroid offset significance (i.e., offset divided by measurement uncertainty), num- ber of planet candidates, and the AO contrast curve of the host star in the absence of stellar compan- ion detection. Wang et al. (2013) provided details in the procedures of deriving these inputs and the methodology for the PSV method. The output of the PSV code is the planet confidence, the ratio between planet prior and the sum of the planet prior and possible false positives. Table 7 pro- vides the results of PSV. There are 7 planet candi- dates that have planet confidences over 0.997. Their planet statuses are therefore validated. These plan- ets include three planets with a single transit (KIC- 3558849b or KOI-4307b, KIC-5951458b, and KIC- 8540376c), 3 planets with double transits (KIC- 8540376b, KIC-9663113b or KOI-179b, and KIC- 10525077b or KOI-5800b), and 1 planet with 4 tran- sits (KIC-5437945b or KOI-3791b). The notation for each planet (e.g., b, c, d) starts from the inner- most planet candidate. Since this work contains single and double-transit planet candidates that are typically overlooked by the Kepler mission (Borucki et al. 2010, 2011; Batalha et al. 2013; Burke et al. 2014), it is infor- mative to investigate the false positive rate for this population of planet candidates. Out of 24 candi- date systems for which we have AO data, 6 have detected stellar companions (see Table 6). Depend- ing on which star hosts the transiting object, 4 systems may be false positives due to an underes- timated radius (see further discussions in the fol- lowing sections). These systems are KIC-8510748 (1-transit), KIC-8636333 (2-transit), KIC-11465813 (3-transit), and KIC-12356617 (2-transit). Two other systems have planet candidates whose radii remain in the planetary regime despite the flux di- lution effect (KIC-5732155 and KIC-10255705). In addition, two candidate systems have planet con- fidences lower than 0.85 (KIC-10024862 and KIC- 11716643). If considering all candidate systems with detected stellar companions and candidate systems with planet confidences lower than 0.85 as false pos- itives, an aggressive estimation of the false positive rate for single or double-transit planet candidates is 33%. If considering only the 4 candidates systems that may be false positives due to flux dilution, a conservative estimation of the false positive rate is 17%. 4.2. Single-Transit Systems 6 Planet Hunters VIII 3558849 This star is listed as KOI-4307 and has one planet candidate with period of 160.8 days, but KOI-4307.01 does not match with the single transit event. Therefore this is an additional planet can- didate in the same system. The additional single- transit planet candidate KIC-3558849b is validated with a planet confidence of 0.997. 5010054 This target is not in the threshold cross- ing event (TCE) or KOI tables. Three visible tran- sits are attributed to two planet candidates. The first two at BKJD 356 and 1260 (Schmitt et al. 2014a) are from the same object (they are included in the following Double-Transit Systems section). The third transit at BKJD 1500 is different in both transit depth and duration, so it is modeled here as a single transit from a second planet in the system. 5536555 There are two single-transit events for this target (BKJD 370 and 492). We flag the one at BKJD 370 as a cosmic-ray-induced event. It is caused by Sudden Pixel Sensitivity Dropout (SPSD, Christiansen et al. 2013; Kipping et al. 2015). Af- ter a cosmic ray impact, a pixel can lose its sen- sitivity for hours, which mimics a single-transit event. A cosmic ray hitting event is marked as a SAP QUALITY 128 event when cosmic ray hits pix- els within photometric aperture and marked as a SAP QUALITY 8192 event when cosmic ray hits adjacent pixels of a photometric aperture. The single-transit event at BKJD 370 coincides with with a SAP QUALITY 128 event, so we caution that it may be an artifact. However, the single- transit event at BKJD 492 is still a viable candidate. Single-transit events that are caused by SPSD are also found for other Kepler stars. We list here the SPSDs found by Planet Hunters and the associated BKJDs: KIC-9207021 (BKJD 679), KIC-9388752 (BKJD 508), and KIC-10978025 (BKJD 686). 5951458 This planet candidate KIC-5951458b is validated with a planet confidence of 0.998. 8540376 There are only two quarters of data for this target (Q16 and Q17). However, there are three planet candidates in this system. One starts at BKJD 1499.0 and has an orbital period of 10.7 days. One has only two observed transits with an orbital period of 31.8 days. The two-transit system will be discussed in the following section (§4.3). There is a single-transit event (BKJD 1516.9), which appears to be independent of the previous two planet candi- dates. This single-transit event would be observed again soon because its orbital period has a 1-σ upper limit of 114.1 days. The planet candidate exhibit- ing single transit (KIC-8540376c) is validated with a planet confidence of 0.999. 9704149 There is a second possible transit at BKJD 1117, but only ingress is recorded here and the rest of the transit is lost due to a data gap. If the second transit is due to the same object, then the orbital period is 697.3 days, which is at odds with the estimated period at 1199.3 days. 10024862 In addition to the single transit event, there is also a second object with three visible tran- sits (P = 567.0 days, see §4.4). The triple-transit system was also reported in Wang et al. (2013), but there were only two visible transits at that time. 10403228 This is a transit event from a planet around a M dwarf. Despite the deep transit (∼5%), the radius of the transiting object is within plan- etary range (RP = 9.7 R⊕). However, the transit is v-shaped, suggesting a grazing transit and the true nature of the transiting object is uncertain. For the M star, we adopt stellar mass and radius from Huber et al. (2014) which uses the Dartmouth stellar evolutionary model (Dotter et al. 2008). 10842718 The orbital period distribution given the constraints from transit duration and stellar density (§2.3) has two peaks. One is at ∼1630 days, the other one is at ∼10,000 days. The bi- modal distribution suggests that the orbital period of this transiting object could be much longer than reported in Table 1, however the probability for a transiting planet with a period of ∼10,000 days is vanishingly low, giving stronger weight to the shorter period peak. 4.3. Double-Transit Systems object is 3756801 This first mentioned in Batalha et al. (2013) and designated as KOI- 1206. Surprisingly, it appears that only one transit was detected by Batalha et al. (2013). It does not appear in the Kepler TCE table because a third transit was not observed. 5732155 A stellar companion has been detected in KS band that is 4.94 magnitudes fainter. The separation of the stars is 1′′ (Table 6). The flux con- tamination does not significantly change the transit depth and thus does not affect planet radius estima- tion. The stellar companion is so faint that even a total eclipsing binary would not yield the observed transit depth. 6191521 This target is listed as KOI-847 and has one planet candidate with orbital period of 80.9 days. Here, we report a second, longer-period planet candidate that was not previously detected in the system. 8540376 There are only two quarters of data for this target (Q16 and Q17), but there are three planet candidates in this system. The longer period single-transit event has been discussed in §4. The double-transit event starts at BKJD 1520.3 and has a period of 31.8 days. The shortest period planet (10.7 days) has transits that begin at BKJD 1499.0. The double-transit planet candidate KIC-8540376b is validated with a planet confidence of 0.999. 8636333 This target is listed as KOI-3349 and has two planet candidates. One is KOI-3349.01 with period of 82.2 days; the other one was re- ported in Wang et al. (2013) with period of 804.7 days. Here, we report the follow-up observations for this star: a fainter stellar companion has been detected in H and KS bands (Table 6) with differ- ential magnitudes of 1.58 and 1.71 in these filters respectively. We estimate their Kepler band magni- tudes to be different by ∼ 3 mag. Based on Fig. 11 in Horch et al. (2014), a correction for the radius of the planet that accounts for flux from the stel- lar companion would increase the planet radius by a small amount, ∼3%. However, if the two candi- dates are transiting the fainter secondary star, then Long-Period Exoplanets 7 their radii would increase by a factor of ∼3. In this case, although the radii for both candidates would remain in planetary range, the longer-period candi- date would be at the planetary radius threshold. 9663113 This target is listed as KOI-179 and has two planet candidates. One is KOI-179.01 with period of 20.7 days. KOI-179.02 was reported in Wang et al. (2013) with period of 572.4 days with two visible transits. The expected third transit at BKJD 1451 is missing, but the expected position is in a data gap. The double-transit planet candidate KIC-9663113b is validated with a planet confidence of 0.999. was This target 10255705 reported in Schmitt et al. (2014a). Follow-up AO ob- servation shows that there is a nearby stellar companion (Table 6). The companion is ∼2 mag fainter in Kepler band. If the planet candidate orbits the primary star, then the planet radius ad- justment due to flux contamination is small. If the planet candidate orbits around the newly detected stellar companion, then the planet radius is revised upward by a factor of ∼2 (Horch et al. 2014), but the adjusted radius is still within planetary range. 10460629 This target is listed as KOI-1168 and has one planet candidate that matches with the double-transit event. There are two deep v-shaped dips in the lightcurve at BKJD 608.3 and 1133.3, likely indicating an eclipsing binary within the planet orbit. These v-shaped transits are so deep (about 13%) that they could easily be followed up from the ground. If the planet interpretation is cor- rect for the other two transit events, then this could be an circumbinary planet candidate. However, this system is likely to be a blending case in which two stars are in the same photometric aperture. This system is discussed further in §5.2. 10525077 This target is listed as KOI-5800 and has one planet candidate with period of 11.0 days. The second planet candidate was reported in Wang et al. (2013) with period of 854.1 days. There are two transits at BKJD 355.2 and 1189.3. In between these two transits, there is a data gap at 762.3, preventing us from determining whether the orbital period is 854.1 days or half of the value, i.e, 427.05 days. This planet candidate KIC-10525077b is validated with a planet confidence of 0.998. 12356617 This target is listed as KOI-375 and has one planet candidate that matches with the double-transit event. Follow-up AO observation shows that there is one faint stellar companion at 3.12′′ separation. If the transit occurs for the pri- mary star, the radius adjustment due to flux con- tamination is negligible. If the transit occurs for the secondary star, then this is a false positive. 4.4. Triple or Quadruple Transit Systems candidates two planet 5437945 This target is listed as KOI-3791and in 2:1 resonance. has KOI-3791.01 was reported in Wang et al. (2013) and Huang et al. (2013). The fourth transit ap- pears at BKJD 1461.8. The longer-period planet candidate KIC-5437945b is validated with a planet confidence of 0.999. 5652983 This target is listed as KOI-371 and has one planet candidate that matches with the triple- transit event. The radius of the transiting object is too large to be a planet, and thus the triple-transit event is a false positive, which is supported by the notes from CFOP that large RV variation has been observed. 6436029 This target is listed as KOI-2828 and has two planet candidates. KOI-2828.02 with pe- riod of 505.5 days matches the triple-transit event. KOI-2828.02 was reported in Schmitt et al. (2014a), but there were only two visible transits. 7619236 This target is listed as KOI-5205 and has one planet candidate that matches with the triple-transit event. It exhibits significant transit timing variations (TTVs). The time interval be- tween the first two transits is different by ∼27 hours from the time interval between the second and the third transit. 10024862 8012732 This object was reported in Wang et al. (2013). It exhibits significant TTVs. The time in- terval between the first two transits is different by ∼20 hours from the time interval between the sec- ond and the third transit. 9413313 This object was reported in Wang et al. (2013). It exhibits significant TTVs. The time in- terval between the first two transits is different by ∼30 hours from the time interval between the sec- ond and the third transit. This reported in Wang et al. (2013), but only two transit were observed then. The third transit is observed at BKJD 1493.8. It exhibits significant TTVs. The time interval between the first two transits is different by ∼41 hours from the time interval between the second and the third transit. 10850327 This target is listed as KOI-5833 and has one planet candidate that matches with the triple-transit event. The object was reported in Wang et al. (2013), but there were only two tran- sit observed then. object was 11465813 This target is listed as KOI-771 and has one planet candidate that matches with the triple-transit event. The transit depth is varying. This target also has a single transit at BKJD 1123.5. A stellar companion has been detected (Table 6). From the colors of the companion, we estimate the differential magnitude to be 0.7 mag. The radius of the object would be revised upward by 23% or 150% depending on whether the object orbits the primary or the secondary star. In either case, it is likely that this object is a false positive. 11716643 This target is listed as KOI-5929 and was reported in Wang et al. (2013), but there were only two transits observed then. It exhibits TTVs. The time interval between the first two transits is different by ∼2.7 hours from the time interval be- tween the second and the third transit. 4.5. Notable False Positives In addition to the systems with single-transit events flagged as SPSDs in §4.2, we list other tran- siting systems that are likely to be false positives. 1717722 This target is listed as KOI-3145 with 8 Planet Hunters VIII two known planet candidates. Neither candidate matches the single transit event at BKJD 1439. This single transit is likely spurious, as pixel cen- troid offset between in- and out-of-transit are seen for this transit. 3644071 This target is listed as KOI-1192 and has one false positive (02) and one candidate (01). The epoch for candidate KOI-1192.01 matches with the epoch of the single transit event in this paper. According to notes on CFOP, the KOI-1192 event is “due to video crosstalk from an adjacent CCD readout channel of the image of a very bright, highly saturated star”. This effect causes the varying tran- sit depth and duration. The explanation is further supported by the apparent pixel offset between in- and out-of-transit for both KOI-1192.01 and KOI- 1192.02. So KOI-1192.01 is also likely to be a false positive. 9214713 This target is listed as KOI-422 and has one planet candidate that matches with the double- transit event found by Planet Hunters. Significant pixel centroid offset is found between in- and out- of-transit although follow-up AO and spectroscopic observations show no sign of nearby stellar compan- ions. 10207400 This target is currently not in either the Kepler KOI or TCE tables. There is a pixel centroid offset between in- and out-of-transit. 5. SUMMARY AND DISCUSSION 5.1. Summary We report 41 long-period planet candidates around 38 Kepler stars. These planet candidates are identified by the Planet Hunters based on the archival Kepler data from Q0 to Q17. We conduct AO imaging observations to search for stellar com- panions and exclude false positive scenarios such as eclipsing binary blending. In total, we obtain AO images for 33 stars. We detect stellar compan- ions around 6 stars within 4′′, KIC-5732155, KIC- 8510748, KIC-8636333 (KOI-3349), KIC-10255705, KIC-11465813 (KOI-771), and KIC-12356617 (KOI- 375). The properties of these stellar companions are given in Table 6. For those stars with non- detections, we provide AO sensitivity limits at dif- ferent angular separations (Table 5). We obtain high-resolution spectra for a total of 6 stars. We use the stellar spectra to infer stellar properties such as stellar mass and radius which are used for orbital period estimation for single-transit events. The stellar properties of planet host stars are given in Table 4. We model the transiting light curves with TAP to obtain their orbital parameters. Ta- ble 1, Table 2 and Table 3 give the results of light curve modeling for single-transit, double-transit, and triple/quardruple-transit systems, respectively. Based on transiting and follow-up observations, we calculate the planet confidence for each planet can- didate. Seven planet candidates have planet con- fidence above 0.997 and are thus validated. These planets include 3 planets with a single transit (KIC- 3558849b or KOI-4307b, KIC-5951458b, and KIC- 8540376c), 3 planets with double transits (KIC- 8540376b, KIC-9663113b or KOI-179b, and KIC- 10525077b or KOI-5800b), and 1 planet with 4 tran- sits (KIC-5437945b or KOI-3791b). We estimate the false positive rate to be 17%-33% for 1-transit and 2-transit events. 5.2. KIC-10460629: An Interesting Case At first glance, KIC-10460629 might be an ex- treme circumbinary planetary system if confirmed with a P = 856.7 days planet and a P = 525 days eclipsing binary star. The ratio of semi-major axis of the transiting planet to the eclipsing sec- ondary star is ∼1.4. The tight orbital configura- tion makes the system dynamically unstable. Ac- cording to Equation 3 in Holman & Wiegert (1999), the minimum semi-major axis ratio for a stable or- bit around a binary star is 2.3 for a binary with e = 0 and µ = 0.5, where µ is the mass ratio of the primary to the secondary star estimated from the transit depth (13%). Therefore, KIC-10460629 should be dynamical unstable. Furthermore, the minimum semi-major axis ratio increases with in- creasing eccentricity, which makes the systems even more unstable for eccentric orbits based on the cri- terion from Holman & Wiegert (1999). A more likely explanation for the observed two sets of transits is a blending case, in which two stars are within the photometric aperture and each has one set of transits. There are two possibilities in the blending case. For the first case (referred to as Case A), the deep transit takes place around the brighter star and the shallower transit takes place around the fainter star. In this case, the fainter star needs to brighter than 7.8 differential magnitude in the Kepler band, otherwise it does not produce the observed 784 ppm transit even with a total ecplise. Our AO observations are not deep enough to rule out this scenario. For second possible blending case (referred to as Case B), the deep transit takes place around the fainter star and the shallower transit takes place around the brighter star. In this case, the fainter star needs to be brighter than 2.2 dif- ferential magnitude in the Kepler band in order to produce the observed 13% transit depth. This pos- sibility is ruled out by our AO observations for an- gular separations larger than 0′′.1. The blending has to happen within 0′′.1 angular separation for Case B. There might be a Case C, in which the two shallower transits are not caused by the same ob- ject. However, there is no evidence that this is the case given the similarity of the two transits (see Fig. 3). Follow-up observations are necessary to determine the nature of this transiting system. For Case A, a deeper AO observation is required to confirm or rule out the fainter star. For Case B, a high-resolution spectroscopy of the target would reveal the fainter source since it is at least 13% as bright as the brighter star. Long time-baseline RV observations can also differentiate the two cases. For Case A, a clear stellar RV signal should be observed. In case B, the precision of RV measurements may not be adequate to map out the orbit of the transiting planet candidate with a Neptune-size given the low- Long-Period Exoplanets 9 mass and the faintness of the host star (KP = 14.0). Ground-based transiting follow-up observations can certainly catch the transit of the secondary star at 13% depth. The next transit of the secondary star will on UT June 1st 2016. The transit depth of the planet candidate is ∼800 ppm, which may be de- tected by ground-based telescopes. The next shal- lower transit will be on UT August 30 2016. This may confirm or rule out Case C. 5.3. Evidence of Additional Planets in Systems with Long-Period Transiting Planets TTVs indicate the likely presence of additional components in the same system that dynamically interacting with transiting planet candidates. For the 10 systems with 3-4 visible transits for which we can measure TTVs, 50% (5 out of 10) exhibit TTVs ranging from ∼2 to 40 hours. Four systems have synodic TTV larger than 20 hours, making them the “queens” of transit variations as opposed to the 12- hr “king” system KOI-142 (Nesvorn´y et al. 2013). Excluding two likely false positives, KIC-5652983 (large RV variation) and KIC-11465813 (blending), the fraction of systems exhibiting TTVs goes up to 68%. All such systems host giant planet can- didates with radii ranging from 4.2 to 12.6 R⊕. Based on Equation 10 in Deck & Agol (2015), or Fig. 6 in Nesvorn´y & Vokrouhlick´y (2014), an or- der of magnitude estimation for the mass of the per- turber is a few Jupiter masses, assuming an orbital separation corresponding to a period ratio . 2, and low eccentricity orbits. In general, to maintain the same amplitude synodic TTV, the mass of the per- turber would need to increase for wider separations, and decrease for smaller. This result suggests that most long-period tran- siting planets have at least one additional compan- ion in the same system. This finding is consistent with the result in Fischer et al. (2001) that almost half (5 out of 12) of gas giant planet host stars exhibit coherent RV variations that are consistent with additional companions. This finding is further supported by a more recent study of companions to systems with hot Jupiters (Knutson et al. 2014; Ngo et al. 2015), in which the stellar and plane- tary companion rate of hot Jupiter systems is es- timated to be ∼50%. While we emphasize the dif- ferent planet populations between previous studies (short-period planets) and systems reported in this paper (long-period planets), the companion rate for stars with gas giant planets is high regardless of the orbital period of a planet. Dawson & Murray-Clay (2013) found that giant planets orbiting metal-rich stars show signatures of planet-planet interactions, suggesting that multi- planet systems tend to favorably reside in metal- rich star systems. We check the metallicities of the five systems exhibiting TTVs. The median metal- licity is 0.07 ± 0.18. In comparison, the median metallicity for the entire sample is −0.06± 0.38 dex. While there is a hint that the TTV sample is more metal rich, the large error bars and the small sam- ple prevent us from further studying the metallic- ity distribution of systems exhibiting TTVs. How- ever, studying the metallicity of planet host stars remains a viable tool and future follow-up observa- tions would allow us to use the tool to test planet formation theory. 5.4. The Occurrence Rate of Long-Period Planets The presence of long-period planets may affect the evolution of multi-planet systems by dynami- cal interaction (e.g., Rasio & Ford 1996; Dong et al. 2014). The dynamical effects result in observable effects such as spin-orbit misalignment which pro- vides constraints on planet migration and evolu- tion (e.g., Winn et al. 2010). Therefore, measur- ing the occurrence rate of long-period planets is essential in determine their role in planet evolu- tion. Cumming et al. (2008) estimated that the oc- current rate is 5-6% per period decade for long- period gas giant planets. Knutson et al. (2013) es- timated that 51% ± 10% of hot-Jupiter host stars have an additional gas giant planet in the same sys- tem. However, these studies are sensitive to planets with mass higher than ∼0.3 Jupiter mass. The Ke- pler mission provides a large sample of small planets (likely to be low-mass planets), which can be used to infer the occurrence rate for small, long-period planets. However, such analysis is limited to periods up to ∼500 days (Dong & Zhu 2013; Petigura et al. 2013; Rowe et al. 2015). The upper limit is due to the 3-transit detection criterion for Kepler planet candidates. With the long-period planet candidates in this paper, we will be able to probe the occurrence rate of planets between 1 and 3 AU. To accomplish this goal, a proper assessment of planet recovery rate of the Planet Hunters is required. The frame- work has already been provided by Schwamb et al. (2012b) and this issue will be addressed in a future paper. Estimating the occurrence rate of Neptune to Jupiter-sized planets between 1 and 3 AU will be an important contribution of Planet Hunters to the exoplanet community. 5.5. K2 and TESS The current K2 mission and future TESS (Tran- siting Exoplanet Survey Satellite) missions have much shorter continuous time coverage than the Ke- pler mission. Each field of the K2 mission receives ∼75 days continuous observation (Howell et al. 2014). For the TESS mission, the satellite stays in the same field for 27.4 days (Ricker et al. 2015). Despite longer time coverage for a portion of its field, the majority of sky coverage of TESS will re- ceive only 27.4 days observation. Given the scan- ning strategy of these two missions, there will be many single-transit events. Estimating the orbital periods for these events is crucial if some the tar- gets with a single transit have significant scientific value, e.g., planets in the habitable zone. More generally, estimating orbital period helps to predict the next transit and facilitates follow-up observa- tions, especially for those searching for the next transit (Yee & Gaudi 2008). Once more than one transits are observed, more follow-up observations can be scheduled such as those aiming to study tran- 10 Planet Hunters VIII siting planets in details, e.g., CHEOPS (CHaracter- ising ExOPlanet Satellite) and JWST (James Webb Space Telescope). Acknowledgements The authors would like to thank the anonymous referee whose comments and sug- gestions greatly improve the paper. We are grateful to telescope operators and supporting astronomers at the Palomar Observatory and the Keck Observa- tory. Some of the data presented herein were ob- tained at the W.M. Keck Observatory, which is op- erated as a scientific partnership among the Cal- ifornia Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possi- ble by the generous financial support of the W.M. Keck Foundation. The research is made possible by the data from the Kepler Community Follow- up Observing Program (CFOP). The authors ac- knowledge all the CFOP users who uploaded the AO and RV data used in the paper. We thank Kather- ine M. Deck for insightful comments on TTV sys- tems and the dynamical stability of KIC-10460629. This research has made use of the NASA Exoplanet Archive, which is operated by the California Insti- tute of Technology, under contract with the Na- tional Aeronautics and Space Administration un- der the Exoplanet Exploration Program. JW, DF and TB acknowledge the support from NASA under Grant No. NNX12AC01G and NNX15AF02G. The Robo-AO system was developed by collab- orating partner institutions, the California Insti- tute of Technology and the Inter-University Cen- tre for Astronomy and Astrophysics, and with the support of the National Science Foundation under Grant No. AST-0906060, AST-0960343 and AST- 1207891, the Mt. Cuba Astronomical Foundation and by a gift from Samuel Oschin.C.B. acknowl- edges support from the Alfred P. Sloan Foundation. KS gratefully acknowledges support from Swiss Na- tional Science Foundation Grant PP00P2 138979/1 5m PO:1.5m (Robo-AO) PO: Facilities: (PHARO) KO: 10m (NIRC2) Baranec, C., et al. 2013, Journal of Visualized Experiments, Huang, X., Bakos, G. ´A., & Hartman, J. D. 2013, MNRAS, REFERENCES 72, e50021 —. 2014, ApJ, 790, L8 Barclay, T., et al. 2013, ApJ, 768, 101 Batalha, N. M., et al. 2013, ApJS, 204, 24 Borucki, W. J., et al. 2010, Science, 327, 977 —. 2011, ApJ, 736, 19 Brandl, B., Hayward, T. L., Houck, J. R., Gull, G. E., Pirger, B., & Schoenwald, J. 1997, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 3126, Adaptive Optics and Applications, ed. R. K. Tyson & R. Q. Fugate, 515 Burke, C. J., et al. 2014, ApJS, 210, 19 Burruss, R. S., et al. 2014, in Presented at the Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol. 9148, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Cassan, A., et al. 2012, Nature, 481, 167 Christiansen, J. L., et al. 2013, ApJS, 207, 35 Cumming, A., Butler, R. P., Marcy, G. W., Vogt, S. S., Wright, J. T., & Fischer, D. A. 2008, PASP, 120, 531 Dawson, R. I., & Murray-Clay, R. A. 2013, ApJ, 767, L24 Deck, K. M., & Agol, E. 2015, ApJ, 802, 116 Demarque, P., Woo, J.-H., Kim, Y.-C., & Yi, S. K. 2004, ApJS, 155, 667 Dong, S., Katz, B., & Socrates, A. 2014, ApJ, 781, L5 Dong, S., & Zhu, Z. 2013, ApJ, 778, 53 Dotter, A., Chaboyer, B., Jevremovi´c, D., Kostov, V., Baron, E., & Ferguson, J. W. 2008, ApJS, 178, 89 Dressing, C. D., Adams, E. R., Dupree, A. K., Kulesa, C., & McCarthy, D. 2014, AJ, 148, 78 Fischer, D. A., Marcy, G. W., Butler, R. P., Vogt, S. S., Frink, S., & Apps, K. 2001, ApJ, 551, 1107 Fischer, D. A., et al. 2012, MNRAS, 419, 2900 Fressin, F., et al. 2013, ApJ, 766, 81 Fruchter, A. S., & Hook, R. N. 2002, PASP, 114, 144 Gaudi, B. S. 2010, ArXiv e-prints Gazak, J. Z., Johnson, J. A., Tonry, J., Dragomir, D., Eastman, J., Mann, A. W., & Agol, E. 2012, Advances in Astronomy, 2012, 30 Gonzalez, G., & Vanture, A. D. 1998, A&A, 339, L29 Hayward, T. L., Brandl, B., Pirger, B., Blacken, C., Gull, G. E., Schoenwald, J., & Houck, J. R. 2001, PASP, 113, 105 Holman, M. J., & Wiegert, P. A. 1999, AJ, 117, 621 Horch, E. P., Howell, S. B., Everett, M. E., & Ciardi, D. R. 2014, ApJ, 795, 60 Howell, S. B., et al. 2014, PASP, 126, 398 429, 2001 Huber, D., et al. 2014, ApJS, 211, 2 Jenkins, J. M., et al. 2010, ApJ, 713, L87 Kipping, D. M., Huang, X., Nesvorn´y, D., Torres, G., Buchhave, L. A., Bakos, G. ´A., & Schmitt, A. R. 2015, ApJ, 799, L14 Knutson, H. A., et al. 2013, ArXiv e-prints —. 2014, ApJ, 785, 126 Kurucz, R. L. 1993, SYNTHE spectrum synthesis programs and line data Lintott, C. J., et al. 2013, AJ, 145, 151 Lovis, C., et al. 2011, A&A, 528, A112 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Nesvorn´y, D., Kipping, D., Terrell, D., Hartman, J., Bakos, G. ´A., & Buchhave, L. A. 2013, ApJ, 777, 3 Nesvorn´y, D., & Vokrouhlick´y, D. 2014, ApJ, 790, 58 Ngo, H., et al. 2015, ApJ, 800, 138 Oppenheimer, B. R., & Hinkley, S. 2009, ARA&A, 47, 253 Perryman, M. A. C., et al. 2001, A&A, 369, 339 Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Proceedings of the National Academy of Sciences Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954 Ricker, G. R., et al. 2015, Journal of Astronomical Telescopes, Instruments, and Systems, 1, 014003 Rowe, J. F., et al. 2014, ApJ, 784, 45 —. 2015, ApJS, 217, 16 Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153 Schmitt, J. R., et al. 2014a, AJ, 148, 28 —. 2014b, ApJ, 795, 167 Schwamb, M. E., et al. 2012a, ArXiv e-prints —. 2012b, ApJ, 754, 129 —. 2013, ApJ, 768, 127 Sneden, C. A. 1973, PhD thesis, THE UNIVERSITY OF TEXAS AT AUSTIN. Sousa, S. G., et al. 2008, A&A, 487, 373 Torres, G., et al. 2011, ApJ, 727, 24 Wang, J., Fischer, D. A., Xie, J.-W., & Ciardi, D. R. 2015, ApJ, 813, 130 Wang, J., Xie, J.-W., Barclay, T., & Fischer, D. A. 2014, ApJ, 783, 4 Wang, J., et al. 2013, ApJ, 776, 10 Winn, J. N. 2010, Exoplanet Transits and Occultations, ed. S. Seager, 55–77 Winn, J. N., Fabrycky, D., Albrecht, S., & Johnson, J. A. 2010, ApJ, 718, L145 Long-Period Exoplanets 11 Wizinowich, P. L., Acton, D. S., Lai, O., Gathright, J., Lupton, W., & Stomski, P. J. 2000, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4007, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. P. L. Wizinowich, 2–13 Yee, J. C., & Gaudi, B. S. 2008, ApJ, 688, 616 Yelda, S., Lu, J. R., Ghez, A. M., Clarkson, W., Anderson, J., Do, T., & Matthews, K. 2010, ApJ, 725, 331 1.0001 1.0000 0.9999 0.9998 0.9997 KIC 2158850 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 KIC 3558849 −30 −20 −10 0 10 20 30 −15 −10 −5 0 5 10 15 x u l f d e z i l a m r o N x u l f d e z i l a m r o N x u l f d e z i l a m r o N 1.0005 1.0000 0.9995 0.9990 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 0.9940 0.9930 KIC 5536555 −30 −20 −10 0 10 20 30 KIC 8410697 1.0008 1.0006 1.0004 1.0002 1.0000 0.9998 0.9996 0.9994 0.9992 1.0002 1.0001 1.0000 0.9999 0.9998 0.9997 0.9996 1.0006 1.0004 1.0002 1.0000 0.9998 0.9996 0.9994 0.9992 1.0004 1.0002 1.0000 0.9998 0.9996 0.9994 0.9992 0.9990 0.9988 KIC 5010054 −30 −20 −10 0 10 20 30 KIC 5951458 KIC 5536555 −30 −20 −10 0 10 20 30 −20 −10 0 10 20 1.0005 1.0000 0.9995 KIC 8540376 KIC 8510748 −20 −10 0 10 20 −30 −20 −10 0 10 20 30 −20 −10 0 10 20 Phase (hours) Phase (hours) Phase (hours) Fig. 1.— Transiting light curves for 1-transit planet candidates. Blue open circles are data points and black solid line is the best-fitting model. Orbital parameters can be found in Table 1. 12 Planet Hunters VIII x u l f d e z i l a m r o N x u l f d e z i l a m r o N x u l f d e z i l a m r o N 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 0.9975 0.9970 0.9965 0.9960 1.0100 1.0000 0.9900 0.9800 0.9700 0.9600 0.9500 0.9400 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 0.9975 KIC 9838291 KIC 9704149 −10 −5 0 5 10 −30 −20 −10 0 10 20 30 1.0020 1.0000 0.9980 0.9960 KIC 10403228 0.9940 KIC 10842718 1.0020 1.0000 0.9980 0.9960 0.9940 0.9920 0.9900 0.9880 1.0010 1.0005 1.0000 0.9995 0.9990 0.9985 KIC 10024862 −15 −10 −5 0 5 10 15 KIC 10960865 −60 −40 −20 0 20 40 60 −40 −20 0 20 40 −40 −20 0 20 40 1.0010 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 0.9975 0.9970 KIC 11558724 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 KIC 12066509 Phase (hours) −15 −10 −5 0 5 10 15 −15 −10 −5 0 5 10 15 Phase (hours) Phase (hours) Fig. 2.— Transiting light curves for 1-transit planet candidates. Blue open circles are data points and black solid line is the best-fitting model. Orbital parameters can be found in Table 1. 1.0010 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 0.9940 1.0005 1.0000 0.9995 0.9990 0.9985 1.0010 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 KIC 3756801 −40 −30 −20 −10 0 10 20 30 40 KIC 6191521 x u l f d e z i l a m r o N x u l f d e z i l a m r o N x u l f d e z i l a m r o N x u l f d e z i l a m r o N Long-Period Exoplanets 13 1.0001 1.0001 1.0000 1.0000 0.9999 0.9999 KIC 5522786 1.0060 1.0040 1.0020 1.0000 0.9980 0.9960 0.9940 KIC 5732155 KIC 5010054 −20 −10 0 10 20 −40 −30 −20 −10 0 1.0020 10 20 30 40 1.0020 −40−30−20−10 0 10 20 30 40 1.0010 1.0000 0.9990 0.9980 0.9970 KIC 8636333 KIC 8540376 0 5 10 0.9960 15 1.0010 −20 −10 0 10 20 1.0010 1.0000 0.9990 KIC 9662267 0.9980 −20 −15 −10 −5 1.0020 0 5 10 15 20 −20 −15 −10 −5 1.0005 0 5 10 15 20 −15 −10 −5 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 0.9975 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 0.9940 1.0000 0.9995 0.9990 0.9985 1.0010 1.0005 1.0000 0.9995 0.9990 0.9985 KIC 9663113 −30 −20 −10 0 10 20 30 KIC 12356617 1.0005 1.0000 0.9995 0.9990 0.9985 KIC 10460629 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 KIC 10525077 KIC 10255705 −40 −20 0 20 40 −20 −10 0 10 20 −30 −20 −10 0 10 20 30 Phase (hours) Phase (hours) KIC 12454613 −10 −5 0 5 10 −20 −10 0 10 20 Phase (hours) Phase (hours) Fig. 3.— Transiting light curves for 2-transit planet candidates. Blue and red open circles are data points for odd- and even-numbered transits. Black solid line is the best-fitting model. Orbital parameters can be found in Table 2. 14 Planet Hunters VIII x u l f d e z i l a m r o N x u l f d e z i l a m r o N x u l f d e z i l a m r o N 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 0.9975 0.9970 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 0.9940 1.0050 1.0000 0.9950 0.9900 0.9850 0.9800 0.9750 KIC 5437945 −20 −10 0 10 20 KIC 8012732 −15 −10 −5 0 5 10 15 KIC 11465813 1.0005 1.0000 0.9995 0.9990 0.9985 0.9980 1.0020 1.0000 0.9980 0.9960 0.9940 0.9920 0.9900 1.0020 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 1.0030 1.0020 1.0010 1.0000 0.9990 0.9980 0.9970 1.0020 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 1.0010 1.0000 0.9990 0.9980 0.9970 0.9960 0.9950 0.9940 0.9930 KIC 7619236 KIC 6436029 −20 −10 0 10 20 −10 −5 0 5 10 1.0005 1.0000 0.9995 0.9990 KIC 10850327 KIC 10024862 KIC 5652983 −10 −5 0 5 10 KIC 9413313 −20 −10 0 10 20 −30 −20 −10 0 10 20 30 −20 −10 0 10 20 Phase (hours) Phase (hours) KIC 11716643 −40 −30 −20 −10 0 10 20 30 40 −20 −10 0 10 20 Phase (hours) Phase (hours) Fig. 4.— Transiting light curves for 3-transit planet candidates. Blue and red open circles are data points for odd- and even-numbered transits. Black solid line is the best-fitting model. Orbital parameters can be found in Table 3. 12 10 8 6 4 2 s m e t s y S f o r e b m u N 0 −3 −2 −1 1 2 3 0 P−¯P δP s m e t s y S f o r e b m u N 14 12 10 8 6 4 2 0 0 1 2 3 4 5 δP P Fig. 5.— Left: distribution of the difference between the period estimated from individual transit ( ¯P ) and the period estimated from the time interval of consecutive transits (P ) for 24 candidate planetary systems with 2-3 visible transits. The difference is normalized by measurement uncertainty of δP . Right: distribution of the fractional error δP/P . Long-Period Exoplanets 15 Fig. 6.— Scatter plot of planet radii vs. orbital periods for planet candidates discovered with Kepler data. Black dots are Kepler planet candidates. Red filled circles are planet candidates from this work that are identified by Planet Hunters. Planet candidates with a single transit are marked with red crosses. Yellow filled circles are planet candidates from previous Planet Hunters papers: PH I (Fischer et al. 2012), PH II: (Schwamb et al. 2012a), PH III: (Schwamb et al. 2013), PH IV: (Lintott et al. 2013), PH V: (Wang et al. 2013), PH VI: (Schmitt et al. 2014a), PH VII: (Schmitt et al. 2014b). Long-period planet candidates are predominantly discovered by Planet Hunters. KIC KOI P Mode (days) P Range (days) a/R∗ Inclination RP/R∗ µ1 µ2 Comments See §4.2 for details Orbital Parameters (1 visible transit) TABLE 1 1 6 04307 2158850† 3558849† 5010054† 5536555† 5536555† 5951458† 8410697† 8510748†,†† 8540376† 9704149† 9838291† 10024862† 10403228 10842718 10960865 11558724 12066509 1203.8 1322.3 1348.2 3444.7 1188.4 1320.1 1104.3 1468.3 75.2 1199.3 3783.8 735.7 88418.1 1629.2 265.8 276.1 984.6 [1179.3..2441.6] [1311.1..1708.4] [1311.2..3913.9] [1220.7..9987.4] [1098.6..4450.9] [1167.6..13721.9] [1048.9..2717.8] [1416.0..5788.4] [74.1..114.1] [1171.3..2423.2] [1008.5..8546.1] [713.0..1512.8] [846.5..103733.3] [1364.7..14432.2] [233.7..3335.9] [267.0..599.3] [959.0..1961.7] 1037.9+225.7 −274.6 576.7+21.5 −50.2 825.1+134.8 −264.2 908.4+119.4 −191.1 431.8+71.1 −97.8 278.1+109.1 −66.0 446.0+7.7 −17.1 569.5+145.4 −203.5 103.9+14.1 −14.1 600.9+71.2 −121.8 930.4+72.1 −97.5 324.1+41.2 −36.2 13877.4+400.0 −408.4 347.5+19.8 −23.9 99.7+13.7 −28.6 181.1+10.1 −25.8 460.8+89.4 −72.3 (deg) 89.956+0.058 −0.110 89.973+0.023 −0.031 89.963+0.140 −0.250 89.965+0.024 −0.038 89.887+0.068 −0.130 89.799+0.090 −0.120 89.976+0.024 −0.031 89.938+0.044 −0.073 89.701+0.160 −0.160 89.955+0.059 −0.076 89.974+0.069 −0.063 89.905+0.030 −0.022 89.996+0.011 −0.009 89.938+0.010 −0.008 89.703+0.240 −0.530 89.897+0.021 −0.032 89.925+0.050 −0.036 0.013+0.002 −0.001 0.063+0.002 −0.002 0.021+0.002 −0.002 0.024+0.003 −0.003 0.024+0.002 −0.001 0.040+0.089 −0.008 0.072+0.001 −0.001 0.012+0.001 −0.001 0.018+0.004 −0.005 0.054+0.003 −0.003 0.043+0.001 −0.001 0.098+0.004 −0.004 0.269+0.022 −0.024 0.071+0.002 −0.002 0.024+0.003 −0.003 0.043+0.002 −0.002 0.062+0.003 −0.003 RP (R⊕) 1.6+1.0 −0.8 6.9+1.0 −0.9 3.4+1.8 −1.6 2.7+1.4 −1.1 2.7+1.2 −1.0 6.6+26.3 −4.2 9.8+4.9 −4.7 3.6+2.4 −2.1 2.4+1.9 −1.4 5.0+1.4 −1.3 5.1+1.8 −1.8 11.8+3.7 −3.4 9.7+2.4 −2.2 9.9+5.4 −5.0 3.9+3.0 −2.5 5.9+2.9 −2.7 7.1+2.3 −2.2 Epoch (BKJD) 411.791+0.008 −0.008 279.920+0.440 −0.300 1500.902+0.008 −0.009 370.260+0.033 −0.038 492.410+0.009 −0.008 423.463+0.010 −0.013 542.122+0.001 −0.001 1536.548+0.013 −0.015 1516.911+0.020 −0.020 419.722+0.007 −0.007 582.559+0.003 −0.004 878.561+0.004 −0.004 744.843+0.013 −0.013 226.300+1.100 −0.520 1507.959+0.007 −0.006 915.196+0.003 −0.003 632.090+0.004 −0.004 0.520+0.330 0.350+0.300 −0.230 0.400+0.380 0.500+0.350 −0.330 0.510+0.340 0.520+0.330 −0.350 0.410+0.130 −0.130 0.500+0.300 −0.300 0.510+0.340 −0.340 0.490+0.330 0.280+0.250 −0.180 0.370+0.310 −0.250 0.550+0.310 −0.370 0.700+0.180 0.510+0.330 0.470+0.330 0.360+0.330 −0.250 Note. — † : Targets with AO follow-up observations. ††: Targets with detected stellar companions as reported in Table 6. The AO detection limits are given in Table 5. −0.350 −0.070+0.450 −0.430 0.220+0.370 −0.460 −0.280 −0.040+0.420 −0.440 Multi 0.070+0.430 −0.420 −0.340 −0.140+0.410 −0.410 Multi, Validated Cosmic artifact Validated Binary Multi, Validated, Q16 and Q17 data only 0.000+0.430 −0.420 0.210+0.240 −0.230 0.000+0.450 −0.450 0.000+0.440 −0.420 0.430+0.280 −0.370 0.280+0.370 −0.500 0.050+0.420 −0.410 −0.320 −0.080+0.450 −0.440 Possible incomplete second transit Multi V-shape −0.240 −0.150+0.400 −0.340 −0.010+0.450 −0.310 −0.130+0.460 0.240+0.390 −0.500 −0.430 −0.450 P l −0.300 Bimodal in inferred period a n e t H u n t e r s V I I I KIC KOI P (days) a/R∗ Inclination RP/R∗ µ1 µ2 Comments See §4.3 for details 3756801 5010054† 5522786† 5732155†,†† 6191521 8540376 8636333†† 9662267† 9663113 10255705†,†† 10460629 10525077 10525077 12356617†† 12454613† 01206 00847 03349 00179 01168 05800 05800 00375 422.91360+0.01608 −0.01603 904.20180+0.01339 −0.01212 757.09520+0.01176 −0.01211 644.21470+0.01424 −0.01598 1106.24040+0.00922 −0.00954 31.80990+0.00919 −0.00933 804.71420+0.01301 −0.01500 466.19580+0.00850 −0.00863 572.38470+0.00583 −0.00567 707.78500+0.01844 −0.01769 856.67100+0.01133 −0.01039 854.08300+0.01628 −0.01697 427.04150+0.01487 −0.01628 988.88111+0.00137 −0.00146 736.37700+0.01531 −0.01346 92.2+21.0 −27.0 291.9+26.0 −62.0 330.3+45.0 −77.0 204.3+15.0 −31.0 326.6+30.0 −26.0 34.7+3.7 −6.9 343.8+21.0 −52.0 357.1+37.0 −82.0 153.5+23.0 −15.0 92.1+27.0 −11.0 275.4+15.0 −40.0 239.3+46.0 −52.0 130.9+14.0 −30.0 1059.5+29.0 −53.0 257.0+140.0 −50.0 Orbital Parameters (2 visible transits) TABLE 2 (deg) 89.620+0.280 −0.360 89.918+0.057 −0.093 89.913+0.062 −0.083 89.894+0.073 −0.100 89.862+0.020 −0.020 89.300+0.490 −0.720 89.946+0.038 −0.062 89.931+0.049 −0.081 89.768+0.095 −0.062 89.510+0.210 −0.110 89.932+0.048 −0.075 89.861+0.096 −0.098 89.800+0.140 −0.220 89.966+0.003 −0.004 89.820+0.120 −0.064 0.036+0.003 −0.002 0.028+0.001 −0.001 0.009+0.001 −0.001 0.059+0.002 −0.002 0.068+0.002 −0.002 0.030+0.002 −0.002 0.044+0.002 −0.002 0.035+0.002 −0.002 0.041+0.001 −0.001 0.034+0.002 −0.003 0.028+0.001 −0.001 0.050+0.003 −0.003 0.049+0.003 −0.002 0.069+0.001 −0.001 0.033+0.002 −0.002 RP (R⊕) 5.1+2.2 −1.9 4.6+2.2 −2.0 1.9+0.4 −0.3 9.9+5.2 −4.7 6.0+0.8 −0.6 4.1+2.2 −1.9 4.5+0.5 −0.5 4.5+1.7 −1.6 4.6+0.6 −0.7 8.9+3.6 −3.5 3.8+0.8 −0.8 5.5+0.9 −0.8 5.4+0.9 −0.8 12.5+2.4 −2.3 3.2+0.6 −0.6 Epoch (BKJD) 448.494+0.008 −0.008 356.412+0.009 −0.008 282.995+0.009 −0.008 536.702+0.006 −0.005 382.949+0.007 −0.008 1520.292+0.006 −0.006 271.889+0.009 −0.012 481.883+0.006 −0.006 306.506+0.004 −0.004 545.741+0.014 −0.013 228.451+0.008 −0.006 335.236+0.012 −0.012 335.238+0.011 −0.012 239.224+0.001 −0.001 490.271+0.014 −0.012 0.260+0.310 −0.180 0.460+0.330 −0.310 0.320+0.360 0.410+0.320 −0.270 0.480+0.340 −0.320 0.570+0.300 −0.350 0.420+0.340 −0.280 0.590+0.280 0.450+0.330 −0.270 0.620+0.250 −0.350 0.420+0.320 0.500+0.310 −0.310 0.500+0.310 −0.310 0.650+0.230 0.460+0.360 Note. — All targets have AO imaging observations. Targets with follow-up spectroscopic observations are marked with a † . †† : Targets with detected stellar companions as reported in Table 6. The AO detection limits are given in Table 5. −0.230 −0.060+0.360 −0.400 0.410+0.340 −0.500 0.050+0.440 −0.450 0.040+0.440 −0.460 0.150+0.390 −0.400 0.030+0.440 −0.420 0.080+0.430 −0.470 0.040+0.390 −0.420 0.250+0.350 −0.280 −0.350 −0.060+0.460 −0.440 Binary Multi Multi, Validated, Q16 and Q17 data only Multi, Binary Multi, Validated Binary −0.280 −0.070+0.420 −0.420 EB, Unstable, Likely blending (§5.2) L o n g - Multi, Validated, Uncertain period (P = 427 or 854 Multi, Validated, Uncertain period (P = 427 or 854 0.140+0.410 −0.430 0.160+0.410 −0.440 −0.320 −0.050+0.460 −0.310 −0.030+0.450 −0.430 −0.330 Binary P e r i o d E x o p l a n e t s 1 7 Orbital Parameters (3-4 visible transits) TABLE 3 1 8 KIC KOI P (days) a/R∗ Inclination RP/R∗ 5437945 5652983 6436029 7619236† 8012732† 9413313† 10024862† 10850327 11465813†† 11716643† 03791 00371 02828 00682 05833 00771 05929 440.78130+0.00563 −0.00577 498.38960+0.01166 −0.01131 505.45900+0.04500 −0.04102 562.70945+0.00411 −0.00399 431.46810+0.00358 −0.00365 440.39840+0.00275 −0.00282 567.04450+0.02557 −0.02936 440.16700+0.01738 −0.01671 670.65020+0.01018 −0.01018 466.00010+0.00799 −0.00775 158.9+5.1 −12.0 215.8+29.0 −33.0 155.5+32.0 −39.0 311.9+16.0 −14.0 160.2+5.4 −4.6 352.1+7.2 −15.0 230.9+37.0 −81.0 124.9+36.0 −21.0 85.2+1.1 −1.1 380.5+24.0 −61.0 (deg) 89.904+0.066 −0.086 89.721+0.049 −0.072 89.661+0.072 −0.150 89.851+0.012 −0.011 89.741+0.018 −0.015 89.966+0.023 −0.028 89.868+0.095 −0.190 89.570+0.120 −0.100 89.535+0.013 −0.012 89.947+0.037 −0.059 0.047+0.001 −0.001 0.111+0.061 −0.057 0.047+0.012 −0.005 0.077+0.002 −0.002 0.074+0.001 −0.002 0.080+0.001 −0.001 0.046+0.003 −0.003 0.032+0.003 −0.003 0.136+0.002 −0.002 0.047+0.002 −0.002 RP (R⊕) 6.4+1.6 −1.6 35.9+24.7 −20.0 4.1+1.3 −0.7 9.9+1.9 −1.8 9.8+4.1 −3.9 12.6+7.2 −6.9 5.5+2.0 −1.6 3.5+0.7 −0.6 13.8+1.1 −1.1 4.2+0.5 −0.4 Epoch (BKJD) 139.355+0.003 −0.003 244.083+0.008 −0.008 458.092+0.035 −0.031 185.997+0.002 −0.002 391.807+0.002 −0.002 485.608+0.002 −0.002 359.666+0.017 −0.021 470.358+0.011 −0.011 209.041+0.004 −0.004 434.999+0.005 −0.005 µ1 µ2 Comments See §4.4 for details 0.320+0.180 −0.160 0.550+0.310 −0.360 0.510+0.340 −0.360 0.410+0.370 −0.280 0.560+0.290 −0.320 0.380+0.130 −0.130 0.410+0.370 −0.280 0.570+0.310 −0.370 0.420+0.260 −0.240 0.490+0.280 −0.290 −0.290 Multi, Validated Large RV variation, likely a false positive 0.290+0.270 0.000+0.420 −0.430 0.000+0.430 −0.440 Multi 0.230+0.360 −0.440 TTV 0.000+0.430 −0.360 TTV 0.450+0.200 −0.250 TTV 0.070+0.430 −0.480 TTV 0.120+0.390 −0.360 0.340+0.360 −0.380 Multi, Binary, Varying depth, Likely a false positive 0.230+0.370 −0.420 TTV Note. — All targets have AO imaging observations. Targets marked with a † are systems displaying TTVs. †† : Targets with detected stellar companions as reported in Table 6. The AO detection limits are given in Table 5. P l a n e t H u n t e r s V I I I Long-Period Exoplanets 19 TABLE 4 Stellar Parameters KIC KOI α δ (h m s) (d m s) Kp (mag) Teff (K) log g (cgs) [Fe/H] (dex) M∗ (M⊙) R∗ (R⊙) Orbital Solutions in Table 2158850 3558849 3756801 5010054† 5437945 5522786† 5536555 5652983 5732155† 5951458 6191521 6436029 7619236 8012732 8410697 8510748 8540376 8636333 9214713† 9413313 9662267† 9663113 9704149 9838291 10024862 10255705† 10403228†† 10460629 10525077 10842718 10850327 10960865 11465813 11558724 11716643 12066509 12356617 12454613† 04307 01206 03791 00371 00847 02828 00682 03349 00422 00179 01168 05800 05833 00771 05929 00375 19 24 37.875 +37 30 55.69 19 39 47.962 +38 36 18.68 19 35 49.102 +38 53 59.89 19 25 59.610 +40 10 58.40 19 13 53.962 +40 39 04.90 19 13 22.440 +40 43 52.75 19 30 57.482 +40 44 10.97 19 58 42.276 +40 51 23.36 19 53 42.132 +40 54 23.76 19 15 57.979 +41 13 22.91 19 08 37.032 +41 33 56.84 19 18 09.317 +41 53 34.15 19 40 47.518 +43 16 10.24 18 58 55.079 +43 51 51.18 18 48 44.594 +44 26 04.13 19 48 19.891 +44 30 56.12 18 49 30.607 +44 41 40.52 19 43 47.585 +44 45 11.23 19 21 33.559 +45 39 55.19 19 41 40.915 +45 54 12.56 19 47 10.274 +46 20 59.68 19 48 10.901 +46 19 43.32 19 16 39.269 +46 25 18.48 19 39 02.134 +46 40 39.11 19 47 12.602 +46 56 04.42 18 51 24.912 +47 22 38.89 19 24 54.410 +47 32 59.93 19 10 20.830 +47 36 00.07 19 09 30.737 +47 46 16.28 18 47 47.285 +48 13 21.36 19 06 21.895 +48 13 12.97 18 52 52.675 +48 26 40.13 19 46 47.666 +49 18 59.33 19 26 34.094 +49 33 14.65 19 35 27.665 +49 48 01.04 19 36 12.245 +50 30 56.09 19 24 48.286 +51 08 39.41 19 12 40.656 +51 22 55.88 10.9 14.2 13.6 14.0 13.8 9.3 13.5 12.2 15.2 12.7 15.2 15.8 13.9 13.9 13.4 11.6 14.3 15.3 14.7 14.1 14.9 14.0 15.1 12.9 15.9 12.9 16.1 14.0 15.4 14.6 13.0 14.2 15.2 14.7 14.7 14.7 13.3 13.5 6108+203 −166 6175+168 −194 5796+162 −165 6300+400 −400 6340+176 −199 8600+300 −300 5996+155 −159 5198+95 −95 6000+400 −400 6258+170 −183 5665+181 −148 4817+181 −131 5589+102 −108 6221+166 −249 5918+157 −152 7875+233 −309 6474+178 −267 6247+175 −202 6200+400 −400 5359+167 −143 6000+400 −400 6065+155 −180 5897+155 −169 6123+141 −177 6616+169 −358 5300+300 −300 3386+50 −50 6449+163 −210 6091+164 −213 5754+159 −156 6277+155 −187 5547+196 −154 5520+83 −110 6462+177 −270 5830+155 −164 6108+149 −192 5755+112 −112 5500+280 −280 −0.30 −0.30 −0.26 −0.30 −0.60 −1.96+0.34 −0.27 −0.42+0.28 −0.22 −0.02+0.24 −0.28 0.02+0.22 −0.28 −0.25 −0.38+0.28 −0.30 0.07+0.14 −0.59 −0.28 −0.48+0.30 −0.26 −0.30 −0.32 4.48+0.14 4.44+0.07 4.12+0.26 4.30+0.50 −0.50 4.16+0.22 4.20+0.20 −0.20 4.49+0.06 3.61+0.02 −0.02 4.20+0.50 4.08+0.28 4.56+0.05 4.50+0.08 −0.84 4.23+0.13 −0.12 4.29+0.12 −0.38 4.37+0.14 3.70+0.28 −0.10 4.31+0.10 4.49+0.04 4.40+0.50 4.40+0.13 −0.39 4.50+0.50 4.42+0.08 4.53+0.03 4.47+0.05 4.33+0.08 −0.31 3.80+0.40 4.92+0.06 −0.07 4.23+0.16 4.42+0.06 4.38+0.12 4.43+0.07 4.05+0.34 −0.26 4.47+0.04 −0.14 4.32+0.10 4.54+0.03 4.47+0.04 −0.30 4.10+0.14 −0.13 4.60+0.30 −0.30 −0.50 −0.04+0.22 −0.23 −0.50+0.30 −0.27 −0.58+0.34 −0.26 0.42+0.06 −0.24 0.34+0.10 −0.14 0.20+0.16 −0.32 −0.24 −0.42+0.30 −0.26 0.04+0.17 −0.38 −0.33 −0.16+0.23 −0.27 −0.34+0.26 −0.50 −0.30+0.26 −0.30 0.02+0.28 −0.26 −0.50 −0.06+0.22 −0.26 −0.28+0.28 −0.28 −0.16+0.24 −0.29 −0.14+0.22 −0.30 0.07+0.19 −0.39 −0.40 −0.12+0.33 −0.30 0.00+0.10 −0.10 −0.27 −0.32+0.24 −0.30 −0.04+0.22 −0.24 −0.06+0.26 −0.28 −0.46+0.28 −0.30 0.02+0.26 −0.26 0.48+0.08 −0.16 −0.35 −0.08+0.22 −0.28 −0.14+0.24 −0.28 0.07+0.15 −0.33 0.24+0.14 −0.14 0.00+0.24 −0.24 −0.30 −0.30 −0.30 · · · −0.30 −0.30 −0.26 −0.32 [0.72..0.96] [0.87..1.09] [0.89..1.19] [0.87..1.32] [0.90..1.24] [1.86..2.19] [0.69..0.98] [1.13..1.61] [0.87..1.33] [0.77..1.19] [0.77..0.92] [0.79..0.88] [0.93..1.12] [0.77..1.07] [0.74..1.08] [1.36..2.40] [0.84..1.23] [0.86..1.03] [0.84..1.17] [0.72..1.17] [0.88..1.21] [0.85..1.10] [0.73..0.99] [0.76..1.08] [0.89..1.24] [0.98..1.40] [0.27..0.37] [0.94..1.29] [0.89..1.13] [0.74..1.12] [0.87..1.10] [0.73..1.19] [0.88..1.03] [0.81..1.22] [0.79..0.93] [0.80..1.11] [0.98..1.25] [0.82..1.00] [0.63..1.54] [0.90..1.11] [0.87..1.75] [0.87..2.13] [0.95..1.53] [1.63..2.12] [0.67..1.38] [2.70..3.23] [0.82..2.25] [0.70..2.34] [0.75..0.88] [0.75..0.84] [0.98..1.36] [0.75..1.69] [0.66..1.85] [1.20..4.23] [0.70..1.82] [0.87..1.01] [0.79..1.66] [0.66..2.24] [0.79..1.52] [0.91..1.15] [0.67..1.02] [0.72..1.42] [0.82..1.39] [1.62..3.23] [0.28..0.38] [1.02..1.47] [0.91..1.11] [0.65..1.90] [0.90..1.10] [0.62..2.42] [0.87..0.99] [0.70..1.80] [0.77..0.87] [0.76..1.32] [1.39..1.96] [0.77..1.00] 1 1 2 1,2 3 2 1 3 2 1 2 3 3 3 1 1 1,2 2 3 2 2 2 1 1 1,3 2 1 2 2 1 3 1 3 1 3 1 2 2 Note. — Targets with follow-up spectroscopic observations are marked with an †. Their stellar properties are based on MOOG analysis. We report 1-σ range for stellar mass and radius. †† : Stellar mass and radius are adopted from Huber et al. (2014). 20 Planet Hunters VIII TABLE 5 AO Sensitivity to Companions KIC KOI Kmag [mag] Kepler i J H K [mag] [mag] [mag] [mag] Observation Limiting Delta Magnitude Companion Instrument Filter within 5′′ 0.1 [′′] 0.2 [′′] 0.5 [′′] 1.0 [′′] 2.0 [′′] 4.0 Orbital [′′] in 2158850 3558849 3756801 5010054 5010054 5437945 5437945 5522786 5522786 5536555 5652983 5732155 5732155 5951458 6191521 6436029 7619236 8012732 8410697 8510748 8540376 8540376 8636333 8636333 9413313 9413313 9662267 9662267 9663113 9704149 9704149 9838291 10024862 10024862 10255705 10255705 10255705 10255705 10460629 10525077 10850327 11465813 11465813 11465813 11716643 12356617 12454613 12454613 04307 01206 03791 03791 00371 00847 02828 00682 03349 03349 00179 01168 05800 05833 00771 00771 00771 05929 00375 10.863 14.218 13.642 13.961 13.961 13.771 13.771 9.350 9.350 13.465 12.193 15.195 15.195 12.713 15.201 15.768 13.916 13.922 13.424 11.614 14.294 14.294 15.292 15.292 14.116 14.116 14.872 14.872 13.955 15.102 15.102 12.868 15.881 15.881 12.950 12.950 12.950 12.950 13.997 15.355 13.014 15.207 15.207 15.207 14.692 13.293 13.537 13.537 10.726 14.035 13.408 13.710 13.710 13.611 13.611 9.572 9.572 13.285 11.895 14.978 14.978 12.556 14.970 15.369 13.692 13.727 13.238 11.638 14.151 14.151 15.113 15.113 13.835 13.835 14.667 14.667 13.765 14.897 14.897 12.703 15.712 15.712 12.678 12.678 12.678 12.678 13.851 15.163 12.872 15.068 15.068 15.068 14.485 13.111 13.306 13.306 9.855 13.092 12.439 12.797 12.797 12.666 12.666 9.105 9.105 12.313 10.723 14.006 14.006 11.640 13.935 14.041 12.688 13.094 12.281 10.967 13.259 13.259 14.192 14.192 12.733 12.733 13.670 13.670 12.823 13.896 13.896 11.826 14.846 14.846 11.560 11.560 11.560 11.560 12.923 14.143 11.993 13.678 13.678 13.678 13.483 12.137 12.326 12.326 9.570 12.819 12.099 12.494 12.494 12.429 12.429 9.118 9.118 11.971 10.289 13.705 13.705 11.381 13.585 13.506 12.378 12.794 11.933 10.898 13.013 13.013 13.890 13.890 12.335 12.335 13.385 13.385 12.545 13.538 13.538 11.548 14.551 14.551 11.105 11.105 11.105 11.105 12.672 13.868 11.711 13.317 13.317 13.317 13.095 11.842 11.929 11.929 9.529 12.766 12.051 12.412 12.412 12.367 12.367 9.118 9.118 11.933 10.169 13.621 13.621 11.323 13.569 13.429 12.260 12.790 11.922 10.861 12.965 12.965 13.880 13.880 12.227 12.227 13.339 13.339 12.502 13.454 13.454 11.496 14.541 14.541 11.021 11.021 11.021 11.021 12.595 13.753 11.666 13.253 13.253 13.253 13.092 11.791 11.867 11.867 no no no no no no no no no no no yes no no no no no no no yes no no yes yes no no no no no no no no no no yes yes yes yes no no no yes yes yes no yes no no NIRC2 NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 Robo-AO NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 Robo-AO NIRC2 NIRC2 NIRC2 NIRC2 NIRC2 NIRC2 NIRC2 PHARO NIRC2 Robo-AO KS KS KS KS i J KS KS i KS KS KS i KS KS KS KS i KS KS KS i H KS KS i KS i KS KS i KS KS i H J KS i KS KS KS H J KS KS KS KS i 3.8 3.3 2.1 2.0 0.2 2.4 2.8 1.5 0.0 3.6 2.2 2.1 0.6 3.8 2.0 1.2 1.8 0.2 3.9 4.9 4.0 0.2 1.0 1.0 1.9 0.4 1.6 0.4 2.2 1.7 0.6 4.3 0.7 0.8 2.5 2.0 2.4 0.2 2.1 1.9 2.4 1.0 0.9 1.5 2.1 0.1 2.0 0.3 3.8 3.3 4.2 3.9 0.5 3.5 4.6 4.6 0.4 3.6 4.2 3.5 0.9 3.8 3.7 2.7 3.8 0.5 4.6 4.9 4.0 0.4 2.0 2.3 3.7 0.7 3.4 0.6 3.9 3.2 0.8 4.3 1.8 1.1 4.0 3.3 4.2 0.3 3.7 3.4 4.1 2.5 2.0 3.2 3.4 0.8 4.2 0.4 5.9 5.4 5.2 4.9 2.3 5.3 6.3 5.9 2.7 5.3 5.5 4.8 2.2 5.9 4.9 4.4 4.8 1.8 6.5 6.5 4.3 1.9 3.6 3.9 4.8 2.4 4.9 2.4 4.8 4.4 2.1 6.4 2.2 2.0 5.8 5.2 5.6 1.7 4.7 4.5 5.1 4.7 4.2 4.1 4.2 2.5 5.3 1.8 6.7 6.9 5.3 5.0 3.8 5.7 6.9 6.5 4.6 5.6 5.7 4.9 3.2 6.7 5.1 4.7 5.0 3.5 7.0 6.8 4.3 3.4 4.8 4.6 4.9 3.5 5.0 3.2 4.9 4.7 3.1 7.2 2.2 2.6 6.1 5.6 5.7 3.1 4.8 4.8 5.4 5.2 4.6 4.2 4.3 4.0 5.4 3.4 6.7 7.0 5.3 4.9 4.6 5.7 7.0 6.5 6.9 5.5 5.8 5.0 3.4 6.5 5.1 4.7 5.0 4.3 7.1 6.8 4.3 4.1 5.0 4.7 4.9 3.9 5.1 3.8 4.9 4.8 3.5 7.4 2.2 2.8 6.3 5.7 5.8 4.6 4.7 4.8 5.4 5.0 4.6 4.3 4.3 4.9 5.5 4.5 6.7 6.8 5.3 4.9 4.7 5.7 6.9 6.5 8.0 5.5 5.7 5.0 3.6 6.6 5.1 4.6 4.9 4.3 7.1 6.6 4.3 4.2 5.0 4.6 4.9 3.9 5.1 3.8 4.9 4.7 3.5 7.2 2.3 2.8 6.2 5.7 5.8 5.0 4.7 4.8 5.3 5.2 4.7 4.3 4.3 5.0 5.5 4.7 Long-Period Exoplanets 21 TABLE 6 AO Detections KIC KOI Kmag [mag] Kepler i J H K [mag] [mag] [mag] [mag] sep. P.A. [′′] [deg] Companion ∆i ∆J ∆H ∆K [mag] [mag] [mag] [mag] 5732155 8510748 8636333 10255705 11465813 12356617 03349 00771 00375 15.195 11.614 15.292 12.950 15.207 13.293 14.978 11.638 15.113 12.678 15.068 13.111 14.006 10.967 14.192 11.560 13.678 12.137 13.705 10.898 13.890 11.105 13.317 11.842 13.621 10.861 13.880 11.021 13.253 11.791 0.93 0.17 0.32 1.06 1.77 3.10 221.1 111.9 266.2 164.1 282.7 305.4 1.94 2.27 0.77 1.58 2.37 0.74 4.94 3.13 1.71 2.40 0.65 4.42 Note. — Typical uncertainties for companion separation (sep.), position angle (P. A.) and differential magnitude ∆Mag are 0′′ .05, 0◦.5 and 0.1 mag. The uncertainties are estimated based on companion injection simulation (Wang et al. 2015). 22 Planet Hunters VIII TABLE 7 Planet Confidence KIC Epoch Depth Offset BKJD (ppm) (mas) δ Offset Significance #Planet (mas) (σ) Candidates RP (R⊕) Confidence Comment Orbital Solutions in Table 2158850 3558849 3756801 5010054 5010054 5437945 5522786 5536555 5536555 5652983 5732155 5951458 6191521 6436029 7619236 8012732 8410697 8510748 8540376 8540376 8636333 9413313 9662267 9663113 9704149 9838291 10024862 10024862 10255705 10460629 10525077 10850327 11465813 11716643 12356617 12454613 411 279 448 1500 356 139 282 370 492 244 536 423 382 458 185 391 542 1536 1516 1520 271 485 481 306 419 582 878 359 545 228 335 470 209 434 239 490 169 3969 1296 441 784 2209 81 576 576 12321 3481 1600 4624 2209 5929 5476 5184 144 324 900 1936 6400 1225 1681 2916 1849 9604 2116 1156 784 2500 1024 18496 2209 4761 1089 0.2 1.6 0.7 0.7 0.3 0.3 0.3 0.7 0.5 0.1 5.0 0.1 0.7 7.0 1.6 1.1 0.4 0.2 1.0 0.2 0.9 1.5 0.5 0.6 1.7 0.3 2.9 0.2 0.9 0.1 1.7 0.2 1.0 1.7 0.5 1.2 0.3 2.1 1.3 0.9 0.8 0.8 0.3 1.3 0.9 1.1 4.1 0.6 2.8 10.6 1.2 1.6 1.4 0.3 1.5 1.6 2.1 1.5 1.7 1.2 3.3 0.7 5.1 5.4 0.7 1.5 3.8 0.6 3.3 2.8 1.2 1.8 0.5 0.8 0.6 0.8 0.4 0.4 1.1 0.5 0.6 0.1 1.2 0.1 0.3 0.7 1.3 0.7 0.3 0.7 0.7 0.1 0.4 1.0 0.3 0.5 0.5 0.4 0.6 0.0 1.3 0.0 0.5 0.4 0.3 0.6 0.4 0.7 1 1 1 1 1 2 1 1 1 2 1 1 1 2 1 1 1 1 3 3 2 1 1 2 1 1 1 1 1 1 2 1 1 1 1 1 1.6 6.9 5.1 3.4 4.6 6.4 1.9 2.7 2.7 35.9 9.9 6.6 6.0 4.1 9.9 9.8 9.8 3.6 2.4 4.1 4.5 12.6 4.5 4.6 5.0 5.1 11.8 5.5 8.9 3.8 5.5 3.5 13.8 4.2 12.5 3.2 0.946 0.997 0.980 0.984 0.883 0.999 0.960 0.922 0.922 · · · · · · 0.998 0.966 0.955 0.984 0.972 0.996 · · · 0.999 0.999 · · · 0.983 0.961 0.999 0.965 0.992 0.730 0.304 · · · 0.973 0.998 0.977 · · · 0.708 · · · 0.960 Validated Validated Large RV and radius Nearby Companion Validated Nearby Companion Validated Validated Nearby Companion Validated Nearby Companion Validated Nearby Companion Nearby Companion 1 1 2 1 2 3 2 1 1 3 2 1 2 3 3 3 1 1 1 2 2 3 2 2 1 1 1 3 2 2 2 3 3 3 2 2
0911.2196
1
0911
2009-11-11T18:13:11
Effects of Gamma Ray Bursts in Earth Biosphere
[ "astro-ph.EP", "astro-ph.HE", "q-bio.PE" ]
We continue former work on the modeling of potential effects of Gamma Ray Bursts on Phanerozoic Earth. We focus on global biospheric effects of ozone depletion and show a first modeling of the spectral reduction of light by NO2 formed in the stratosphere. We also illustrate the current complexities involved in the prediction of how terrestrial ecosystems would respond to this kind of burst. We conclude that more biological field and laboratory data are needed to reach even moderate accuracy in this modeling
astro-ph.EP
astro-ph
GRB’s in Earth’s Biosphere Original Research Effects of Gamma Ray Bursts in Earth’s Biosphere Osmel Martin1, Rolando Cardenas2, Mayrene Guimaraes3, Liuba Peñate4, Jorge Horvath5, Douglas Galante6 1,2 Department of Physics, Universidad Central de Las Villas, Santa Clara, Cuba. Phone 53 42 281109 Fax 53 42 281109. e-mail: 1 [email protected], 2 [email protected] 3 Marine Ecology Group, Center for Research of Coastal Ecosystems, Cayo Coco, Ciego de Avila, Cuba Phone 53 33 301104 ext 117 e-mail: [email protected] 4Department of Biology, Universidad Central de Las Villas, Santa Clara, Cuba. Phone 53 42 281109 Fax 53 42 281109 e-mail: [email protected] 5, 6 Department of Astronomy, Instituto de Astronomia, Geofísica e Ciências Atmosféricas, Universidade de São Paulo, São Paulo, Brazil. e-mail: 4 [email protected]; 5 [email protected] Abstract: We continue former work on the modeling of potential effects of Gamma Ray Bursts on Phanerozoic Earth. We focus on global biospheric effects of ozone depletion and show a first modeling of the spectral reduction of light by NO2 formed in the stratosphere. We also illustrate the current complexities involved in the prediction of how terrestrial ecosystems would respond to this kind of burst. We conclude that more biological field and laboratory data are needed to reach even moder ate accuracy in this modeling. 1 Introduction The idea of strong astrophysical influence in the course of the Earth‟s biological evolution has been discussed by several authors. According with that, large asteroid impacts, giant solar flares, supernovae explosions (SNe) or Gamma Ray Bursts (GRB‟s), could have acted as triggers of extinctions in the Earth‟s geological past. For instance, recently it has been suggested a connection between supernovae and the extinction of tropical American mollusks that took place around the Pliocene - Pleistocene boundary (Benitez, Maiz-Apellaniz & Canelles, 2002). Additionally, the potential influence of SNe is used to define concepts such as Galactic Habitable Zone, in a more astrobiological context (Gonzalez, Brownlee & Ward, 2001). For the case of transient radiation events like SNe or GRB‟s, beyond differences between them, their main influences on the Earth‟s biota could be similar. Both events are strong sources of highly energetic gamma radiation, capable of inducing severe perturbations on the chemistry of planetary atmospheres. Their main effects on the Earth‟s biota are strongly dependent on the atmospheric composition, the presence or not of an active O2/O3 ultraviolet radiation (UVR) blocking system, and ecosystem specificities. They are associated to an increase of typical UV levels reaching the ground at least in two forms: the so called reemitted UV flash and the increase of solar UV by depletion of the ozone layer (Galante & Horvath, 2007). The relative importance of these effects appears to be a strong function of the free oxygen content in the atmosphere. For contemporary Earth-like atmospheres (rich in O2), the main influence is the depletion of the ozone layer through the catalytic effect of NOx species formed during the burst. The total recovery is determined mainly by the atmospheric chemistry and the 1 transport processes. The time for the recovery of the ozone is around a decade (Thomas et al, 2005). Other potential influences on the climate and biosphere may be induced by abnormal nitrate deposition due to rainout of NOx in the form of nitric acid rain or cooling effects and reduction of sunlight in the visible range due to high NO2 levels. Modeling the action of an UVR excess on Earth‟s biosphere is a highly complicated task, given the variability in species sensitivity, possible thresholds, and non-linearities. We can restrict our consideration to the so called primary producers of the biosphere (phytoplankton, algae, higher plants), as they form the basis of the food web , so any perturbation on them should be reflected in higher levels of the trophic assemblage (herbivores, carnivores, omnivores). As the biosphere contributes to the CO2 fixation and O2 evolution, important perturbations on it by an UVR excess, coming from any source (solar or extrasolar), have the potential for global climate changes. However, the biosphere is formed by many interacting ecosystems, whose respective responses to UVR excess is even today an open question (Thomas, 2008; Hader et al, 2007). In this work, we continue exploring global and regional effects that a GRB could cause on Earth‟s biosphere, due to the aforementioned ozone depletion and enhanced atmospheric opacity due to the formation of NO2 as consequences of the burst impact. 2 Basic assumptions 2.1 The effects on the atmosphere Ionizing radiation dissociates N2 and O2 in the atmosphere, releasing important quantities of atomic nitrogen and oxygen. These very reactive chemical species then form considerable quantities of nitrogen oxides, catalizers of the ozone dissociation. As stated in (Thomas et al. 2005), the „„typical‟‟ nearest burst in the last billion years would cause a globally averaged ozone depletion of up to 38% and significant global depletion (at least 10%) would persist up to seven years. This would imply: - an enhanced irradiation of the planet‟s surface with the solar ultraviolet radiation (UVR), atmospheric opacity reducing visible sunlight in a few percent because of the formation of NO2, with potential global cooling, and, - deposition of nitrate through rainout of nitric acid, slightly greater than that currently caused by lightning, lasting several years. In order to account for the spectral reduction of irradiance at planet surface due to the formation of NO2 we used the solar spectrum at surface as given in (ASTM G173 - 03e1). Then, considering that in (Thomas et al. 2005) total irradiance reduction in the range (0-10) % due to the formation of NO2 is reported, we calculated which columns of NO2 would make reductions of total irradiance I given by several values of the fraction number f: - (1) where from now on the subscripts after and before mean after and before the impact of the GRB. We used the values for f of 0.98, 0.96, 0.94 and 0.92, representing irradiance reductions of 2, 4, 6 and 8 % respectively. The values of total irradiances after and before the burst are given by: 2 )(0IfIIbeforeafter 𝜆 (3) 𝜆 (2) 700 𝑛𝑚 𝐼𝑎𝑓𝑡𝑒𝑟 = 280 𝑛𝑚 700 𝑛𝑚 𝐼𝑏𝑒𝑓𝑜𝑟𝑒 = 280 𝑛𝑚 𝐼0 𝜆 𝑑 𝐼0 𝜆 𝑒 −𝜏 𝑑 where 𝜏 is the optical path of photons in the NO2 column. This magnitude gives the clue to estimate the quantity of NO2 needed to reduce the total irradiance in a given f. The above procedure neglects the increase of irradiance due to ozone depletion, but as the Sun peaks in the visible part of the spectrum, that contribution to the total irradiance 𝐼𝑎𝑓𝑡𝑒𝑟 is very small, something that we checked using the radiative transport code NCAR/ACD TUV: Tropospheric Ultraviolet & Visible Radiation Model (NCAR/ACD). 2.2 Estimation of wide-scale damage on the biosphere It is clear that the first and second atmospheric effects mentioned above could affect many photosynthetic species: more solar UV can damage DNA and inhibit photosynthesis to some extent, while less visible sunlight (i. e., photosynthetic active radiation, PAR) would reduce the energy available for photosynthesis and therefore for primary production. However, the third effect can offset, at least partially, the above mentioned inhibition of photosynthesis, and could even cause eutrofication (over-enrichment of nutrients) in some freshwater and coastal ecosystems. It is true that the nitric acid rain could stress portions of the biosphere, but, after titration, the increased nitrate deposition could be helpful to photosynthetic organisms, especially to land plants. This effect requires further attention and is not a focus in this paper. Due to the considerable variability in species sensitivity to radiations and to non linearities, the accurate modeling of how the biosphere would behave in excess of UVR is very complicated. However, a rough idea of the biological effects of ozone depletion is the radiation amplification factor (RAF), relating the biological effective irradiances E* with the ozone columns N, after and before the ionizing event: 𝑅𝐴𝐹 ∗ 𝐸𝑎𝑓𝑡𝑒𝑟 ∗ 𝐸𝑏𝑒𝑓𝑜𝑟𝑒 (4) 𝑁𝑏𝑒𝑓𝑜𝑟𝑒 = 𝑁𝑎𝑓𝑡𝑒𝑟 The RAF‟s are dependent both upon the group of species and upon the organismal process to be considered (represented by a biological weighting function BWF). BWF‟s are typically measured in controlled laboratory conditions; so they are of limited value to estimate the actual response of living beings to UVR. Under the action of UVR, organisms can enzymatically reverse the photochemical reaction or re- synthesize the affected molecules. These processes, generically known as repair, depend not only on the species, but also on environmental variables. For instance, it is well known the interaction repair – temperature for several species of phytoplankton: at very low temperatures repair is very slow, while at intermediates temperature repair is good. In general, repair is not properly taken into account when BWF‟s are measured; therefore the biological amplification factor (BAF) is the quantity which would give us more accurate information on the biological effects of UVR: 𝑃𝑎𝑓𝑡𝑒𝑟 𝑃𝑏𝑒𝑓𝑜𝑟𝑒 = 𝐵𝐴𝐹 × ∗ 𝐸𝑏𝑒𝑓𝑜𝑟𝑒 ∗ 𝐸𝑎𝑓𝑡𝑒𝑟 (5) 3 where P is the rate of an organismal process (for example, photosynthesis). Unfortunately, very few BAF‟s have been measured, though the alternative exposure - response curves (ERC‟s) for several species have been reported. Anyway, RAF‟s and BAF‟s could be useful for a first rough approach to estimate global damage on the biosphere of a Gamma Ray Burst, but a more detailed modeling implies to look at specific ecosystems, the building blocks of the biosphere. 2.3 Gamma Ray Bursts at ecosystem level From the broadest biophysiological point of view, the biosphere is the global ecological system integrating all living beings and their relationships, including their interaction with the elements of the lithosphere, hydrosphere, and atmosphere. The biosphere can also be considered as the sum of all ecosystems (aquatic, terrestrial and hybrid). Studies of ecosystems usually focus on the movement of energy and matter through the system, but these processes will depend on the kind of ecosystem. However, some generic characteristics can be stated: - On energy: Almost all ecosystems run on energy captured from the Sun by primary producers (phytoplankton, algae, higher plants) via photosynthesis, this energy then flows through the food chains to primary consumers (herbivores, who eat and digest the plants), and on to secondary and tertiary consumers (either carnivores or omnivores). - On matter: It is incorporated into living organisms by the primary producers. Photosynthetic plants fix carbon from carbon dioxide and nitrogen from atmospheric nitrogen or nitrates present in the soil to produce amino acids. Much of the carbon and nitrogen contained in ecosystems is created by such plants, and is then consumed by secondary and tertiary consumers and incorporated into themselves. Nutrients are usually returned to the ecosystem via decomposition. The entire movement of chemicals in an ecosystem is termed a biogeochemical cycle, and includes the carbon and nitrogen cycle. To study the effects of GRB‟s at regional or local level implies modeling the action of UVR excess on several different ecosystems. In this work we have chosen lakes, one of the reasons being that the selected model of lake successfully describes the process of eutrofication (over enrichment by nutrients, primarily nitrogen and phosphorus) and it has been predicted that one of the atmospheric effects of a GRB would be an increased rainout of nitrogen compounds, thus contributing to the eutrofication of terrestrial ecosystems (Thomas et al 2005). We expect then that our results might be a rough proxy of what could happen in a considerable proportion of inland waters and coastal ecosystems after the incidence of the UVR perturbation, because many of these systems often show some degree of eutrofication due to the influence of land masses. We admit that a more accurate modeling of the action of an UVR excess at ecosystem level would require specific models for other specific ecosystems, both aquatic and terrestrial, something which we leave for future work. 2.3.1 The Comprehensive Aquatic Simulation Model The Comprehensive Aquatic Simulation Model (CASM) has successfully described the key features of the eutrophication process in real lakes (Amemiya et al, 2007). This 4 process is associated to the over enrichment by nutrients, primarily phosphorus and nitrogen, with a consequent increase of phytoplankton levels, while other species such as fish and zooplankton become rather scarce. As we said in the above subsection, eutrofication by nitrate deposition is one of the potential consequences of a GRB striking our atmosphere (Thomas et al, 2005), making this model attractive for our purposes. In this model there is an external input IN of the limiting nutrient N to the ecosystem, which in our case would include the atmospheric deposition of nitrates after the GRB by rainout. Equation (6) below represents the dynamics of nutrients in the ecosystem, where rN is the loss rate of nitrogen by diverse causes (for instance, sedimentation, flow out, etc.), while the third term of right hand side (rhs) models the consumption of nutrients by the primary consumers (phytoplankton X). The form of this term is inspired in the Michaelis- Menten kinetics, firstly applied to simple processes in which enzymes participate. In our case, γ is the ratio of nutrient mass (nitrate mass) to biomass, r1 is the maximum growth rate of phytoplankton and k1 is a half saturation constant (when N = k1, the whole term will be divided by two after cancelling N, hence the denomination half saturation). Finally, the fourth term of rhs of eq. (6) represents the input of nutrient N, via decomposition of detritus matter D, considering that d4 is the decomposition rate of D. (6) The primary production of the ecosystem is represented by equation (7) below, where phytoplankton X consumes nutrients via the first term of rhs (compare it with the third term of rhs on the equation (6)), and the second term shows how zooplankton Y predates on phytoplankton. In this term, f1 is the feeding rate of zooplankton and k2 is the half saturation constant for this term (because when X2= k2, the cancellation of X2 ensures that the whole term is divided by two). The last term of rhs of the equation contains the mortality d1 of phytoplankton and its removal rate from the ecosystem e1. (7) Equation (8) below represents the dynamics of the primary consumer, zooplankton Y. The first term of rhs shows how it predates on phytoplankton (compare it with the second term of rhs of above equation), while the third term says how zooplankton is eaten by the secondary consumer, the zooplanktivorous fish Z. The parameter η represents the assimilation efficiency of zooplankton, the meanings of the other parameters can readily be deduced from the explanations given of the first two equations. (8) The dynamics of the secondary consumer, the zooplanktivorous fish, is given by equation below. Here the new parameter Z*, the low equilibrium biomass of zooplanktivorous fish, avoids the unrealistic situation of former versions of CASM, in which fish could appear from states in which it was already extinct. 5 DdNkNXrrIdtdNNN411XedXkYXfNkNXrdtdX11222111YedYkZYfXkYXfdtdY2223222221 (9) Finally, we should consider that there are sources of detritus matter D in the ecosystem (fecal material and dead X, Y and Z), whose decomposition returns nutrients to the ecosystem. This is very important in all ecosystems: an important fraction of nutrients is returned to the ecosystem via decomposition of feces and dead beings, as stated in equation below: (10) We remind that di are death or decomposition rates and ei are removal rates from the system. As can be seen from equations (6)-(10), CASM has five dynamical variables and 19 parameters. In general, we refer the interested reader to (Amemiya et al, 2007) for more details. 2.3.2 The inclusion of radiative transport in the Comprehensive Aquatic Simulation Model The formulation of CASM model above does not take into consideration the vertical distribution of the living species in the water column. This is an important omission when considering any situation of UVR stress, given the attenuation of radiation due to the phenomena of absorption and dispersion in the water column. To account for this we considered phytoplankton to be the only trophic level stressed by UVR, as they are obligated to have an adequate solar exposure in order to perform photosynthesis. We can then imagine all phytoplankton living in an effective depth and receiving increased UVR levels after a GRB. Thus, to include the role of some components of the ecosystem as UV screeners in the water column (detritus and phytoplankton themselves), we modified the CASM model considering the mortality rate coefficient of phytoplankton (d1) no longer a constant, but an explicit function of such components of the form The above exponential dependence is motivated by the well known Beer‟s law for the are coefficients for UVR absorption of light by any liquid solution, and attenuation by phytoplankton and detritus matter, while is the lethality rate coefficient of the phytoplankton when no UV blocking effect is considered. 3 Results and General Discussion 3.1 Global damage: the biosphere level 3.1.1 The effects of ozone depletion As mentioned above, in (Thomas et al. 2005) it is shown that the typical nearest burst in the last billion years would cause an averaged global ozone depletion of up to 38%, (11) 6 *332į22ZZedYkZYfdtdZ,11443212į3222221DedZdYdXdYkZYfXkYXfdtdDdedDhXhDX1XhDhd 0.31 1.27 1.20 1.12 1.05 0.51 RAF 1.67 - 2.2 2.22 - 2.85 1.82 – 2.20 1.45 – 1.63 1.19 – 1.26 30 1.12 20 1.07 10 1.03 ∗ 𝐸𝑎𝑓𝑡𝑒𝑟 ∗ 𝐸𝑏𝑒𝑓𝑜𝑟𝑒 38 1.16 for several values of ozone depletion (%) which would persist several years. For instance, seven years after the burst, 10% ozone depletion would be expected. Considering this, in Table 1 below we show the fractional increase of the effective biological irradiances for several values of ozone depletion and several biological weighting functions. Biological Weighting Function Photoinhibition of a marine phytoplankton Photoinhibition of land plants DNA damage Table 1 Radiation Amplification Factors and fractional increase of effective biological irradiances for some biological weighting functions and for ozone depletions of 38, 30, 20 and 10 % The table above suggests that DNA damage is in general the main influence of a GRB over the biosphere and that land plants might suffer more than phytoplankton. However, it should be noticed that RAF‟s are typically measured in controlled conditions very different from the natural conditions in which organisms live. Therefore, the use of biological amplification factors (BAF‟s) or exposure response curves (ERC‟s) should give us a much better picture of the response of the biosphere to UVR perturbations. Unfortunately, very few BAF‟s or ERC‟s have been measured for the most common primary producers in the biosphere, such as the main species of marine phytoplankton. Therefore, we are lacking biological field data to make more accurate accounts of the potential global effects of a GRB on the biosphere. The good news is that several studies are now underway which will supply useful biological data, therefore the next future looks promissory. 3.1.2 The effects of irradiance reduction due to NO2 formation We followed the methodology explained in subsection 2.1 to calculate the spectral reduction of light as a consequence of the enhanced formation of NO2. The Table 2 below shows a slightly selective absorption in the visible band (PAR), while a more pronounced absorption in the UV-A band and in the photorepair band (350-450nm) appears. The photorepair light is needed to execute the most efficient repair pathway of DNA damages caused by UV-B. f UV-A f 0.98 0.92 0.96 0.85 0.94 0.78 0.92 0.71 f PAR 0.98 0.95 0.93 0.90 f 350-450 nm 0.92 0.84 0.77 0.70 7 Table 2 Ratio of irradiances after and before the burst (eq. 1), both for global irradiance and for some bands We additionally checked that 30% depletion of the standard ozone column of 340 Dobson units implied a 22% increase of UV-B, but only a 0.37 % increase of UV-A, therefore the ozone depletion contribution to the increase of UV-A is much smaller than the decrease of this band due to NO2 formation, which in this case depletes light in around 10 %. Thus, the net global biological effect of a GRB suggests a combination of more damages due to more UV-B reaching the ground (because of ozone depletion) and a less efficient repair of DNA damages because less light in the photorepair band (350nm–450 nm) reaches the ground. Also, less light (PAR) would be available for photosynthesis. Additionally, the total reduction of sunlight in the percents stated in this work has the potential of global cooling, something which per se deserves considerable future investigation. 3.2 Regional damage: the ecosystem level As stated in subsection 2.3, to take into account the combined effects of the depletion of the ozone layer and the (spectral) reduction of sunlight, our modification of the CASM model for lakes was explored with increments of the mortality rate coefficient of phytoplankton (d1). In Figure 1 it is shown how the qualitative behavior of the model changes as a function of the parameter (d1). Fig. 1: Bi-stability appears for mortality rates of phytoplankton (d1) in the approximate range (0.095 – 0.125) day-1 (solid lines represent stable or oscillatory states, dashed 8 lines transient ones). The Hopf bifurcation at d1= 0.105 day-1 marks the transition from a stable state to an oscillatory one. When the parameter d1 increases to d1=0.105, only a 5 percent above the referenced value d1= 0.1 in (Amemiya et al, 2007), the steady clear state emerges as an oscillating state. At higher values (around d1 = 0.125), the bi-stability of the system is broken and the oscillating state emerges as a unique possibility. Such alternative states are exhibited by CASM for other parameter regions (Amemiya et al, 2007). Radiative transport analysis in oscillating regimes appears interesting because the optical properties of the water column are continually varying in the time. Some components as detritus matter (D) and phytoplankton (X) play additional UV protection to the main underwater species. Taking into account our modified expression for the mortality rate coefficient (eq. 11) and equal contributions to the attenuation of UV photons by phytoplankton and by detritus (h = hX = hD), we found the behavior shown in Figure 2. Fig. 2 Self protection from UV photons of detritus and phytoplankton could cause regime shift or oscillations around the clear state Now, according with the information in Figure 2, if self-protection is not too high, the oscillating regime around the clear state persists, with minor corrections in the amplitude and oscillation period. If the self-protection reaches some threshold value, phytoplankton population recoveries progressively in time and the ecosystem comes back to the original turbid state. The above results illustrate the complexities involved in predicting how terrestrial ecosystems would recover if stressed by a GRB. The status towards which a given lake 9 will evolve might depend on several variables and parameters, but phytoplankton, being the primary producer, would play the determinant role. However, a more accurate modeling of the recovery of aquatic ecosystems after a GRB needs a closer look at the behavior of the more common species of phytoplankton under UVR stress, and also other environmental variables are probably to be taken into consideration. 4 Conclusions Given non-linearity and variability in the response of biological systems to radiations, it is difficult to predict the damage and recovery of the biosphere under the impact of the „„typical‟‟ nearest GRB in the last billion years. In this work we have estimated some global effects on the biosphere, but the lack of data on biological amplification factors for the more abundant species of primary producers actually limit the predictive power of present studies. However, largely motivated by today‟s ozone depletion, some researchers are currently making studies on the response of the most abundant primary producers to UVR; therefore soon we should be able to make more detailed modeling on the potential global biological effects of a GRB. On the ecosystem (regional) scale a similar situation holds, but again we are optimist concerning the next arrival of new field data. This could serve as a discriminating tool to reveal towards which state several terrestrial ecosystems would shift their equilibrium after the action of a nearby GRB. Acknowledgments The authors thank to the Brazilian federal organization CAPES for financial support of the project “Influence of cosmic radiations on Earth‟s environment and biosphere” and to the Cuban Ministry for Science, Technology and Environment for funding our research project „‟Mathematical Modeling of Ecosystems‟‟. References Amemiya, T., Enomoto, T., Rossberg, A., Yamamoto, T., Inamori,Y., Itoh, K.: Stability and dynamical behaviour in a real lake model and implications for regime shifts in real lakes. Ecological modeling 206, 54-62, (2007) ASTM G173 - 03e1. ASTM G173 - 03e1 Standard Tables for Reference Solar Spectral Irradiances. http://www.astm.org/Standards/G173.htm Benitez, N., Maiz-Apellaniz, J., Canelles, M.: Evidence for Nearby Supernova Explosions. Phys. Rev. Lett. 88 081101 (2002) Galante, D., Horvath, J.: Biological effects of gamma-ray bursts: distances for severe damage on the biota. International Journal of Astrobiology 6 (1): 19–26 (2007) Gonzalez, G., Brownlee, D., Ward, P.: The Galactic Habitable Zone I. Galactic Chemical Evolution. Icarus 152, 185 (2001) 10 Hader, D., Kumar, H., Smith, R., Worrest, R.: Effects of solar UV radiation on aquatic ecosystems and interactions with climate change. Photochem. Photobiol. Sci., 6, 267– 285 (2007) Martin, O., Galante, D., Cardenas, R., Horvath, J.: Short-term effects of gamma ray bursts on Earth. Astrophys Space Sci (2009) 321: 161–167. DOI 10.1007/s10509-009- 0037-3 NCAR/ACD. NCAR/ACD TUV: Tropospheric Ultraviolet & Visible Radiation Model, NCAR Atmospheric Chemistry Division (ACD). NCAR/ACD. http://cprm.acd.ucar.edu/Models/TUV/ Thomas, B., Melott, A., Jackman, C., Laird, C., Medvedev, M., Stolarski, R., Gehrels, N., Cannizzo, J., Hogan, D., Ejzak, L.: Gamma-Ray Bursts and the Earth: Exploration of Atmospheric, Biological, Climatic and Biogeochemical Effects, Astrophys. J., 634, 509-533 (2005) 11
1902.00047
1
1902
2019-01-31T19:36:57
Topography of (exo)planets
[ "astro-ph.EP" ]
Current technology is not able to map the topography of rocky exoplanets, simply because the objects are too faint and far away to resolve them. Nevertheless, indirect effect of topography should be soon observable thanks to photometry techniques, and the possibility of detecting specular reflections. In addition, topography may have a strong effect on Earth-like exoplanet climates because oceans and mountains affect the distribution of clouds \citep{Houze2012}. Also topography is critical for evaluating surface habitability \citep{Dohm2015}. We propose here a general statistical theory to describe and generate realistic synthetic topographies of rocky exoplanetary bodies. In the solar system, we have examined the best-known bodies: the Earth, Moon, Mars and Mercury. It turns out that despite their differences, they all can be described by multifractral statistics, although with different parameters. Assuming that this property is universal, we propose here a model to simulate 2D spherical random field that mimics a rocky planetary body in a stellar system. We also propose to apply this model to estimate the statistics of oceans and continents to help to better assess the habitability of distant worlds.
astro-ph.EP
astro-ph
Topography of (exo)planets F. Landais (1*), F. Schmidt (1), and S. Lovejoy (2) February 4, 2019 (1) GEOPS, Univ. Paris-Sud, CNRS, Université Paris Saclay, Rue du Belvedere, Affiliations: Bat. 504 -- 509, 91405 Orsay, France (2) Physics department, McGill University, 3600 University st., Montreal, Que. H3A 2T8, Canada *Correspondence to: F. Landais ([email protected]) Abstract Current technology is not able to map the topography of rocky exoplanets, simply because the objects are too faint and far away to resolve them. Nevertheless, indirect effect of to- pography should be soon observable thanks to photometry techniques, and the possibility of detecting specular reflections. In addition, topography may have a strong effect on Earth-like exoplanet climates because oceans and mountains affect the distribution of clouds (Houze, 2012). Also topography is critical for evaluating surface habitability (Dohm and Maruyama, 2015). We propose here a general statistical theory to describe and generate realistic synthetic topographies of rocky exoplanetary bodies. In the solar system, we have examined the best- known bodies: the Earth, Moon, Mars and Mercury. It turns out that despite their differences, they all can be described by multifractral statistics, although with different parameters. As- suming that this property is universal, we propose here a model to simulate 2D spherical random field that mimics a rocky planetary body in a stellar system. We also propose to 1 apply this model to estimate the statistics of oceans and continents to help to better assess the habitability of distant worlds. Keywords: planetary systems, planets and satellites: surfaces, planets and satellites: terrestrial planets, methods: numerical 1 Introduction Efforts to detect and study exoplanets in other solar systems were initially restricted to gas giants (Mayor and Queloz, 1995) but multiple rocky exoplanets have now been discovered (Wordsworth et al., 2011). Their climates depend mainly on their atmospheric composition, stellar flux and orbital parameters (Wang et al., 2014; Forget and Leconte, 2014). But topography also plays a role in atmospheric circulation (Blumsack, 1971) and is an important trigger for cloud formation (Houze, 2012). Furthermore, the presence of an ocean filled with volatile compounds at low albedo is of a prime importance to the climate (Charnay et al., 2013). Last but not least, surface habitability relies on the presence of the three elements: the atmosphere, ocean and land (Dohm and Maruyama, 2015). Topography is also the determinant of ocean and land cover. Thanks to different observations techniques, measurements of the atmospheres of hot Jupiter planets have been achieved (Seager, 2010). Significantly, the detection of clouds has been reported (Demory et al., 2013) indicating strong heterogeneity in their spatial distribution. The detection of the first atmospheric transmission spectra of a super-Earth (Bean et al., 2010) and the discovery of a rocky exoplanet in the habitable zone around a dwarf star opens a new area in exoplanet science (de Wit et al., 2016). Such observations are expected to be increasingly frequent (Tian, 2015). Nevertheless, with current technology, direct imaging of exoplanets is very difficult because the objects are too faint and too far away. For the moment, the only way to determine the topography is by statistical models. In the near future, photometry techniques should improve our knowledge of exoplanet topography , even if the bodies are not resolved in ways similar to the small bodies in our Solar System (see for 2 instance Lowry et al. (2012) for estimates of the shape of comet 67P before the Rosetta landing). In addition, if oceans or lakes are present, their specular reflection should be detectable , for example, as also observed through the haze of Titan (Stephan et al., 2010) . Even if exoplanets are too far to be resolved, their topographies should be studied now. We offer here a framework to prepare and interpret future observations. Recently, we reported the first unifying statistical similarity between the topographic fields of the best known bodies in the Solar System: Earth, Moon, Mars and Mercury (Landais et al., 2018). All these topographies seems to be well described by a mathematical scaling framework called «multifractals». The multifractal model, initially proposed for topography by Lavallee et al. (1993) describes the distribution and correlation of slopes at different scales. More precisely, we consider here the "universal multifractal" model developed by Schertzer and Lovejoy (1987). The accuracy of such a model has been tested in the case of different available topographic fields on Earth (Gagnon et al., 2006), Mars (Landais et al., 2015), Mercury and the Moon (Landais et al., 2018). This model has the advantage to reproduce closely the statistical properties of natural topography: the scaling properties, but also the intermittency (both rough and smooth regions can be found on the planets). Universal multifractals depend on only 3 parameters: H controls how the roughness changes from one scale to another and C1 controls the spatial heterogeneity of the roughness near the mean and α quantifies how rapidly the properties change as we move away from the mean topographic level. The bodies studied show transitions at ~10 km and are characterized by specific multifractal parameters (Landais et al., 2018). The scaling law at large scales (> 10 km) is characterized for the Moon by H = 0.2, Mercury by H = 0.3, Mars and Earth by H = 0.5. The α ∼ 1.9 for the Earth, Mars and Mercury but α ∼ 1.4 for the Moon. The C1 ∼ 0.1 for Earth and Mars, with lower values C1 ∼ 0.06 for Mercury and C1 ∼ 0.03 for the Moon. These differences are interpreted to be linked to dynamical topography and variation of elastic thickness of the crust (see table 1). Assuming that exoplanets are statistically similar to those observed in our own Solar System, we propose here a stochastic topographic model. Such models will be very useful for investigating 3 the distribution of exoplanet oceans, for studying the effect of topography on exoplanet climates, and for studying the effect of topography on their orbital motions or for determining the effect of topography and roughness on photometry. It can also be used to study the early climate on Earth. The purpose of this article to first present our statistical model and its implementation on the sphere we then discuss the distribution of oceans and land cover. An introduction to the multi- fractal formalism can be found in the next section. 2 Method 2.1 Universal multifractals The first application of fractional dimensions on topography was by B. Mandelbrot in his article "how long is the coast of Britain" (Mandelbrot, 1967). Fractals are geometrical sets of points that have scaling, power law, deterministic or statistical relations from one scale to another. This type of behavior has been observed in geophysical phenomenon including turbulence - clouds, wind, ocean gyres - but also faults in rock, geogravity, geomagnetism and topography (Lovejoy and Schertzer, 2007) . The most common way to test scaling is to study the dependence of various statistics as functions of scale. Topographic level contours (isoheights) are fractals if for example the length of the In this case, the contour is a power law function of the resolution at which it is measured. level set is "scaling" and the exponent is its fractal dimension. In real topography, each level set has its own different fractal dimension so that the topography itself is a multifractal (Lavallee et al., 1993). Numerous studies haves shown that in several contexts, topography is scaling over a significant range of scales (see the review in Lovejoy and Shertzer, 2013). If the topography is multifractal, fractal dimensions measured locally appear to vary from one location to another. Indeed multifractal fields can be thought as a hierarchy of singularities whose exponents are random variables. Modern developments have introduced the notion of multifractal processes for such fields. For such processes, a local estimate of a fractal exponent is expected be different from a 4 location to another without requiring different processes to generate it. With multifractals, it is possible to interpret the topography of regions that exhibit completely different slope distributions in a unified statistical framework. These models suggest global topography analyses are relevant despite of their diversity and complexity. Previous studies (Gagnon et al., 2006; Lavallee et al., 1993) have established the accuracy of multifractal global statistical approach in the case of Earth's topography. More precisely, a particular class of multifractal has been considered: the universal multifractal, a stable and attractive class (Schertzer and Lovejoy, 1987). In our previous analysis (Landais et al., 2015), we performed the same kind of global analysis on the topographic data from Mars, from MOLA laser altimeter measurement (Smith et al., 2001). This analysis also find a good agreement with universal multifractal but on a restricted range of scale (Landais et al., 2015). Indeed the statistical structure has been found to be different at small scale (monofractal) and large scale (multifractal) with a transition occurring around 10 km. Fluctuations ations. The simplest fluctuation that can be used to describe topography is the distribution of In order to interpret topography as a multifractal, we must quantify its fluctu- changes in altitude ∆h over horizontal distances ∆x. There are many other ways to define fluc- tuations, the general framework being wavelets. The simple altitude difference corresponds to the so called "poor man's" wavelet and can be efficiently replaced by the Haar wavelet that tends to converge faster and is useful over a wider range of geophysical process. Over an interval ∆x, the Haar fluctuation is the average elevation over the first half of the interval minus the average eleva- tion over the second half (see Lovejoy, 2014; Lovejoy and Schertzer, 2012) and paragraph below for a precise definition of Haar fluctuations). The computation of fluctuations can be performed for each pair of elevation data in order to accumulate a huge amount of slope fluctuations. From this, a global planetary average M (∆x) can be performed and will reflect the mean fluctuation of slopes at the scale ∆x. Scaling By estimating fluctuations at different scales, we can observe the structure of the sta- tistical dependance of the ensemble mean fluctuation at scale ∆x: M (∆x) . If the topographic 5 field is fractal, this dependance is a power-law corresponding to equation 1 where H is a power law exponent (named in Honor of Ewin Hurst and equal to the Hurst exponent in the monofractal, Gaussian case): M (∆x) ∼ ∆xH (1) Statistical moments Additionally, instead of simply considering the average (i. e.the first statistical moment of the fluctuations), we can compute any statistical moment Mq of order q defined by Mq =< ∆hq > ; Mq is called the qth order structure function. If q = 2, it simply corresponds to the usual (variance based) structure function. In principle, all orders (including non-integer orders) must be computed to fully characterize the full variability of the data. Multifractality Mq allows us to introduce two distinct statistical structures of interest: monofrac- tal and multifractal. For a detailed description of the formalism we apply in this study, the readers can refer to Lovejoy and Shertzer (2013) briefly summed in Landais et al. (2015). We quickly recall the main notions here : • In the monofractal case the parameters H is sufficient to describe the statistics of all the moments of order q (equation 2). In this case, no intermittency is expected, meaning that the roughness of the field is spatially homogenous despite of its fractal variability regarding to scales. Typically, the value H = 0.5 corresponds to the classic Brownian motion. This kind of model has been used in many local and regional analysis of natural surfaces (Orosei et al., 2003; Rosenburg et al., 2011), but it fails to account for the intermittency (and strongly non-Gaussian statistics) commonly observed on large topographic datasets. Mq ∼ ∆xqH (2) • In the multifractal case, H is no longer sufficient to fully describe the statistics of the moments 6 of order q. An additional convex function K(q) depending on q is required (3). Mq ∼ ∆xqH−K(q); (3) • The moment scaling function K modifies the scaling law of each moment. The consequence on the corresponding field appears clearly on simulations: the field exhibit a juxtaposition of rough and small places that are clearly more realistic in the case of natural surfaces (Gagnon et al., 2006). Moreover, it is possible to restrain the generality of the function K(q) by considering universal multifractals, a stable and attractive class proposed by Schertzer and Lovejoy (1987) for which the multifractality is completely determined by the mean (codimension of the mean) and the curvature α of the function intermittency C1 = K, α = 1 evaluated at q = 1 (the degree of multifractality). In this case the expression of K is simply given by equation 4 (cid:16) dK(q) d2K(q) C1 dq2 (cid:17) dq q=1 K(q) = C1 α − 1 (qα − q) (4) 2.2 Spherical multifractal simulation Simulations in 1D or 2D with multifractal properties and specific values for α, H and C1 can be obtained by the procedure defined by Schertzer and Lovejoy (1987); Wilson et al. (1991). The necessary steps are briefly reminded here after : • Step 1 : Generation of a un-correlated Levy noise γα(r). When α = 2, it simplifies to a gaussian white noise whereas α < 2 corresponds to an extremal levy variable with negative extreme values. • Step 2 : Convolution of γα(r) with a singularity gα(r) defined by equation 5 to obtain a Levy-generator Γα(r), by using a convolution denoted by "(cid:63)" gα(r) = r−2/α (5) 7 Γ(r) = C 1/α 1 g(r) (cid:63) γα(r) • Step 3 : Exponentiation of the generator to obtain the multifractal noise ε ε = eΓ (6) (7) • Step 4 : The final field is then obtained by fractional integration of order H (another convo- lution similar to step 2) Whereas the convolutions required for step 2 and 4 can easily be performed in Fourier space for the cartesian case, the generalization to spherical case is not straightforward, but as shown in appendix 5D of Lovejoy and Shertzer (2013), it can be done using spherical harmonics. Let θ and ϕ being respectively the colatitude and longitude angle, the singularity can be expressed by equation 8. As it is symmetric by rotation along ϕ, gα(θ, ϕ) only depend on θ. gα(θ, ϕ) = θ−2/α (8) Let the spherical harmonic expansion of gα(θ, ϕ) be given by equation 9, where Ylm is the spherical harmonic of order m and l. As gα(θ, ϕ) does not depend on ϕ, all the Ylm for m (cid:54)= 0 are equal to zero. (cid:88) σlYl,0 gα(θ, ϕ) = Let the spherical harmonic expansion of γα(θ, ϕ) be given by : (cid:88) γα(θ, ϕ) = ulmYl,m(θ, ϕ) Then the convolution C of gα(θ, ϕ) and γα(θ, ϕ) is given by : (cid:115) (cid:88) l,m σl C = 4π 2l + 1 ulmYl,m(θ, ϕ) 8 (9) (10) (11) Figure 1: Mean fluctuations of topography of Earth, Mars, Moon and Mercury, as a function of scale. All dataset are normalized in order to be equal to 1 at the scale 10km. The normalization does not modify the scaling behavior but emphasize the transition occurring at around 10km. The errors bars are smaller than the size of the points. 3 Results 3.1 Solar System In this section, we recall the main results of the planetary bodies of the Solar System. On Figure 1, we have plotted the mean normalized fluctuations of altitude as a function of scale on a log-log plot. The easiest way to define fluctuations at a given scale ∆x is to take the simple difference of altitude between two points separated by the distance ∆x. We average all of these fluctuations over the whole planetary body. As we are focusing on statistical properties, the results on figure 9 ScalesNormalized mean Haar fluctuations 1 have been normalized in order to emphasize the transition between 2 distinct range of scales. The global average have been normalized in order to be similar around 10 km. As a consequence of this normalization, it is not possible to compare the absolute altitude and roughness values on this plot, only the scaling laws. One can see the similarity between curves at lower scales (<10 km) and distinct scaling behaviors at higher scales (>10km). Still in each case, the dependance towards scales remains roughly linear on a log-log plot revealing a simple power-law behavior. The parameters H is taken as a function of the linear coefficient of the fit and thus control how the mean fluctuations of elevations behave towards scales. This kind of linear behavior is called fractal or monofractal. Moreover the multifractal model includes two other parameters (C1 and α) that control the spatial distribution of roughness. Thanks to C1 and α, it is possible to have a global description, in a common statistical framework, including regions with heterogeneous roughness at a given scales. More details about the two non-trivial parameters may be found in the appendices. Global mea- sures of H, C1 and α in the case of Earth, Mars, Moon and Mercury have produced satisfying results (see table 1and Landais et al., 2018). We analyzed the generated random field and show that the estimated H, C1 and α are in agreement with the expected values for a large range of parameter space. 3.2 Exoplanets Table 1: Estimates of the parameters H, α and C1 Earth low high 0.8 0.5 0.1 0.001 NA 1.9 Mars Moon low high low high 0.9 0.2 0.7 0.5 0.11 0.004 0.03 0.04 1.8 NA 1.4 NA Mercury low high 0.7 0.3 0.06 0.004 NA 1.9 H C1 α Given its simplicity and its accuracy in the case of several real topographies, the multifractal model should be a good candidate for producing artificial topographies of (exo)planets. Figure 2 provides several examples of spherical topography obtained by our simulation model for varying 10 values of C1 and H. One can see the interesting multifractal features . In the case of non-zero C1, the roughness level is highly heterogeneous with an alternation of smooth and rough terrains depending on the altitude. This features makes the multifractal simulations much more realistic by (implicitly) taking into account the possible occurrence of oceans or large smooth volcanic plains that are statistically different from deeply cratered terrains or mountainous areas where the level of roughness is high. Whereas the value of H controls the rate at which the roughness changes with scale (see figure 2a), the value of C1 = 0.1 controls the proportion of rough and smooth places (see figure 2b). A high value increases the roughness discrepancies between locations. One has to remember that only the scaling laws are simulated here, neither the absolute height, nor the radius of the planet. Vertical exaggeration has been set arbitrarily in order to maximize the visual impression. Nevertheless, the variety of shapes and roughnesses produced are astonishing and in addition to terrestrial planets, could potentially even be realistically applied to small bodies including asteroids and comets. To estimate the properties of potential exoplanet surfaces, we conducted a statistical analysis of oceans and continents obtained from 500 simulated multifractal topography fields at 1° spatial resolution with the set of parameters obtained for the global estimates on Earth (H = 0.5, α = 1.9, C1 = 0.1). In order to deal with the notion of oceans and continents, one must first define the sea/land cover. We define the sea level s, as a quantile of the global topographic distribution. This definition simply means that at quantile s, the sea level is such as s is also the surface proportion of the sea. For instance, (i) s = 0.5 is the median altitude and half of the planet is ocean covered , half by land; (ii) s = 0.9 means that 90% of the planet area is ocean covered and 10% is land. Oceans and continents are respectively defined as disconnected areas located beneath or above the sea level s. We plotted on figure 3 a example of synthetic multifractal topographies with varying ratio s. On Figure 4 we plotted the size of the largest continent and largest ocean as functions of s. We summarized the 500 experiments by computing the average, standard deviation and mini- mum/maximum. As one can see, the simulations produce typically one large ocean or one large 11 (a) Spherical simulations at 0.1° resolution for different values of H (α=1.9 and C1 = 0.1). H varies from 0.2 to 0.99. Synthetic bodies with low H values have little large-scale altitude fluctuations and are rough at small scales. As a result, their shape is similar to a regular sphere but with a rough texture. When H increases, this behavior tends to be reversed : large altitude variations appear at large scales deforming the body, which has a smoother texture. (b) Spherical simulations at 0.1° resolution for two values of C1 (α=1.9 and H = 0.5 constant). From left to right C1 is 0 and 0.1. The left simulation (C1 = 0) is characterized by a spatially homogeneous roughness. On the contrary, themultifractal simulation on the right shows alternating smooth and rough areas Figure 2: Several example of synthetic spherical topographic fields by varying H and C1 12 C1=0C1=0.1 Figure 3: Synthetic multifractal topography at 0.1° resolution as a function of sea level. The fraction of the planet's surface covered by ocean is noted s. The simulation is set for the Earth/Mars like planet (H = 0.5, α = 1.9, C1 = 0.1). Low altitude regions are smoother than high altitude ones. See also video 2 in sup.mat. 13 s=0.1s=0.3s=0.5s=0.7longitude 0-180longitude 180-360 Figure 4: The Ocean/continent relationship. The size (as proportion of the total planet surface) of the largest continent (blue) and ocean (green)) for different values of sea level s. The diamond indicate the mean size with one standard deviation bars, whereas the circles indicate the mini- mum and maximum value in each case. The blue and green lines ) correspond to proportions of the remaining area covered by continents and ocean . These results are based on 500 synthetic topography simulations of an Earth-like planet (H = 0.5, α = 1.9, C1 = 0.1). The red diamonds are for the Earth. 14 continent with a size close to the maximum available area indicating that it is highly improbable to obtain two disconnected large areas. However, at respectively very small or very large values of s, the available area is split between several small oceans (conversely, large s and small continents). Finally, we apply the same analysis on the particular case of Earth based on ETOPO1 (Amante and Eakins, 2009) and use red diamonds to indicate the size of the largest ocean and continents as functions of the terrestrial value of s (s ≈ 0.66). The points are satisfyingly close to those obtained by multifractal simulations supporting the accuracy of the model. Following Dohm and Maruyama (2015), we investigate the interface between ocean, atmosphere and land. From our results, on average the size of the largest ocean or continent is always close to the maximum available size (near the 90% line). The congruent part of the surface covered by ocean (or land) is split up into smaller but more numerous islands (or lakes), as also observed on the Earth (Downing et al., 2006). There are some extreme cases, where the largest continent is very small. Interestingly, this case happens more for small sea levels. If s = 0.1, the extreme case can even reach 25%, meaning that the largest ocean only covers 25% of the ocean surface , 75% are thus covered by smaller lakes. The symmetric situation occurs for s = 0.9 : the largest continent only covers 25% of the land, 75% are thus covered by small islands. The Earth corresponds to the average situation since all the major oceans are connected through the thermo-haline circulation. From this study, we can exclude the situation of two large unconnected oceans, representing a global sea surface > 50%. The same for two large unconnected continents, representing a global sea surface > 50%. As a summary, the interface between land and sea, so important for habitability, can be statistically constrained by this model. 4 Conclusion Multifractal simulations on spheres are able to statistically reproduce the morphology of planetary bodies, and even potentially small bodies such asteroids and comets. In addition, it offers a wide field of investigation for evaluating the role of the topography in exoplanet signals , thanks to 15 photometry and specular reflection, this is especially true for transiting objects . The simulations will serve as a starting point for future studies aimed at characterizing the overall photometric response of unresolved rotating bodies. Our synthetic numerical topographies can be integrated into the development of realistic exoplanet climate simulations in different contexts by integrating the roles of clouds and surface / atmosphere interactions. In particular, exoplanets in gravitational lock are subjected to climatic instabilities (Kite et al., 2011). In particular, our results suggest that it is statistically highly unlikely to have two major united oceans on either side of the globe. If the dark side is too cold and the sunny side too hot to allow the presence of liquid water, the topography could contribute to creating to a global glacier, continually moving the volatile elements from the illuminated side to the dark side. This dynamic state should significantly increase the presence of liquid water at the terminator with consequences for habitability. By construction the statistical properties of all our simulations are isotropic. The procedure used can be modified to generate anisotropic topographies but poses a number of technical problems that have not yet been addressed. Anisotropy adds degrees of freedom that make the problem more complex both in generation but also in determining parameters on real data. To deal with this question, we should consider implementing the formalism of generalized scale invariance (GSI, Schertzer, 2011) as a future work. We provide a 3D visualization of some examples with varying parameters (https://data.ipsl.fr/exotopo/). In addition, a dataset of synthetic spherical topographies can be downloaded by the reader (http://dx.doi.org/10.14768/20181024001.1) Acknowledgement We acknowledge support from the "Institut National des Sciences de l'Univers" (INSU), the "Centre National de la Recherche Scientifique" (CNRS) and "Centre National d'Etudes Spatiales" (CNES) through the "Programme National de Planétologie" and the "Programme National de Télédétec- tion spatiale", the MEX/OMEGA and the MEX/PFS programs. We thank R. Orosei and the 16 Assistant Editor M. Hollis for their constructive reviews. We thank C. Marmo for the develop- ment of the 3D visualization tool and W. Pluriel for his contribution to the project. We also thank ESPRI-IPSL for hosting the data. References Amante, C., and B. Eakins (2009), Etopo1 1 arc-minute global relief model: Procedures, data sources and analysis., NOAA Technical Memorandum NESDIS NGDC-24. National Geophysical Data Center, NOAA. Bean, J. L., E. M.-R. Kempton, and D. Homeier (2010), A ground-based transmission spectrum of the super-earth exoplanet gj 1214b, Nature, 468 (7324), 669 -- 672. Blumsack, S. L. (1971), On the effects of topography on planetary atmospheric circulation, Journal of the Atmospheric Sciences, 28 (7), 1134 -- 1143, doi:10.1175/1520-0469(1971)028<1134: OTEOTO>2.0.CO;2. Charnay, B., F. Forget, R. Wordsworth, J. Leconte, E. Millour, F. Codron, and A. Spiga (2013), Exploring the faint young sun problem and the possible climates of the archean earth with a 3-D GCM, J. Geophys. Res. Atmos., 118 (18), 10,414 -- 10,431, doi:10.1002/jgrd.50808. de Wit, J., et al. (2016), A combined transmission spectrum of the earth-sized exoplanets trappist- 1 b and c, Nature, 537 (7618), 69 -- 72, doi:10.1038/nature18641. Demory, B.-O., et al. (2013), Inference of inhomogeneous clouds in an exoplanet atmosphere, The Astrophysical Journal Letters, 776 (2), L25. Dohm, J. M., and S. Maruyama (2015), Habitable trinity, Geoscience Frontiers, 6 (1), 95 -- 101, doi: http://dx.doi.org/10.1016/j.gsf.2014.01.005, special Issue: Plate and plume tectonics: Numerical simulation and seismic tomography. 17 Downing, J. A., et al. (2006), The global abundance and size distribution of lakes, ponds, and impoundments, Limnology and Oceanography, 51 (5), 2388 -- 2397, doi:10.4319/lo.2006.51.5.2388. Forget, F., and J. Leconte (2014), Possible climates on terrestrial exoplanets, Philosophical Trans- actions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 372 (2014), doi:10.1098/rsta.2013.0084. Gagnon, J.-S., S. Lovejoy, and D. Schertzer (2006), Multifractal earth topography, Nonlinear Processes in Geophysics, 13 (5), 541 -- 570, doi:10.5194/npg-13-541-2006. Houze, R. A. (2012), Orographic effects on precipitating clouds, Reviews of Geophysics, 50 (1), n/a -- n/a, doi:10.1029/2011RG000365, rG1001. Kite, E. S., E. Gaidos, and M. Manga (2011), Climate instability on tidally locked exoplanets, The Astrophysical Journal, 743 (1), 41. Landais, F., F. Schmidt, and S. Lovejoy (2015), Universal multifractal martian topography, Non- linear Processes in Geophysics, 22 (6), 713 -- 722, doi:10.5194/npg-22-713-2015. Landais, F., F. Schmidt, and S. Lovejoy (2018), Multifractal topography of several planetary bodies in the solar system, Icarus, http://arxiv.org/abs/1805.11249. Lavallee, D., S. Lovejoy, D. Schertzer, and P. Ladoy (1993), Nonlinear variability and landscape topography: analysis and simulation. fractals in geography, Eds. L. De Cola, N. Lam, 158-192, PTR, Prentice Hall. Lovejoy, S. (2014), A voyage through scales, a missing quadrillion and why the climate is not what you expect, Climate Dynamics. Lovejoy, S., and D. Schertzer (2007), Scaling and multifractal fields in the solid earth and topog- raphy, Nonlinear Processes in Geophysics, 14 (4), 465 -- 502. 18 Lovejoy, S., and D. Schertzer (2012), Haar wavelets, fluctuations and structure functions: convenient choices for geophysics, Nonlinear Processes in Geophysics, 19 (5), 513 -- 527, doi: 10.5194/npg-19-513-2012. Lovejoy, S., and D. Shertzer (2013), The Weather and Climate :Emergent Laws and Multifractal Cascades, Cambridge. Lowry, S., S. Duddy, B. Rozitis, S. F. Green, A. Fitzsimmons, C. Snodgrass, H. H. Hsieh, and O. Hainaut (2012), The nucleus of comet 67p/churyumov-gerasimenko-a new shape model and thermophysical analysis, Astronomy & Astrophysics, 548, A12. Mandelbrot, B. (1967), How long is the coast of britain? statistical self-similarity and fractional dimension, Science, 156 (3775), 636 -- 638, doi:10.1126/science.156.3775.636. Mayor, M., and D. Queloz (1995), A jupiter-mass companion to a solar-type star, Nature, 378 (6555), 355 -- 359. Orosei, R., R. Bianchi, A. Coradini, S. Espinasse, C. Federico, A. Ferriccioni, and A. I. Gavr- ishin (2003), Self-affine behavior of martian topography at kilometer scale from mars or- biter laser altimeter data, Journal of Geophysical Research: Planets, 108 (E4), n/a -- n/a, doi: 10.1029/2002JE001883. Rosenburg, M. A., O. Aharonson, J. W. Head, M. A. Kreslavsky, E. Mazarico, G. A. Neumann, D. E. Smith, M. H. Torrence, and M. T. Zuber (2011), Global surface slopes and roughness of the moon from the lunar orbiter laser altimeter, Journal of Geophysical Research: Planets, 116 (E2), n/a -- n/a, doi:10.1029/2010JE003716. Schertzer, D., and S. Lovejoy (1987), Physical modeling and analysis of rain and clouds by anisotropic scaling multiplicative processes, Journal of Geophysical Research: Atmospheres, 92 (D8), 9693 -- 9714, doi:10.1029/JD092iD08p09693. 19 Schertzer, S., D. Lovejoy (2011), Multifractals, generalized scale invariance and complexity in geophysics, International Journal of Bifurcation and Chaos, 21 (12), 3417 -- 3456, doi:10.1142/ S0218127411030647. Seager, D., Sara; Deming (2010), Exoplanet atmospheres, Annual Review of Astronomy and As- trophysics, vol. 48, p.631-672. Smith, D. E., et al. (2001), Mars orbiter laser altimeter: Experiment summary after the first year of global mapping of mars, Journal of Geophysical Research: Planets, 106 (E10), 23,689 -- 23,722, doi:10.1029/2000JE001364. Stephan, K., et al. (2010), Specular reflection on titan: liquids in kraken mare, Geophysical Re- search Letters, 37 (7). Tian, F. (2015), Observations of exoplanets in time-evolving habitable zones of pre-main-sequence m dwarfs, Icarus, 258, 50 -- 53, doi:http://dx.doi.org/10.1016/j.icarus.2015.06.004. Wang, Y., F. Tian, and Y. Hu (2014), Climate patterns of habitable exoplanets in eccentric orbits around m dwarfs, The Astrophysical Journal Letters, 791 (1), L12. Wilson, J., D. Schertzer, and S. Lovejoy (1991), Continuous Multiplicative Cascade Models of Rain and Clouds, pp. 185 -- 207, Springer Netherlands, Dordrecht. Wordsworth, R. D., F. Forget, F. Selsis, E. Millour, B. Charnay, and J.-B. Madeleine (2011), Gliese 581d is the first discovered terrestrial-mass exoplanet in the habitable zone, The Astrophysical Journal Letters, 733 (2), L48. 20
1701.02534
1
1701
2017-01-10T12:03:41
Visible spectra of (474640) 2004 VN112-2013 RF98 with OSIRIS at the 10.4 m GTC: evidence for binary dissociation near aphelion among the extreme trans-Neptunian objects
[ "astro-ph.EP" ]
The existence of significant anisotropies in the distributions of the directions of perihelia and orbital poles of the known extreme trans-Neptunian objects (ETNOs) has been used to claim that trans-Plutonian planets may exist. Among the known ETNOs, the pair (474640) 2004 VN112-2013 RF98 stands out. Their orbital poles and the directions of their perihelia and their velocities at perihelion/aphelion are separated by a few degrees, but orbital similarity does not necessarily imply common physical origin. In an attempt to unravel their physical nature, visible spectroscopy of both targets was obtained using the OSIRIS camera-spectrograph at the 10.4 m Gran Telescopio Canarias (GTC). From the spectral analysis, we find that 474640-2013 RF98 have similar spectral slopes (12 vs. 15 %/0.1um), very different from Sedna's but compatible with those of (148209) 2000 CR105 and 2012 VP113. These five ETNOs belong to the group of seven linked to the Planet Nine hypothesis. A dynamical pathway consistent with these findings is dissociation of a binary asteroid during a close encounter with a planet and we confirm its plausibility using N-body simulations. We thus conclude that both the dynamical and spectroscopic properties of 474640-2013 RF98 favour a genetic link and their current orbits suggest that the pair was kicked by a perturber near aphelion.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 6 (2017) Preprint 7 September 2018 Compiled using MNRAS LATEX style file v3.0 Visible spectra of (474640) 2004 VN112 -- 2013 RF98 with OSIRIS at the 10.4 m GTC: evidence for binary dissociation near aphelion among the extreme trans-Neptunian objects⋆ J. de León,1,2† C. de la Fuente Marcos3 and R. de la Fuente Marcos3 1Instituto de Astrofísica de Canarias, C/ Vía Láctea s/n, E-38205 La Laguna, Tenerife, Spain 2Departamento de Astrofísica, Universidad de La Laguna, E-38206 La Laguna, Tenerife, Spain 3Universidad Complutense de Madrid, Ciudad Universitaria, E-28040 Madrid, Spain Accepted 2017 January 4. Received 2016 December 23; in original form 2016 October 20 ABSTRACT The existence of significant anisotropies in the distributions of the directions of perihe- lia and orbital poles of the known extreme trans-Neptunian objects (ETNOs) has been used to claim that trans-Plutonian planets may exist. Among the known ETNOs, the pair (474640) 2004 VN112 -- 2013 RF98 stands out. Their orbital poles and the directions of their perihelia and their velocities at perihelion/aphelion are separated by a few degrees, but orbital similarity does not necessarily imply common physical origin. In an attempt to unravel their physical nature, visible spectroscopy of both targets was obtained using the OSIRIS camera- spectrograph at the 10.4 m Gran Telescopio Canarias (GTC). From the spectral analysis, we find that 474640 -- 2013 RF98 have similar spectral slopes (12 vs. 15 %/0.1 µm), very different from Sedna's but compatible with those of (148209) 2000 CR105 and 2012 VP113. These five ETNOs belong to the group of seven linked to the Planet Nine hypothesis. A dynamical path- way consistent with these findings is dissociation of a binary asteroid during a close encounter with a planet and we confirm its plausibility using N-body simulations. We thus conclude that both the dynamical and spectroscopic properties of 474640 -- 2013 RF98 favour a genetic link and their current orbits suggest that the pair was kicked by a perturber near aphelion. Key words: techniques: spectroscopic -- techniques: photometric -- astrometry -- celestial mechanics -- minor planets, asteroids: individual: (474640) 2004 VN112 -- minor planets, as- teroids: individual: 2013 RF98. 1 INTRODUCTION The dynamical and physical properties of the extreme trans- Neptunian objects or ETNOs (semimajor axis, a, greater than 150 au, and perihelion distance, q, greater than 30 au, Trujillo & Shep- pard 2014) are intriguing in many ways. Their study can help probe the orbital distribution of putative planets going around the Sun be- tween the orbit of Pluto and the Oort Cloud as well as understand the formation and evolution of the Solar system as a whole. The first ETNO was found in 2000, (148209) 2000 CR105, and its dis- covery was soon recognised as a turning point in the study of the outer Solar system (e.g. Gladman et al. 2002; Morbidelli & Levison 2004). The current tally stands at 21 ETNOs. Trujillo & Sheppard (2014) were first in suggesting that the dynamical properties of the ETNOs could be better explained if a yet to be discovered planet of several Earth masses is orbiting ⋆ Based on observations made with the GTC telescope, in the Spanish Ob- servatorio del Roque de los Muchachos of the Instituto de Astrofísica de Ca- narias, under Director's Discretionary Time (program ID GTC2016-055). † E-mail: [email protected] c(cid:13) 2017 The Authors the Sun at hundreds of au. This interpretation was further sup- ported by de la Fuente Marcos & de la Fuente Marcos (2014) with a Monte Carlo-based study confirming that the observed patterns in ETNO orbital parameter space cannot result from selection ef- fects and suggesting that one or more trans-Plutonian planets may exist. A plausible multi-planet dynamical scenario was explored by de la Fuente Marcos, de la Fuente Marcos & Aarseth (2015). Based on observational data, and analytical and numerical work, Batygin & Brown (2016) presented their Planet Nine hypothesis that was further developed by Brown & Batygin (2016), but ques- tioned by Shankman et al. (2016). The orbits of seven ETNOs -- Sedna, 148209, (474640) 2004 VN112, 2007 TG422, 2010 GB174, 2012 VP113 and 2013 RF98 -- were used by Brown & Batygin (2016) to predict the existence of the so-called Planet Nine, most probably a trans-Plutonian super-Earth in the sub-Neptunian mass range. Out of the 21 known ETNOs, only Sedna has been observed spectroscopically (see Fornasier et al. 2009). Among the known ETNOs, the pair 474640 -- 2013 RF98 clearly stands out (de la Fuente Marcos & de la Fuente Marcos 2016). The directions of their perihelia (those of the vector from 2 J. de León, C. de la Fuente Marcos and R. de la Fuente Marcos the Sun to the respective perihelion point) are very close (angu- lar separation of 9.◦8), their orbital poles are even closer (4.◦1), and consistently the directions of their velocities at perihelion/aphelion are also very near each other (9.◦5), although improved values are given in Section 3.3; in addition, they have similar aphelion dis- tances, Q (589 au vs. 577 au). Assuming that the angular orbital elements of the ETNOs follow uniform distributions (i.e. they are unperturbed asteroids moving in Keplerian orbits around the Sun), the probability of finding by chance two objects with such a small angular separation between their directions of perihelia and, what is more important, also between their orbital poles is less than 0.0001, which suggests a common dynamical origin. However, a probable common dynamical origin does not imply a common physical ori- gin. In an attempt to unravel their physical nature, visible spec- troscopy of the two targets was obtained on 2016 September using the OSIRIS camera-spectrograph at the 10.4 m Gran Telescopio Canarias (GTC) telescope, located in La Palma (Canary Islands, Spain). Here, we present and discuss the results of these observa- tions. This Letter is organized as follows. Section 2 reviews the state of the art for this pair of ETNOs. The new observations -- including spectroscopy, photometry and astrometry -- and their re- sults are presented in Section 3. The possible origin of this pair is explored in Section 4, making use of the new observational results and N-body simulations. Conclusions are summarized in Section 5. 2 THE PAIR 474640 -- 2013 RF98: STATE OF THE ART Asteroid (474640) 2004 VN112 was discovered on 2004 Novem- ber 6 by the ESSENCE supernova survey observing with the 4 m Blanco Telescope from Cerro Tololo International Observatory (CTIO) at an apparent magnitude R of 22.7 (Becker et al. 2007). Its absolute magnitude, H = 6.4 (assuming a slope parameter, G = 0.15), suggests a diameter in the range 130 -- 300 km for an as- sumed albedo in the range 0.25 -- 0.05. The orbital solution (2016 August) for this object is based on 31 observations spanning a data- arc of 5113 d or 14 yr, from 2000 September 26 to 2014 September 26, its residual rms amounts to 0.′′19.1 Such an object would have been visible to ESSENCE for only about 2 per cent of its orbit, suggesting that the size of the population of minor bodies mov- ing in orbits similar to that of 474640 could be very significant (Becker et al. 2008). Sheppard (2010) gives a normalised spectral gradient of 11±4 %/0.1 µm for this object based on Sloan g′, r′, i′ optical photometry acquired in 2008. Some additional photome- try was obtained with the Hubble Wide Field Camera 3 (Fraser & Brown 2012). Its optical colours are relatively neutral and this was interpreted by Brown (2012) as a sign that it is not dominated by methane irradiation products. Asteroid 2013 RF98 was discovered on 2013 September 12 by the Dark Energy Camera (DECam, Flaugher et al. 2015) observing from CTIO for the Dark Energy Survey (DES, Abbott et al. 2005) at an apparent magnitude z of 23.5 (Abbott et al. 2016). Its absolute magnitude, H = 8.7, suggests a diameter in the range 50 -- 120 km. The orbital solution (2016 August) for this ETNO is rather poor as it is based on 38 observations spanning a data-arc of just 56 d, from 2013 September 12 to 2013 November 7, its residual rms amounts to 0.′′12.2 No other data, besides these 38 observations, its H value, and its orbital solutions, are known about this ETNO. Asteroid 474640 has been classified as an extreme detached object by Sheppard & Trujillo (2016) and their integrations show that it is a stable ETNO within the standard eight-planets-only So- lar system paradigm; this result is consistent with that in Brown & Batygin (2016). In Sheppard & Trujillo (2016), 2013 RF98 is classified as an extreme scattered object and found to be unstable within the standard paradigm over 10 Myr time-scales due to Nep- tune perturbations. Both ETNOs are rather unstable within some in- carnations of the Planet Nine hypothesis (de la Fuente Marcos, de la Fuente Marcos & Aarseth 2016). The heliocentric orbits avail- able in 2016 August1,2 have been used to compute the values of the angular separations quoted in the previous section. In spite of the limitations of the orbital solution of 2013 RF98, the uncertain- ties mainly affect orbital elements other than inclination, i, longi- tude of the ascending node, Ω, and argument of the perihelion, ω. These three orbital parameters are the only ones involved in the calculation of the directions of perihelia, location of orbital poles, and directions of velocities at perihelion (de la Fuente Marcos & de la Fuente Marcos 2016). In sharp contrast, the value of Q of 2013 RF98 is affected by a significant uncertainty (12 per cent). 3 OBSERVATIONS 3.1 Spectroscopy Visible spectra of the two targets were obtained using OSIRIS (Cepa et al. 2000) at 10.4 m GTC. The apparent visual magnitude, V, at the time of observation was 23.3 for (474640) 2004 VN112 (heliocentric distance of 47.7 au and phase of 1.◦1) and 24.4 for 2013 RF98 (36.6 au, 1.◦3). For each target, acquisition images in the Sloan r′ filter were obtained in separate nights in order to iden- tify reliably the object in the field of view. This procedure ended up being the most efficient to detect these dim, slow-moving (ap- parent proper motion <2′′/h) targets. Visible spectra were acquired using the low-resolution R300R grism (resolution of 348 measured at a central wavelength of 6635 Å for a 0.′′6 slit width), that covers the wavelength range from 0.49 to 0.92 µm, and a 2′′slit width. Two widely accepted solar analogue stars from Landolt (1992) -- SA93- 101 and SA115-271 -- were observed using the same spectral con- figuration at an airmass identical to that of the targets to obtain the reflectance spectra of the ETNOs. For a given ETNO, the spectrum was then divided by the corresponding spectrum of the solar ana- log. Additional data reduction details are described by de León et al. (2016) and Morate et al. (2016). For 474640 we acquired two spectra of 1800 s each, while for 2013 RF98 we acquired four in- dividual spectra of 1800 s each. Observational details are shown in Table 1. The resulting individual reflectance spectra, normalised to unity at 0.55 µm and offset vertically for clarity, are shown in Fig. 1. In Table 1, the spectral slope (in units of %/0.1 µm) has been computed from a linear fitting to the spectrum in the wavelength range 0.5 -- 0.9 µm. We used an iterative process -- removing a total of 50 points randomly distributed in the spectrum and performing a linear fitting -- and obtained a value of the slope for each iteration. The resulting slope value is the mean of a total of 100 iterations 1 Orbit available from JPL's Small-Body Database and Horizons On-Line Ephemeris System: a = 318 ± 1 au, e = 0.8513 ± 0.0005, i = 25. ◦5748 ± 0. ◦010, referred to the ◦9990 ± 0. ◦0004, Ω = 65. epoch 2457600.5 JD TDB. ◦0007 and ω = 327. ◦121 ± 0. 2 Orbit available from Horizons: a = 307 ± 37 au, e = 0.882 ± 0.015, i = 29. ◦08, Ω = 67. ◦13 and ω = 316◦ ± 6◦, referred to the epoch 2457600.5 JD TDB. ◦62 ± 0. ◦51 ± 0. MNRAS 000, 1 -- 6 (2017) Visible spectra of (474640) 2004 VN112 -- 2013 RF98 3 (474640) 2004 VN112 Spec#1 Spec#2 (474640) 2004 VN112 2013 RF98 e c n a t c e l f e R e v i t a l e R 3.0 2.5 2.0 1.5 1.0 0.5 0.0 2013 RF98 0.4 0.5 0.7 0.6 0.8 Wavelength (µm) 0.9 1.0 e c n a t c e l f e R e v i t a l e R e c n a t c e l f e R e v i t a l e R 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 8 6 4 2 0 Spec#1 Spec#2 Spec#3 Spec#4 Figure 2. Final visible spectra of the pair of ETNOs (see the text for details). e c n a t c e l f e R e v i t a l e R 2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.5 0.55 0.6 (474640) 2004 VN112 2013 RF98 0.7 0.75 0.65 Wavelength (µm) 0.8 0.85 0.9 Figure 3. Comparison between the spectra of (474640) 2004 VN112 and 2013 RF98 smoothed by a Savitzky-Golay filter (see the text) and scaled to match at 0.60 µm. The most prominent absorption band of pure methane ice at 0.73 µm is not seen on either spectra. ative reflectance is both slowly varying and corrupted by random noise, we applied a Savitzky-Golay filter (Savitzky & Golay 1964) to both datasets. Such filters provide smoothing without loss of res- olution, approximating the underlying function by a polynomial as described by e.g. Press et al. (1992). Fig. 3 shows the smoothed spectra after applying a 65 point Savitzky-Golay filter of order 6 to the data in Fig. 2. The smoothed spectra, in particular that of 474640 (blue), show some weak features that might be tentatively identified as pure methane ice absorption bands (Grundy, Schmitt & Quirico 2002). However, the most prominent methane band at 0.73 µm is observed neither in the spectrum of 474640 nor in that of 2013 RF98 (red). The S/N is insufficient to identify reliably any absorption band, but the spectral slopes in the visible of both ob- jects provide some compositional information. Objects with visible spectral slopes in the range 0 -- 10 %/0.1 µm can have pure ices on their surfaces (like Eris, Pluto, Makemake and Haumea), as well as highly processed carbon. Slightly red slopes (5 -- 15 %/0.1 µm) indicate the possible presence of amorphous silicates as in the case of Trojans (Emery & Brown 2004) or Thereus (Licandro & Pinilla- Alonso 2005). In any case, the objects will not have a surface domi- nated by complex organics (tholins). Differences between the spec- tra of 474640 and 2013 RF98 might be the result of their present-day heliocentric distance. Mechanisms that are more efficient in alter- ing the icy surfaces of these objects at smaller perihelion distances include sublimation of volatiles and micrometeoroid bombardment (Santos-Sanz et al. 2009). 0.4 0.5 0.6 0.8 Wavelength (µm) 0.7 0.9 1.0 Figure 1. Individual visible spectra of (474640) 2004 VN112 (top panel) and 2013 RF98 (bottom panel) normalised to unity at 0.55 µm and offset vertically for clarity. Red lines correspond to the linear fitting in the range 0.5 -- 0.9 µm to compute the spectral slope. Table 1. Observational details of the visible spectra (1800 s each) obtained for (474640) 2004 VN112 and 2013 RF98 (see the text for further details). Target Spec # Date UT Start Airmass Slope (%/0.1 µm) 474640 2013 RF98 1 2 1 2 3 4 09-03-16 09-03-16 04:37 05:07 09-08-16 09-08-16 09-08-16 09-08-16 03:07 03:37 04:07 04:38 1.085 1.072 Mean 1.234 1.181 1.151 1.142 Mean 10.2±0.6 13.3±0.6 12±2 14.9±0.8 12.3±0.8 17.1±0.8 16.1±0.8 15±2 and the associated error is the standard deviation of this mean. The process of dividing by the spectra of the solar analogue stars in- troduces an error of 0.3 %/0.1 µm. The error value shown in the last column of Table 1 is sum of these two contributions. For each target, we averaged individual spectra to obtain the average visi- ble spectra for 474640 and 2013 RF98 shown in Fig. 2; the mean spectral slopes in Table 1 are the means of the individual spectra and their errors, the associated standard deviations. The value of the spectral slope of 474640 is consistent with the one obtained by Sheppard (2010). The values in Table 1 are similar to those of scat- tered disc TNOs, Plutinos, high-inclination classical TNOs as well as the Damocloids and comets (see e.g. table 5 in Sheppard 2010). The spectra in Fig. 2 show an obvious resemblance, but the low S/N makes the identification of absorption features difficult. In order to make a better comparison and assuming that the rel- MNRAS 000, 1 -- 6 (2017) 4 J. de León, C. de la Fuente Marcos and R. de la Fuente Marcos 3.2 Photometry We obtained a total of 3 and 11 acquisition images to identify (474640) 2004 VN112 and 2013 RF98, respectively. Images were taken using the Sloan r′ filter and calibrated using the zero point values computed for the corresponding nights. The resulting r′ magnitudes and their uncertainties are shown in Table A1. 3.3 Astrometry We used the acquisition images to compute the celestial coordinates of each target and improve its orbit. This was particularly relevant for 2013 RF98 that prior to our observations had a rather uncertain orbital determination.2 We found (474640) 2004 VN112 within 1′′of the predicted ephemerides, but 2013 RF98 was found nearly 1.′2 away. Both ETNOs were in Cetus. Astrometric calibration of the CCD frames was performed using the algorithms of the Astrom- etry.net system (Lang et al. 2010). The quality of high-precision astrometry with OSIRIS at GTC matches that of data acquired with FORS2/VLT (Sahlmann et al. 2016). The collected astrometry is shown in Tables A2 and A3. The new orbital solution for 2013 RF98 available from Horizons (as of 2016 December 18 18:51:59 UT) is based on 51 observations spanning a data-arc of 1092 d, its resid- ual rms amounts to 0.′′12: a = 349 ± 11 au, e = 0.897 ± 0.003, i = 29.◦572 ± 0.◦003, Ω = 67.◦596 ± 0.◦005 and ω = 311.◦8 ± 0.◦6, referred to the epoch 2457800.5 JD TDB; the time of perihelion passage is 2455125±95 JED (2009 October 20.7289 UT). The time of perihelion passage for 474640 is 2009 August 25.8290 UT. Us- ing the new orbit, the directions of the perihelia of this pair are separated by 14.◦2±0.◦6, their orbital poles by 4.◦056±0.◦003, and the directions of their velocities at perihelion/aphelion by 14.◦1±0.◦6. 4 ORIGIN OF THE PAIR 474640 -- 2013 RF98: FRAGMENTATION VS. BINARY DISSOCIATION From the spectral analysis discussed above, we have found that the members of the pair (474640) 2004 VN112 -- 2013 RF98 show similar spectral slopes, very different from that of Sedna which has ultra- red surface material (spectral gradient of about 26 %/0.1 µm ac- cording to Sheppard 2010, and 42 %/0.1 µm according to Fornasier et al. 2009) but compatible with those of (148209) 2000 CR105 (spectral gradient of about 14 %/0.1 µm, Sheppard 2010) and 2012 VN113 (spectral gradient of about 13 %/0.1 µm, Trujillo & Sheppard 2014). These five objects have been included in the group of seven singled out as relevant to the Planet Nine hypothesis (Brown & Batygin 2016). Such spectral differences suggest that the region of origin of the pair 474640 -- 2013 RF98 may coincide with that of 148209 and 2012 VN113 but not with Sedna's, which is thought to come from the inner Oort Cloud (Sheppard 2010). Other ETNOs with values of their spectral gradient in Sheppard (2010) are 2002 GB32 (∼17 %/0.1 µm) and 2003 HB57 (∼13 %/0.1 µm). Objects with both similar directions of the orbital poles and perihelia could be part of a group of common physical origin (Öpik 1971). This particular pair of ETNOs is very unusual and a model analogous to the one used by de la Fuente Marcos & de la Fuente Marcos (2014) to study the overall visibility of the ETNO popula- tion predicts that the probability of finding such a pair by chance is less than 0.0002. This model uses the new orbital solutions and assumes an unperturbed asteroid population moving in heliocentric orbits. Following Öpik (1971), there are two independent scenar- ios that could explain this level of coincidence: (1) a large object ) o ( n o i t a r a p e s r a l u g n A 4.5 4 3.5 3 2.5 2 1.5 1 0.5 0 -10 -8 -6 -4 Time (Myr) -2 0 Figure 4. Evolution of the angular separation between the orbital poles of the pair (474640) 2004 VN112 -- 2013 RF98 for three representative test cal- culations with different perturbers. Red: a = 549 au, e = 0.21, i = 47◦, m = 16 M⊕. Blue: a = 448 au, e = 0.16, i = 33◦, m = 12 M⊕. Green: a = 421 au, e = 0.10, i = 33◦, m = 12 M⊕. broke up relatively recently at perihelion and these two ETNOs are fragments, or (2) both ETNOs were kicked by an unseen perturber at aphelion. Sekanina (2001) has shown that minor bodies resulting from a fragmentation episode at perihelion must have very differ- ent times of perihelion passage. The fragmentation episode at per- ihelion can thus be readily discarded as the difference in time of perihelion passage for this pair is less than a year. The second scenario pointed out above implies the presence of an unseen massive perturber, i.e. a trans-Plutonian planet. Close encounters between minor bodies and planets can induce fragmen- tation directly via tidal forces (e.g. Sharma, Jenkins & Burns 2006) or indirectly by exciting rapid rotation (e.g. Scheeres et al. 2000; Ortiz et al. 2012). Alternatively, wide binary asteroids can be eas- ily disrupted during close encounters with planets. The existence of wide binaries among the populations of minor bodies orbiting beyond Neptune is well documented (e.g. Parker et al. 2011). Wide binary asteroids have very low binding energies and can be eas- ily dissociated during close encounters with planets (e.g. Agnor & Hamilton 2006; Parker & Kavelaars 2010). Binary asteroid disso- ciation may be able to explain the properties of this pair of ETNOs, but only if there is a massive unseen perturber orbiting the Sun well beyond Pluto. In order to test the viability of this hypothesis, we have per- formed thousands of numerical experiments following the prescrip- tions discussed by de la Fuente Marcos et al. (2016) and aimed at finding the most probable orbital properties of a putative perturber able to tilt the orbital plane of the pair 474640 -- 2013 RF98 from an initial angular separation close to zero at dissociation to the cur- rent value of nearly 4◦. These simulations involve N-body integra- tions backwards in time under the influence of an unseen perturber with varying orbital and physical parameters (per numerical exper- iment). Our preliminary results indicate that a planet with mass, m, in the range 10 -- 20 M⊕ moving in an eccentric (0.1 -- 0.4) and in- clined (20 -- 50◦) orbit with semimajor axis of 300 -- 600 au, may be able to induce the observed tilt on a time-scale of 5 -- 10 Myr. Per- turbers with m < 10 M⊕ or a > 600 au are unable to produce the desired effect. Fig. 4 illustrates the typical outcome of these cal- culations. A detailed account of these numerical experiments will be presented in an accompanying paper (de la Fuente Marcos, de la Fuente Marcos & Aarseth, in preparation). The orbital param- eters of this putative planet are somewhat consistent with those of the object discussed by Holman & Payne (2016). Super-Earths may form at 125 -- 750 au from the Sun (Kenyon & Bromley 2015, 2016). Our analysis favours a scenario in which 474640 -- 2013 RF98 were MNRAS 000, 1 -- 6 (2017) Visible spectra of (474640) 2004 VN112 -- 2013 RF98 5 once a binary asteroid that became unbound after a relatively recent gravitational encounter with a trans-Plutonian planet at hundreds of au from the Sun. An alternative explanation involving an asteroid break-up near aphelion, also after a close encounter with a planet, is possible but less probable because it requires an approach at closer range, 20 planetary radii versus 0.8 au for binary dissociation. 5 CONCLUSIONS In this Letter, we provide for the first time direct indication of the surface composition of the ETNOs (474640) 2004 VN112 and 2013 RF98. Both objects are too faint for infrared spectroscopy, but our results show that they are viable targets for visible spec- troscopy. The analysis of our results gives further support to the trans-Plutonian planets paradigm that predicts the presence of one or more planetary bodies well beyond Pluto. Summarizing: (i) Our estimate of the spectral slope for 474640 is 12±2 %/0.1 µm and for 2013 RF98 is 15±2 %/0.1 µm. These values suggest that the surfaces of these ETNOs can have pure methane ices (like Pluto) and highly processed carbons, in- cluding some amorphous silicates. (ii) Although the spectra of the pair 474640 -- 2013 RF98 are not perfect matches, the resemblance is significant and the dispar- ities observed might be the result of their different present-day heliocentric distance. (iii) By improving the orbital solution of 2013 RF98, we confirm that the pair 474640 -- 2013 RF98 has unusual relative dynam- ical properties. The directions of their perihelia are separated by 14.◦2 and their orbital poles are 4.◦1 apart. (iv) Our numerical analysis favours a scenario in which 474640 -- 2013 RF98 were once a binary asteroid that became unbound after an encounter with a trans-Plutonian planet at very large heliocentric distance. ACKNOWLEDGEMENTS We thank the anonymous referee for a constructive and detailed report, and S. J. Aarseth for providing one of the codes used in this research and for his helpful comments on binary dissociation. Based on observations made with the Gran Telescopio Canarias; we are grateful to all the technical staff and telescope operators for their assistance with the observations. This work was partially sup- ported by the Spanish 'Ministerio de Economía y Competitividad' (MINECO) under grant ESP2014-54243-R. JdL acknowledges fi- nancial support from MINECO under the 2015 Severo Ochoa Pro- gram MINECO SEV-2015-0548. CdlFM and RdlFM thank A. I. Gómez de Castro, I. Lizasoain and L. Hernández Yáñez of the Uni- versidad Complutense de Madrid (UCM) for providing access to computing facilities. Part of the calculations and the data analysis were completed on the EOLO cluster of the UCM, and CdlFM and RdlFM thank S. Cano Alsúa for his help during this stage. EOLO, the HPC of Climate Change of the International Campus of Excel- lence of Moncloa, is funded by the MECD and MICINN. This is a contribution to the CEI Moncloa. In preparation of this Letter, we made use of the NASA Astrophysics Data System, the ASTRO-PH e-print server and the MPC data server. MNRAS 000, 1 -- 6 (2017) REFERENCES Abbott T. et al., 2005, The Dark Energy Survey, eprint arXiv:astro- ph/0510346 Abbott T. et al., 2016, MNRAS, 460, 1270 Agnor C. B., Hamilton D. P., 2006, Nature, 441, 192 Batygin K., Brown M. E., 2016, AJ, 151, 22 Becker A. C., Arraki K. S., Rest A., Wood-Vasey W. M., Marsden B. G., 2007, MPEC Circ., MPEC 2007-S29 Becker A. C. et al., 2008, ApJ, 682, L53 Brown M. E., 2012, Annu. Rev. Earth Planet. Sci., 40, 467 Brown M. E., Batygin K., 2016, ApJ, 824, L23 Cepa J. et al., 2000, in Iye M., Moorwood A. F. M., eds, Optical and IR Telescope Instrumentation and Detectors. Proc. SPIE, 4008, p. 623 de la Fuente Marcos C., de la Fuente Marcos R., 2014, MNRAS, 443, L59 de la Fuente Marcos C., de la Fuente Marcos R., 2016, MNRAS, 462, 1972 de la Fuente Marcos C., de la Fuente Marcos R., Aarseth S. J., 2015, MN- RAS, 446, 1867 de la Fuente Marcos C., de la Fuente Marcos R., Aarseth S. J., 2016, MN- RAS, 460, L123 de León J. et al., 2016, Icarus, 266, 57 de León J., de la Fuente Marcos C., de la Fuente Marcos R., 2016, MPEC Circ., MPEC 2016-U18 Emery J. P., Brown R. H., 2004, Icarus, 170, 131 Flaugher B. et al., 2015, AJ, 150, 150 Fornasier S. et al., 2009, A&A, 508, 457 Fraser W. C., Brown M. E., 2012, ApJ, 749, 33 Gladman B., Holman M., Grav T., Kavelaars J., Nicholson P., Aksnes K., Petit J.-M., 2002, Icarus, 157, 269 Grundy W., Schmitt B., Quirico E., 2002, Icarus, 144, 486 Holman M. J., Payne M. J., 2016, AJ, 152, 80 Kenyon S. J., Bromley B. C., 2015, ApJ, 806, 42 Kenyon S. J., Bromley B. C., 2016, ApJ, 825, 33 Landolt A. U., 1992, AJ, 104, 340 Lang D., Hogg D. W., Mierle K., Blanton M., Roweis S., 2010, AJ, 139, 1782 Licandro J., Pinilla-Alonso N., 2005, ApJ, 630, L93 Morate D., de León J., De Prá M., Licandro J., Cabrera-Lavers A., Campins H., Pinilla-Alonso N., Alí-Lagoa V., 2016, A&A, 586, A129 Morbidelli A., Levison H. F., 2004, AJ, 128, 2564 Öpik E. J., 1971, Irish Astronomical Journal, 10, 35 Ortiz J. L. et al., 2012, MNRAS, 419, 2315 Parker A. H., Kavelaars J. J., 2010, ApJ, 722, L204 Parker A. H., Kavelaars J. J., Petit J.-M., Jones L., Gladman B., Parker J., 2011, ApJ, 743, 1 Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P., 1992, Nu- merical Recipes in FORTRAN. Cambridge Univ. Press, Cambridge, p. 644 Sahlmann J., Lazorenko P. F., Bouy H., Martín E. L., Queloz D., Ségransan D., Zapatero Osorio M. R., 2016, MNRAS, 455, 357 Santos-Sanz P., Ortiz J. L., Barrera L., Boehnhardt H., 2009, A&A, 494, 693 Savitzky A., Golay M. J. E., 1964, Analytical Chemistry, 36, 1627 Scheeres D. J., Ostro S. J., Werner R. A., Asphaug E., Hudson R. S., 2000, Icarus, 147, 106 Sekanina Z., 2001, Publications of the Astronomical Institute of the Academy of Sciences of the Czech Republic, 89, 78 Shankman C., Kavelaars J. J., Lawler S. M., Gladman B. J., Bannister M. T., 2016, AJ, submitted (eprint arXiv:1610.04251) Sharma I., Jenkins J. T., Burns J. A., 2006, Icarus, 183, 312 Sheppard S. S., 2010, AJ, 139, 1394 Sheppard S. S., Trujillo C. A., 2016, AJ, 152, 221 Trujillo C. A., Sheppard S. S., 2014, Nature, 507, 471 6 J. de León, C. de la Fuente Marcos and R. de la Fuente Marcos Table A1. Observational details of the acquisition images obtained to iden- tify the targets. This paper has been typeset from a TEX/LATEX file prepared by the author. Image # Date UT Start Airmass Exp. Time Zero point (s) (mag) 1 2 3 1 2 3 4 5 6 7 8 9 10 11 09-02-16 09-03-16 09-03-16 02:16 04:21 04:32 1.580 1.120 1.100 09-05-16 09-06-16 09-06-16 09-06-16 09-06-16 09-06-16 09-07-16 09-08-16 09-08-16 09-08-16 09-08-16 02:43 02:55 03:05 03:23 03:32 03:37 03:40 02:46 02:51 02:56 03:00 1.380 1.320 1.290 1.240 1.220 1.210 1.210 1.330 1.310 1.290 1.280 120 120 60 180 120 120 120 120 120 180 120 180 120 120 29.240 29.277 29.277 29.281 29.287 29.287 29.287 29.287 29.287 29.287 29.287 29.287 29.287 29.287 r′ (mag) (474640) 2004 VN112 23.63±0.10 23.58±0.05 23.51±0.05 2013 RF98 24.57±0.10 24.39±0.20 24.42±0.15 24.51±0.20 24.45±0.05 24.53±0.20 24.53±0.05 24.56±0.07 24.55±0.10 24.53±0.05 24.49±0.10 Table A2. Observations of (474640) 2004 VN112. All the observations in the r′ filter. Date (UT) RA(J2000) Dec(J2000) (h:m:s) (◦:′:′′) 2016 09 02.09514 2016 09 03.18233 2016 09 03.18923 03:07:52.51 03:07:51.41 03:07:51.40 +07:38:28.1 +07:38:17.7 +07:38:17.5 Table A3. Observations of 2013 RF98. All the observations in the r′ filter. Date (UT) RA(J2000) Dec(J2000) (h:m:s) (◦:′:′′) 2016 09 05.114385 2016 09 06.122606 2016 09 06.129458 2016 09 06.142113 2016 09 06.148191 2016 09 06.151823 2016 09 07.154454 2016 09 08.116547 2016 09 08.120114 2016 09 08.123483 2016 09 08.126326 02:49:09.83 02:49:07.83 02:49:07.81 02:49:07.79 02:49:07.78 02:49:07.77 02:49:05.70 02:49:03.61 02:49:03.60 02:49:03.60 02:49:03.59 −00:10:52.0 −00:11:14.2 −00:11:14.5 −00:11:14.7 −00:11:14.7 −00:11:14.8 −00:11:37.3 −00:11:59.1 −00:11:59.1 −00:11:59.3 −00:11:59.4 APPENDIX A: PHOTOMETRY AND ASTROMETRY DATA TABLES The r′ magnitudes from the acquisition images and their uncertain- ties are shown in Table A1. The astrometry submitted to the Minor Planet Center (MPC) is shown in Tables A2 and A3 (de León, de la Fuente Marcos & de la Fuente Marcos 2016). MNRAS 000, 1 -- 6 (2017)
1509.06773
1
1509
2015-09-22T20:33:55
Surface ages of mid-size Saturnian satellites
[ "astro-ph.EP" ]
The observations of the surfaces of the mid sized Saturnian satellites made by Cassini Huygens mission have shown a variety of features that allows study of the processes that took place and are taking place on those worlds. Research of the Saturnian satellite surfaces has clear implications for Saturn history and surroundings. In a recent paper, the production of craters on the mid sized Saturnian satellites by Centaur objects was calculated considering the current Solar System. We have compared our results with crater counts from Cassini images and we have noted that the number of observed small craters is less than our calculated number. In this paper we estimate the age of the surface for each observed terrain on each mid sized satellite of Saturn. We have noticed that since there are less observed small craters than calculated (except on Iapetus), this results in younger ages. This could be the result of efficient endogenous or exogenous process(es) for erasing small craters and or crater saturation at those sizes. The size limit from which the observed number of smaller craters is less than the calculated is different for each satellite, possibly indicating processes that are unique to each, but other potential common explanations would be crater saturation and or deposition of E ring particles. These processes are also suggested by the findings that the smaller craters are being preferentially removed, and the erasure process is gradual. On Enceladus, only mid and high latitude plains have remnants of old terrains; the other regions could be young; the regions near the South Polar Terrain could be as young as 50 Myr old. On the contrary for Iapetus, all the surface is old and it notably registers a primordial source of craters. As the crater size is decreased, it would be perceived to approach saturation until D less than 2 km craters, where saturation is complete.
astro-ph.EP
astro-ph
Surface ages of mid-size Saturnian satellites Romina P. Di Sistoa,b, Macarena Zanardib aFacultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata bInstituto de Astrof´ısica de La Plata, CCT La Plata-CONICET-UNLP, Paseo del Bosque S/N (1900), La Plata, Argentina Abstract The observations of the surfaces of the mid-sized Saturnian satellites made by Cassini-Huygens mission have shown a variety of features that allows study of the processes that took place and are taking place on those worlds. Research of the Saturnian satellite surfaces has clear implications not only for Saturn's history and Saturn's surroundings, but also for the Solar System. Crater counting from high definition images is very important and could serve for the determination of the age of the surfaces. In a recent paper, we have calculated the production of craters on the mid-sized Saturnian satellites by Centaur objects considering the current configuration of the Solar System. Also, we have compared our results with crater counts from Cassini images by other authors and we have noted that the number of observed small craters is less than our calculated theoretical number. In this paper we estimate the age of the surface for each observed terrain on each mid-sized satellite of Saturn. All the surfaces analyzed appear to be old with the exception of Enceladus. However, we have noticed that since there are less observed small craters than calculated (except on Iapetus), this results in younger ages than expected. This could be the result of efficient endogenous or exogenous process(es) for erasing small craters and/or crater saturation at those sizes. The size limit from which the observed number of smaller craters is less than the calculated is different for each satellite, possibly indicating processes that are unique to each, but other potential common explanations for this paucity of small craters would be crater saturation and/or deposition of E- ring particles. These processes are also suggested by the findings that the smaller craters are being preferentially removed, and the erasure process is Email address: [email protected] (Romina P. Di Sisto) Preprint submitted to Icarus August 15, 2018 gradual. On Enceladus, only mid and high latitude plains have remnants of old terrains; the other regions could be young. In particular, the regions near the South Polar Terrain could be as young as 50 Myrs old. On the contrary for Iapetus, all the surface is old and it notably registers a primordial source of craters. As the crater size is decreased, it would be perceived to approach saturation until D . 2 km-craters, where saturation is complete. Keywords: Saturn, satellites; Cratering; Centaurs 1. Introduction The surfaces of mid-sized Saturnian satellites are a kind of laboratory where it is possible to observe and investigate the physical and dynamic processes that have been taking place around Saturn and also throughout the Solar System. The study of impact craters on the satellite surfaces allows us to better understand what could be happening in the Saturn environment to produce these craters. The Saturnian satellite system was observed and studied in the past by Pioneer 11 and Voyager 1 and 2 spacecrafts, greatly increasing the knowledge of the Saturn system. Voyager images revealed diverse satellite surfaces (Smith et al., 1981, 1982). Crater counts on those images indicated that Saturnian satellites have been variably cratered, which suggest different geologic histories (Plescia and Boyce 1982, 1983, 1985). At present, the Cassini-Huygens mission is visiting the Saturn system, and the detailed observations it renders provide us with new paradigms and physical processes to understand and interpret. The mid-sized icy satellites of Saturn are Mimas, Enceladus, Tethys, Dione, Rhea, and Iapetus. They are regular satellites, mainly composed of water ice and in synchronous rotation. The Cassini-Huygens mission has observed all of them in detail allowing scientists to obtain accurate infor- mation on the shapes, mean radii and densities (Thomas, 2010) and gravity fields (Jacobson et al., 2006). Furthermore, some of these satellites show traces of physical activity and renovation, possibly due to recent geologi- cal processes. The geologic activity could include endogenous activity such as viscous relaxation, volcanism, and/or tectonic or even atmospheric pro- cesses. Cratering itself is a potential process in which the formation of large craters could remove small craters by ejecta emplacement and seismic shak- ing. Also, exogenous processes, such as in fall from the E-ring or debris rings, 2 could erase surface features. Those physical processes would renew the satellite surfaces, erasing old or young surface features such as craters. The most striking case is Ence- ladus whose surface differs markedly from region to region and has present active geysers emanating from four parallel fractures in the south polar re- gion, called "tiger stripes" (Porco et al., 2006). This region is a very active one where the surface is modified, erasing the craters (totally or partially). On Dione, in turn, Cassini observations reveal two different surfaces: the cratered plains (cp), heavily cratered, and the smoother plains (sp), with a lower cratering suggesting a younger surface. On Rhea, measurements by Cassini spacecraft detected a tenuous atmosphere of oxygen and carbon diox- ide (Teolis et al., 2010). Iapetus is the opposite case of Enceladus as it has heavily cratered plains with large degraded basins, which indicates that its surface is ancient. What is more, the count of small craters showed a size distribution indicative of crater saturation (Denk et al., 2010). Kirchoff and Schenk (2009) and (2010), hereafter KS09 and KS10, an- alyzed high-resolution Cassini images and obtained the number and size- frequency distribution (SFD) of craters for the mid-sized icy satellites. With their observed number of craters and the previous cratering rate estimations by Zahnle et al. (2003), they calculated the surface ages of each satellite for some crater diameters. There are a number of factors that must be taken into account in the analysis of the age of a satellite area. First, all possible impactor popula- tions should be considered. The analysis of the images obtained by Voyager (Smith et al., 1981, 1982) implied that the satellites were struck by two dif- ferent impactor populations: Population I, which produced a greater number of large craters (bigger than 20 km), and Population II, which produced a greater number of smaller craters (smaller than 20 km) (Smith et al., 1982). The origin or even the existence of both populations is disputed. Smith et al. (1981, 1982) suggested that Population I was the tail-off of a postaccretional heavy bombardment, while Population II has the form expected for collisional debris from the satellites or other orbiting debris. Horedt and Neukum (1984) concluded that cratering on Saturnian satellites is produced by heliocentric objects as well as by planetocentric impactors. On the other hand, Hartman (1984) noted that crater densities on heavily cratered surfaces throughout the Solar System are all similar due to a "saturation equilibrium". He ar- gued that only one single population of heliocentric planetesimals and their fragments had been recorded. Dobrovolskis and Lissauer (2004) studied the 3 fate of ejecta from the irregularly shaped satellite, Hyperion, suggesting that it does contribute to Population II craters on the inner satellites of Saturn. However, those particles would produce craters with a different morphology than those produced by a heliocentric source. Moreover, consideration of the primordial situation is important for the study of the origin of craters. It is believed that the mass of the primor- dial trans-Neptunian zone, the source of Centaurs, was ∼ 100 times higher than the present one (Morbidelli et al., 2008). This mass might have been depleted by a strong dynamical excitation of the trans-Neptunian region. There were several models that described the mass depletion; for example, in the "Nice Model" and its subsequent versions, the interaction between the migrating planets and planetesimals destabilized the planetesimal disk and scattered the planetesimals all over the Solar System before the time of the LHB (Tsiganis et al., 2005; Levison et al., 2008). Accordingly, there was much primordial mass that struck the planets and their satellites very early in the Solar System history. It would be expected that this event marked the surfaces of the satellites, mainly producing a great number of larger larger craters than at present. Those craters are probably the larger ones that can be observed on the satellites. Moreover, there are papers (Nervorny et al., 2003, 2007) that argue in favor of more primordial irregular satellites and a more active primordial collisional activity around the major planets. There would even be populations of large irregular satellites and debris of catas- trophically disrupted satellites that may have played important roles in the history of the Saturn system (Dones et al., 2007). A primordial crater contribution to the Saturn system could also be con- nected with the crater saturation observed on some of the Saturnian satellites. The formation of craters, especially the larger ones, is a process that is effec- tive in erasing small craters, not only through crater formation, but also by the ejecta blanket and seismic shaking. When a surface is so heavily bom- barded that the formation of craters is equaled by the obliteration of craters, it reaches a crater saturation equilibrium. In this case, the density of craters on a surface does not change proportionally. This process is then critical to the determination of the source of impactors, geological processes, ages, etc (Hartmann and Gaskell, 1997). Crater saturation equilibrium finally causes a distinctive cumulative size frequency distribution (SFD) of craters. Gault (1970) obtained a −2 power-law, from a model based on first principles and laboratory experiments. Hartmann (1984) fitted a −1.83 power-law for the cumulative SFD to crater count data of the surface of various Solar System 4 objects. Richardson (2009) developed a detailed cratered terrain evolution model and observed the way in which crater densities attain equilibrium con- ditions. He found that if the impactor population has a cumulative power-law slope of < −2, crater densities reach a cumulative power-law of about −2; and if the impactor population has a cumulative power-law slope of > −2, crater density equilibrium values follow the shape of the production population. Squyres et al. (1997) argued that cratering on a surface is a random process and as crater obliteration becomes more important, the spatial distribution of craters tends to be uniform. They used both the spatial distribution and size-frequency distribution of craters to study, with statistical techniques, how a cratered surface approaches saturation. They found that at least 25% of the craters on Rhea and Callisto were destroyed by subsequent oblitera- tion. Therefore, crater saturation is especially important on highly cratered surfaces that are in general old. In the supposedly young surfaces, as in the case of Enceladus for exam- ple, geological processes are erasing craters. However, such processes compete with possible exogenous processes, like particle deposition on the surface that can fill in the craters, finally removing them. Mid-sized Saturnian satellites with the exception of Iapetus are embedded in the Saturn E-ring; therefore their particles cross the orbits of the satellites with some probability of im- pact. Enceladus's plumes are thought to be the source and maintenance of E-ring material (Hamilton and Burns, 1994, Kempf et al., 2010). A frac- tion of the plume particles escape to populate the E-ring but other fraction returns to Enceladus hitting its surface (Kempf et al., 2010). Ingersoll and Ewald, (2011) estimated (12 ± 5.5)108 kg for the mass of particles in the E-ring and a particle lifetime in the E-ring of about 8 years. This mate- rial could be an exogenous source causing erasure of craters in the mid-sized Saturnian satellites. In a recent paper, Di Sisto and Zanardi (2013), hereafter DZ13, calculated the production of craters on the mid-sized Saturnian satellites produced by current Centaur objects coming from the Scattered Disk (SD) and plutinos, in the trans-Neptunian region, and compared these calculations with the Cassini observations made by KS09 and KS10. Also obtained was the current cratering rate on each satellite. In that paper, the authors concluded that since the number of observed small craters is lower than their calculated theoretical number, and because there are physical processes on at least some satellite surfaces that erode them, such difference is likely to be caused by those processes. It seems that the satellite surfaces were "reset" by those 5 physical processes and were, therefore, younger and defined by the current geological processes. However, as mentioned, a number of factors has to be considered to test what processes are really taking place on the satellites and how these affect the age of their surfaces. Therefore, in this paper we analyze in detail our theoretical predictions about the current cratering production to obtain satellite surface ages. These age determinations are not only a first approximation but also a way to help determine the time scale of the dynamic and geological processes that are taking place on the mid-sized satellites of Saturn. Since each satellite is a unique world, it is also interesting to test pecu- liarities and different processes on each satellite in our study. 2. Age calculation For the calculation of the number of impact craters and ages on the mid-sized Saturnian satellites all potential populations of impactors should be considered. Those impactors could be main-belt asteroids, Jupiter and Neptune Trojans, Jupiter Family Comets (JFCs) and Centaurs from the SD, irregular satellites, planetocentric bodies and Nearly Isotropic Comets (NICs) (Dones et al. 2009). There is evidence, such as recent collisions and passages of comets by Jupiter and the observations of Saturn-crossing comets and Centaurs (Dones et al., 2009) as well as also previous work by Zahnle et al. 1998 and 2003, that Centaurs and JFCs dominate impact cratering in the Outer Solar System. The only truly unknown potential impactor population is the planetocentric one. This hypothetical source has been discussed by previous papers (see Sect. 1) but it has not been observed so far; its possible contribution or existence will be analyzed for each satellite with our results. Di Sisto and Zanardi (2013) calculated the production of craters on the mid-sized Saturnian satellites by Centaur objects from SD and plutinos. They found that the contribution of plutinos is negligible with respect to Scattered Disk Objects (SDOs). They used a method previously developed by Di Sisto and Brunini (2011), hereafter DB11. Both papers are based on the numerical simulation carried out by Di Sisto and Brunini (2007), who studied the evolution of SDOs for 4.5 Gyrs when they enter the Centaur zone, considering the current configuration of the Solar System. Di Sisto and Brunini (2007) studied only Centaur objects when they enter the giant planet zone from the SD; they stopped their simulation when Centaurs enter the JFCs zone. Thus, the contribution of JFCs to the satellite cratering is 6 not considered here. However, only a ∼ 22% of SDOs enter the JFC zone (Di Sisto and Brunini, 2007). These objects were studied by Di Sisto et al. (2009) through a numerical simulation that included a complete dynamical- physical model. Only a fraction of the initial comets return to the Centaur zone, depending on the size of the comet (a JFC reduces its radius due to sublimation and splitting). We can then assume that this contribution should be approximately one order of magnitude smaller than that of the Centaurs when they enter the Saturn zone from the SD. DZ13 and DB11 used the output files of the encounters of SDOs with Saturn when they enter the Centaur zone (from Di Sisto and Brunini, 2007) to calculate the cumulative number of craters (Nc(> D)) produced by current Centaurs on the Saturnian satellites. It is generally accepted that the mass of the primordial trans-Neptunian zone, the source of centaurs, was ∼ 100 times higher than the present one (Morbidelli et al. 2008 ) and decreased to its present value in at most 1 Gyr ("Nice Model"). This early phase of the Solar System was not modeled in the numerical simulation by Di Sisto and Brunini (2007), which instead started after the initial mass depletion, when the Solar System began to stabilize, and essentially becomes like the current configuration of the Solar System. As mentioned, this simulation was run for 4.5 Gyrs, since the exact duration of this early phase is not known; but it should be representative of the post initial mass depletion time, namely . 4 Gyr. Then, Nc(> D) corresponds to the number of craters produced since the change in configuration of the Solar System predicted by the Nice Model (∼ 4 Ga). In DZ13, we calculated the Nc(> D) for each satellite and for two cases of the SFD of SDOs. That is, we considered a SFD of SDOs as a power law with a differential index s1 = 4.7 for diameters d > 60 km (from Elliot et al., 2005). The SFD breaks for d . 60 km, but given the uncertainty in the SFD for small objects, we considered two values of the differential index: s2 = 2.5 and s2 = 3.5 (the Dohnanyi exponent). In DZ13, we also compared our theoretical number of craters with the number of observed craters obtained by KS09 in Figs. 2-7 of their paper. We noticed that our calculated number of craters for s2 = 3.5 is in general closer to the observed number by KS09. However, for small craters, the observed number of craters is lower than the calculated one. Since in general there would be physical processes on mid- sized Saturnian satellite surfaces that erode them, these processes could erase small craters. The extent of this erasure is specific to each satellite because it depends on the process that is taking place on each one. It is possible, then, 7 to account for this process by calculating the age of each satellite surface. The calculation of the age of each terrain is as follows: first, let us assume (on the basis of previous comments)that the Centaurs from the SD are the main source of craters in the current configuration of the Solar System and that, given the similarity with observations mentioned before, the calculated number of craters is the one obtained from the differential index: s2 = 3.5 of the SFD of small SDOs. This assumption can be considered a first step in the determination of ages. Since Centaurs come mainly from the SD (Di Sisto and Brunini, 2007), they should be the most important current heliocentric source of craters. It should be highlighted that a possible planetocentric source of craters has been excluded since there are no estimations or calculations from this hypothetical source. From DZ13, we know the total number of craters on each satellite greater than a given diameter D, produced by Centaurs from the SD in the current configuration of the Solar System (Nc(> D)). We took the observed number of craters No(> D) from KS09. Then, if No(> D) > Nc(> D), we can state that the terrain has preserved craters that were produced at earlier times of the Solar System, or that there have been other important sources of craters (such as a planetocentric source), and/or there is no current geological activity that could erode the craters. But if No(> D) < Nc(> D), there are Nc(> D) − No(> D) craters that have been erased and then there must be a geological process taking place on the satellite surface. The comparison between the calculated and the observed numbers can be seen in Figs. 2- 7 in DZ13, where it is noticeable that there is a range of crater diameters (different for each satellite) for which No(> D) < Nc(> D). Therefore, those terrains could have a geological activity that has eroded the craters, and the surface would be young. To calculate the age of those surfaces, the temporal dependence of crater- ing was used. According to the method developed by DB11, the cumulative number of craters produced by Centaurs is proportional to the number of encounters of Centaurs with Saturn. Then, the temporal dependence of cra- tering is equal to that of encounters with Saturn (see Eqs. 2-7 in DB11). In Fig. 1 of this paper, the fraction of encounters with Saturn (number of encounters for a given time with respect to the total number of encounters) was plotted as a function of time and found that the whole plot could be fitted by a log-function given by: F (t) = a log t + b, (1) 8 s r e t n u o c n e f o n o i t c a r F 1 0.8 0.6 0.4 0.2 0 0 500 1000 1500 2000 2500 3000 3500 4000 4500 t [Myrs] Figure 1: Fraction of encounters of SDOs with Saturn versus time and the fit to the data. where a = 0.198406 ± 0.0002257 and b = −3.41872 ± 0.004477. Then, the cumulative number of craters on a given satellite depending on time can be obtained from: Nc(> D, t) = F (t)Nc(> D), (2) Differentiating this equation with respect to time, we obtain the cratering rate over time from: C(> D, t) = a t Nc(> D), (3) Zanhle et al. (1998) calculated cratering rates on the Galilean satellites considering the contribution of several heliocentric sources. They also cal- culated the age of the satellite surfaces. To such end, they considered that the cratering rate on each satellite decreases with time as t−1 from a numer- ical simulation developed by Holman and Wisdom (1993) who found that a collisionless Kuiper Belt dissipates as t−1. This is the same behavior of our calculated cratering rate of Eq. 3, since the original simulation by Di Sisto 9 and Brunini (2007) deals with the dynamical evolution of TNOs. Then, we follow a development analogous to Zanhle et al. (1998) to find the age of a satellite surface but now considering Eq. (3) for the cratering rate. Thus, the estimated cratering timescale or the "age" of the surface τ would be given by: τ (> D) = t0(1 − e− No(>D) aNc(>D) ), (4) where t0 = 4.5 Gyrs is the age of the Solar System. τ represents the model age of the surface for a given diameter, because it is a measure of the cratering timescale, taking into account the production of craters and the erosion of the terrain. In the following sections, we show our results for the mid-sized Saturnian satellites. 3. Results As mentioned in the previous section, considering that Centaurs are the current main source of craters, if No(> D) < Nc(> D), there are Nc(> D) − No(> D) craters that could have been erased. This difference can be plotted against D to get an understanding of the magnitude of the process on each satellite. In Fig. 2, we plot Nc(> D) − No(> D) versus D for Mimas, Tethys, Dione and Rhea considering No(> D) from KS09. For all those satellites, the smaller D is, the greater the difference between observation and our model. Notice also the different slopes of the curves. Tethys and Mimas have more flattened curves meanwhile the curves for Dione and Rhea are steeper. Therefore, for Tethys and Mimas, there seems to be a process that is acting for a wider range of diameters than for Rhea and Dione. We calculate the age depending on D from Eq. 5 and plot the results for Mimas, Tethys, Dione and Rhea in Fig. 3. These age-curves correspond to the terrains analyzed by KS09. In all the curves, there seems to be a noticeable correlation between older age and diameter; that is, smaller craters are younger than greater ones. As can be seen, only small craters would be young or, at least, younger than 4.5 Gyrs old. This means that there could be a geological or other physical process that gradually erases craters and modifies the satellite surface; naturally, smaller craters are cleared before greater ones. Furthermore, the slopes of the age curves are different for each satellite. This fact could imply either that each satellite presents different 10 2 m k 6 0 1 r e p o N - c N 1e+09 1e+08 1e+07 1e+06 100000 10000 1000 100 10 1 Tethys Dione-cp Dione-sp Mimas Rhea 2 4 6 8 10 12 14 16 18 20 D [km] Figure 2: Difference between the cumulative number of theoretical craters and observed craters for Tethys, Dione, Mimas and Rhea according to D. geological processes or exogenous erasing agents (like deposition of E-ring particles), or at least, that each process has a different temporal scale. It could also be possible that on a given satellite surface the equilibrium crater saturation had been reached. There also exists a correlation between the size range or erased craters and distance to Saturn. In the ranges of the diameters plotted, Mimas, the innermost mid-sized satellite located in the inner edge of the E-ring, has a wider range of diameters of erased craters than Tethys, which is farther away from Saturn. Then follows Dione in distance to Saturn, and finally Rhea, at the end of the E-ring. This correlation is very interesting, and potentially implies that one can associate a modification surface agent that depends on the distance to the planet. Using Eq. 5, there is a certain diameter D from which No(> D) gets greater than Nc(> D) and then e− aNc(>D) approaches 0. Therefore, the cal- culated ages trend asymptotically to the age of the Solar System and for those diameters, the surface would be old. As mentioned in the previous No(>D) 11 ] r y [ e g A 4.5e+09 4e+09 3.5e+09 3e+09 2.5e+09 2e+09 1.5e+09 1e+09 5e+08 0 Tethys Dione-cp Dione-sp Mimas Rhea 2 4 6 8 10 12 14 16 18 20 D [km] Figure 3: Age of the surface of Tethys, Dione, Mimas and Rhea according to D. section, when No(> D) > Nc(> D), there are more craters on the surface of the satellite than our calculated number. Since Nc(> D) represents the crater contribution from Centaurs in the current configuration of the Solar System, there are craters that were produced earlier in the Solar System history and/or there is another source of craters, such as planetocentric ob- jects. The limit diameters beyond which No(> D) & Nc(> D) (or before which No(> D) < Nc(> D)) are presented in table 1; it can be seen that those limit diameters are different for each satellite and might be associated with different kinds of processes. Therefore, we cannot evaluate other crater sources for all satellites in general, but we can analyze all our results for each satellite in separate subsections. 3.1. Iapetus Iapetus is Saturn's third-largest moon after Titan and Rhea. Its orbit around Saturn has a semimajor axis of 3.56 ×106 km (∼ 59 Saturn radii), low eccentricity and a slightly high inclination of 8.3◦. Cassini images revealed a near equatorial ridge system that extends for more than 110◦ in longitude 12 Table 1: Limit diameter beyond which No(> D) & Nc(> D)) Satellite Mimas Enceladus Tethys Dione-cp Dione-sp Rhea Iapetus D [km] 15 none 25 5 17 5 2 and that rises more than 20 km above the surrounding plains (Porco et al., 2005). The ridge morphology is consistent with an endogenous origin (Ip, 2006). Giese et al. (2008) found substantial topography on Iapetus' leading side from Cassini images with heights in the range of −10 km to +13 km. A distinctive feature of Iapetus is its global albedo dichotomy. The leading side is dark with albedo values of ∼ 0.04 on a roughly elliptical area (named Cassini Regio after the astronomer Jean Cassini who first noticed it); the trailing side, in turn, is relatively bright with albedos of ∼ 0.6. Two classes of theories have been suggested to account for the origin of this dichotomy: the endogenous ones (Smith et al., 1981, 1982) and the exogenous theories which suggest that debris or dust material from the external moons of Saturn impact the dark hemisphere (Cruikshank et al., 1983; Buratti and Mosher, 1995). A model including both endogenous and exogenous causes was developed by Spence and Denk (2010) which demonstrates that the dichotomy can be explained by runaway global thermal migration of water ice, triggered by the deposition of dark material on the leading hemisphere. From a dynamical point of view, the routes of possible dust impactors on Iapetus were studied by Leiva and Briozzo (2013). They analyzed low-energy incoming dust particles and obtained their distribution on the surface of Iapetus. There is also a global color dichotomy found on Iapetus (Denk et al., 2010), the leading side being redder than the trailing side. Spencer and Denk (2010) provided evidence for an exogenous origin for the redder leading-side parts and suggested that it initiated the thermal formation of the global albedo dichotomy. Iapetus is heavily cratered, with degraded basins with diameters nearly the size of Iapetus's radius, which indicate that both the bright and the 13 dark areas are ancient. Plescia and Boyce (1983) analyzed Voyager 2 images, determining the number of craters at large diameters (because of low image resolution) that indicated that the bright terrain is an ancient surface that dates to the LHB (Plescia and Boyce, 1985). Denk et al. (2010) plotted the cumulative crater size-frequency distribution of Iapetus No(> D), combining five individual measurements (from Cassini images) that we reproduce in Fig. 4 (extrapolating their counts to the complete Iapetus surface) together with Plescia and Boyce (1983) counts. In this figure, we also plotted the cumulative number of craters on Iapetus greater than a given diameter D, produced by current Centaurs from the SD (Nc(> D)) calculated from our model. As mentioned, we assumed a SFD of SDOs that breaks for d . 60 km considering two values of the differential index: s2 = 2.5 and s2 = 3.5. The two resulting curves are plotted in Fig. 4. As noticed by DZ13 for the other satellites, our theoretical number of craters for s2 = 3.5 (green curve) is closer to the observed number. Moreover, both Nc(> D) and No(> D) are similar for D . 2 km and for D & 2 km No(> D) > Nc(> D). We already noticed in DZ13, comparing our results with those of KS10, that No(> 5 km) > Nc(> 5 km), which is consistent with an old surface. Denk et al. (2010) found that for D . 5 − 10 km, the crater cumulative SFD follows a −2 power law that corresponds to crater equilibrium saturation. In fact, by fitting power-laws of the form No(> D) = cDb to the observed crater distributions plotted in Fig. 4, we found for each terrain and ranges of diameters, different values of b. These are shown in Table 2. In the anti-Saturn Hemisphere region, Iapetus-dark and even large basins, there seems to be a tendency of a steeper slope for greater sizes. There- fore, considering that for D . 2 km, the crater distribution is saturated, a decrease of b for larger craters might suggest that larger craters are not likely saturated. If this interpretation were correct, then the crater distri- bution would reflect the impactor distribution only for large craters. Our theoretical crater distribution Nc(> D) with s2 = 3.5 (an impactor popula- tion with a cumulative SFD with an index of 2.5) has a power-law index of b = −2.976 for D & 0.08 km. Hence, only for dark terrain and D > 10 km of the anti-Saturn hemisphere, the observed crater distribution approaches our theoretical crater distribution. On the bright hemisphere, however, the observed crater power-law index is −1.7 from Cassini counts and −2.02 from Voyager counts, a much flatter slope than the theoretical law. On the other hand, from Fig. 4 we note that for D & 5 km, No(> D) > Nc(> D) with a difference ∼ 10 to ∼ 30 times greater. Besides, ac- 14 Table 2: Power-law indexes of the observed cumulative crater distribution obtained by fitting to the observed craters by Denk et al. (2010), Plescia and Boyce (1983) and KS10 Region Anti-Saturn hemisphere Center of basin Engelier Large craters on trailing side Basins (global) Iapetus-dark (KS10) Iapetus-bright (KS10) Iapetus-bright (PB83) D-range [km] 5 < D < 35 8 < D < 35 10 < D < 35 2 < D < 14 17 < D < 202 70 < D < 350 350 < D < 684 4 < D < 80 4 < D < 65 20 < D < 142 b -2.2 -2.36 -2.47 -1.66 -1.7 -0.92 -3.48 -2.66 -1.7 -2.02 cording to our calculations, the greatest crater produced by current centaurs has D ∼ 100 km, but there is observed a significant number of craters and basins greater than this diameter. Therefore, since our model represents the present contribution to cratering, and a current planetocentric population of big objects near Iapetus is not expected, the main crater source of Iapetus must be primordial, which is consistent with previous studies (Plescia and Boyce, 1983, 1985). As mentioned, it is believed that the mass of the pri- mordial trans-Neptunian zone was ∼ 100 times higher than the present one (Morbidelli et al., 2008), in agreement with the differences found between observation and present modeled contribution. Therefore, from the comparisons of our model and observations, we can conclude that big craters on Iapetus are old and connected with the primor- dial epoch of the Solar System. As the crater size decreases, it seems to approach saturation until D . 2 km-craters, where saturation is complete (Denk et al., 2010). This crater saturation does not allow us to evaluate the possibility of another current source of craters such as planetocentric small objects. The different observed power-laws between bright and dark terrains, which may be an artifact of analyzing small areas (Kirchoff and Schenk, 2010), cannot be explained by our model. If real, this difference could be connected to primordial times or also to probable apex-antapex asymmetry (Zhanle et al., 2003). 15 1e+09 1e+08 1e+07 1e+06 100000 10000 1000 100 10 1 s r e t a r c f o r e b m u n e v i t a u m u C l 0.1 0.1 1 10 D [km] 100 1000 Figure 4: Cumulative number of craters of Iapetus with respect to D. Points represent the number of craters obtained by Denk et al. (2010) and Plescia and Boyce (1983) (the gray ones), the red curve represents the cumulative number of craters obtained by our model for s2 = 2.5, and the green one for s2 = 3.5. Furthermore, an interesting result obtained from our model is the recent (2010) identified brighter-than- production of small craters. Denk et al. average craters on highest-resolution images on the dark side, with diameters of up to ∼ 200 m, the brightest one being the crater Escremiz, which is ∼ 4 times brighter than its surroundings. They estimated crater ages assuming that brightness and age are correlated and that the darkening process acts uniformly for all fresh craters. They also stated that a new crater is ∼ 10 times brighter than the surrounding dark terrain and that it declines to half brightness in ∼ 10000 years. Therefore, crater Escremiz is estimated to be ∼ 10000 years old. Moreover, they estimated that on any of Iapetus' dark hemisphere, the largest crater with an age similar to Escremiz should have D ∼ 200 m, and that slightly more than 100 craters of 60 m or larger and with an age similar to or younger than Escremiz should exist. From our model, it is possible to calculate the time the production of craters with a diameter greater than a given D. From Eqs. 2 and 1 and considering that 16 Cassini Regio covers ∼ 40% of Iapetus surface, we found that the last crater greater than D was produced ∆t years ago given by: ∆t(> D) = t0(1 − e− 1 a0.4Nc (>D) ), (5) Then, the last crater with D > 60 m in Cassini Regio was produced ∼ 30 years ago, and the last 100 craters with D > 60 m in Cassini Regio were produced in the last ∼ 3000 years. The last crater with D > 200 m in Cassini Regio was produced ∼ 800 years ago. Those ages are somewhat smaller than the Denk et al.'s (2010) predictions. This might suggest that our calculations of the contribution of Centaurs to the recent cratering are more numerous than expected. However, these differences could be related to the probable apex-antapex asymmetry (Zhanle et al., 2003), which is not considered in this model, or to the intrinsic errors of the current number of SDOs at small sizes or even to errors in Denk et al.'s (2010) model of age-brightness correlation. 3.2. Mimas Mimas is the innermost mid-sized Saturnian satellite and its orbit is lo- cated within the inner edge of the E-ring. Mimas has a diameter of 394.4 km, and has a density of 1.149gr/cm3 slightly higher than water. Mimas' surface is fully covered with impact craters and is one of the most heavily cratered bodies in the Solar System. This satellite has a large impact crater of 139 km in diameter (one third the diameter of Mimas) called Herschel. Schmedemann and Neukum (2011) also suggest evidence for two highly de- graded large craters, one of which is 153 km in diameter. After calculating the production of craters by Centaur objects on Mimas, DZ13 found that the largest crater produced by Centaurs has a diameter of 113 km, whose size is similar to Herschel crater. Lastly, KS10 used high resolution images from Cassini to record surface impact craters of the surface and they found a high density of craters with diameters ∼ 10 < D < 30 km. Plescia and Boyce (1982) obtained crater counts in different areas of Mimas from Voyager im- ages. They distinguished an equatorial area near Hershel (of low resolution), and south polar and adjacent areas which generally lack craters > 20 km. Buratti et al. (2011) have failed to find plume activity on the satellite and concluded that there could only be low-level geologic activity. They also limited this plume activity to an upper level of one order of magnitude less than the production for Enceladus. Measurements of Mimas's forced libra- tions by Tajeddine et al. (2014) confirms the libration amplitudes calculated 17 Mimas 4.5e+09 4e+09 ] r y [ e g A 3.5e+09 3e+09 2.5e+09 2e+09 PB-SP PB-T KS09 2 4 6 8 12 10 D [km] 14 16 18 20 Figure 5: Age of the south polar (PB-SP) and intermediate latitudes regions (PB-T) from Voyager and Cassini images (KS09) of Mimas with respect to D. from the orbital dynamics, with the exception of one amplitude that depends on Mima's internal structure. They suggested that a non-hydrostatic core or a subsurface ocean are the two possible interior models of Mimas that are consistent with their observations. It is evident that age can be calculated only if there exist measures of the number of observed craters. For example, in the case of Mimas, as there are no counts for craters with D . 3.5 km, the age-curve in Figs. 3 and 5 begins at this diameter. We calculate the age depending on D from Eq. 5, considering crater counts of the equatorial (E), south polar (SP) and intermediate latitudes regions (T) from Plescia and Boyce (1982), and also crater counts by KS09 from Cassini images of half of Mima' surface. We plot the results in Fig. 5. Crater counts for the equatorial zone in Voyager images are given only for large diameters due to low resolution, and the age is nearly 4.5 Gyrs old. The south polar terrain of Voyager images has a similar age curve to Cassini images for D . 12 km, but for larger craters, the south polar terrain seems to be younger, probably due to the lack of craters > 20 km found by 18 Plescia and Boyce (1982). Since on the one hand, Cassini counts include almost half of Mimas' surface and have a better resolution and on the other hand, SP area is small, it is possible that this drop of the SP age-curve is very local. For T-area and Cassini images, craters with D . 15 km are younger than 4.5 Gyrs old, and smaller craters might be even younger. For example, craters with D . 6 km are younger than ∼ 3.5 Grys old. In Fig. 2, there are ∼ 8000 craters per million km2 with D & 6 km that could have been erased from the surface in the age of the Solar System. Therefore, one possibility is that, effectively, small craters could have been erased due to particle depositions from the E-ring or even from a low-level geologic activity that should not be discarded on Mimas. However, since Mimas is heavily cratered, another possibility is that for D . 20 km, crater density has reached a saturation equilibrium. In fact, KS09 found that the power-law index of the observed cumulative crater SFD is −1.548 for 4 km< D < 10 km and −2.12, for 10 km < D < 20 km, both indexes near to the saturation equilibrium condition. The slope behavior of the age-curve is similar to Tethys and both curves are below the other two curves of the satellites plotted, which shows an erasure process or saturation that acts on a wider range of crater sizes on those satellites. Because of the recently suggested potential for activity on Mimas, we cannot discard an erasure process accounting for the paucity of small craters suggested in our model. However, the more conservative hypothesis would be saturation equilibrium for craters less than 20 km in diameter. 3.3. Tethys Tethys is cold and airless, 1072 km in diameter, and very similar in size and aspect to Dione and Rhea. As it has a density of 0.985gr/cm3 (Thomas, 2010), very similar to liquid water, it is probably almost entirely composed of water ice and a small amount of rock. In fact, spectral observations of Tethys' surface show no other component than H2O ice (Emery et al., 2005). This satellite has two main features on its surface: one of them is a great impact crater known as Odysseus, whose diameter is ∼ 450 km, and the other one is a great graben called Ithaca Chasma, whose length is ∼ 2000 km. The interior of Odysseus seems to be younger than the rest of Tethys' surface. In general, the surface of Tethys seems to be old and highly cratered, showing little geologic diversity. Royer and Hendrix (2014) present observations from the Ultraviolet Imag- ing Spectrograph subsystem of Cassini for Mimas, Tethys and Dione. They 19 found that Tethys and Dione have a leading hemisphere brighter than their trailing hemisphere at far-UV wavelengths, and Tethys shows a narrower op- position effect, reflecting a more porous surface. They attribute this structure to an intense bombardment by E-ring. There are crater counts on Tethys' areas from Voyager images (Plescia and Boyce, 1982 and 1983) and from Cassini images (KS09). From those observed crater counts and our model, we calculate the age depending on D from Eq. 5 for those terrains. The two bands, along the west side of the terminator observed by Voyager and studied by Plescia and Boyce (1982) have ages near 4.5 Gyrs, consistent with the high craterization observed. The other four zones analyzed by Plescia and Boyce (1983) have crater counts for D > 10 km. The heavily cratered terrain near 60◦ longitude is 4.5 Gyrs old. The remaining three areas that include two plain units (CT and P) and the interior of Ithaca Chasma graven (IC) have craters with D . 15 −20 km that are younger than 4.5 Gyrs old, which is consistent with the age calculated by Plescia and Boyce (1985). Crater counts on Cassini images performed by KS09 were done for ∼ 0.2 < D . 90 km and they have similar ages to plain units and IC for D & 10 km. Ages depending on D for CT, P, IC and KS09 terrains are plotted in Fig. 6 and the heavily cratered terrains have been omitted for clarity. In Fig. 3, we plot the age-curve of Tethys only for KS09 crater counts, since it has a wider range of diameters and a better resolution. It shows an age-curve similar to that of Mimas for the range of diameters plotted. Craters with D . 20 km could be young and, like the other satellites, smaller craters are younger than larger ones. KS09 and KS10 calculated the cumulative slopes of the SFD of craters and found that for 0.2 km < D < 10 km the cumulative slope is −1.728, and for 10 km < D < 60 km is −2.2. Then, for small craters those distributions could be close to the saturation equilibrium, especially for D . 10 km. Craters greater than D ∼ 20 km seem not to be affected by the erasing process or saturation, and thus the surface has preserved them, as can also be seen in Fig. 5 of DZ13. Therefore, those craters should be primordial. Like on Mimas, observations by Buratti et al. (2011) do not show plume activity, and the upper limit of them are also one order of magnitude less than the production on Enceladus. However, this does not rule out some possible activity. Since there appear to be small craters that would have been erased, there could, in fact, be some kind of geological activity (even plume activity at a low level, which has not been detected yet) that is acting on the satellite, 20 ] r y [ e g A 4.5e+09 4e+09 3.5e+09 3e+09 2.5e+09 2e+09 1.5e+09 1e+09 5e+08 0 Tethys KS09 PB-CT PB-P PB-IC 5 10 15 D [km] 20 25 30 Figure 6: Age of the Ithaca Chasma (PB-IC) interior terrain, plain units (PB-CT) and (PB-P) from Voyager images and Cassini images (KS09) of Tethys with respect to D. and/or also bombardment of the E-ring flux, as was suggested by Royer and Hendrix (2014). As can be seen in Fig. 2, this process should be able to erase, for example, ∼ 14000 craters per 106 km2 with D & 4 km or ∼ 106 craters per 106 km2 with D & 1 km in 4.5 Gyrs. However, as mentioned before, crater saturation in small crater sizes is the most conservative hypothesis for the paucity of observed counts. 3.4. Dione Dione has a diameter of 1122.8 km and a mean density of 1.478gr/cm3. Its surface is uneven, as it presents heavily cratered terrains with craters greater than 100 km in diameter, as well as smooth plains (Smith et al., 1982; Plescia and Boyce, 1982). It shows a marked asymmetry between its leading and trailing hemispheres, with the leading side brighter than the trailing one. Royera and Hendrix (2014) analyzed the photometric properties of Dione revealing a more absorbent surface. They ascribe this to the fact that Dione is less intensely bombarded by E-ring grains and thus it should 21 4.5e+09 4e+09 3.5e+09 3e+09 2.5e+09 2e+09 1.5e+09 1e+09 5e+08 ] r y [ e g A 0 0 5 10 Dione KS09-cp KS09-sp PB-IC PB-LC 15 D [km] 20 25 30 Figure 7: Age of an intermediate crater plains unit (PB-IC) and a smooth plains unit (PB- LC) from Voyager images and cratered plains (cp) and smooth plains (sp) from Cassini images (KS09) of Dione with respect to D. have less fresh bright water-ice on its surface; and/or to that an exogenous darkening agent could be acting simultaneously on the trailing side. In addition, Buratti et al. (2011) analyzed spectral observations of Dione in search for plume activity and found that the upper limit of water vapor production is two orders of magnitude smaller than that of Enceladus, sug- gesting that this world is inert. However, plume activity cannot be absolutely ruled out because it can be under the detection limit. Further analysis by the Cassini magnetometer suggests that Dione has a tenuous atmosphere of transient origin (Simon et al., 2011). Future observations by Cassini could possibly confirm the existence of an atmosphere and/or some plume activity. Plescia and Boyce (1982) analyzed three terrain types observed by Voy- ager: a heavily cratered rough terrain (HC), an intermediate crater plain unit (IC) and a smooth plain unit (LC). They obtained cumulative size frequency distributions of craters for those terrains, which we used in our model to calculate surface ages. The HC area has crater counts for ∼ 10 < D . 100 22 km and is 4.5 Gyrs old. Ages depending on D of IC and LC plain units are shown in Fig. 7. The IC unit has crater counts for 6 . D . 20 km and it seems that craters between 15 and 20 km are slightly younger than 4.5 Gyrs old. The LC terrain would be younger than 4.5 Gyrs old for all the diameters observed. This is consistent with the low crater density of LC and IC plains with respect to HC terrain found by Plescia and Boyce (1982) and also with the younger ages found by Plescia and Boyce (1985) for those IC and LC plains. On Dione there are two different surfaces recognized by KS10 from Cassini observations. The cratered plains (cp), which are heavily cratered, presenting records of primordial bombardment, and the smooth plains (sp) with less cratering, suggesting a younger surface. In Figs. 7 and 3, we plot the age with respect to D for both terrains on Dione. In these figures we can see that the observed cp surface portion of Dione is generally old, but craters with D . 5 km could be young, and smaller craters would be younger than larger ones. This could be the result of a process in which craters have been gradually erased. This possible erasing process does not appear to affect craters greater than ∼ 5 km. Although there is no data for D . 4 km for sp, the smooth terrain has a lower amount of craters than the cp surface, and the corresponding age curve is below the cp-age curve. For example, for craters with D & 4 km our model suggests that sp has an age of 3.8 Gyrs old while the cp is 4 Gyrs old. Indeed, since the age of sp for D & 8 km is near 4.5 Gyrs old (Fig. 2) and the number of calculated craters is greater than the observed one, this suggests a very slow erasing process for D & 5 km, which is not present in the case of cp. Potential erasing process/es on Dione would need to be able to erase in sp terrain, ∼ 5500 craters per 106km2 of D & 4, while in cp terrain there would be ∼ 4500 craters per 106 km2 of D & 4 erased in 4.5 Gyrs. Comparing age estimations between Voyager and Cassini plains, it is evident that LC plains would be younger than smooth plains, at least in the range of diameter observed by both missions. It could be that LP plains were in fact young due to, for example, the internal activity which resurfaced portions of the surface of Dione, as was suggested by Plescia and Boyce (1985). However, Cassini images have a resolution of 1 to 2 orders of magnitude higher than Voyager images, so the comparison between age calculation with the data of both missions has to be taken with caution. On the other hand, like on the other satellites, KS10 calculated the cu- mulative crater SFD for cratered plains and obtained a cumulative slope of 23 b = −1.640, for 0.25km < D < 4 km, b = −1.166 for 4 km < D < 10 km, b = −2.31 for 10 km < D < 30 km, and −2.9 for 30km < D < 150 km. Therefore, the crater SFD of cp is similar to our model for craters greater than 30 km, but for smaller sizes the slope of the SFD is getting more flat- tened, similar to what occurs for Iapetus, indicating saturation equilibrium for D . 10 − 30 km. For the smooth plains a similar behavior of the crater SFD has been found. Plescia and Boyce (1982) found a cumulative slope of b = −2 for the heavily cratered terrain, indicating saturation equilibrium, and a cumulative slope of b = −4 for the younger units, more similar to our model. Then, crater saturation could also be responsible for the paucity of small craters on Dione although there is morphological evidence that small craters in the sp have also been affected by endogenous resurfacing (Schenk and Moore, 2009). 3.5. Rhea Rhea is the second largest moon of Saturn, with a diameter of 1529 km and a density of 1.237gr/cm3 (Thomas 2010), a value somewhat greater than that of liquid water. Therefore, it is thought to be composed of a homogeneous mixture of ice and rock. Its surface of Rhea is densely cratered with abundant large heavily degraded craters. Rhea seems to be more heavily cratered than Dione and Tethys. An outstanding feature of this moon is a fresh ray crater of ∼ 47.2 km of diameter, known as Inktomi. This crater is probably the youngest feature on Rhea's surface, whose estimated age ranges from 8 to 280 Myrs old (Wagner et al., 2008). Measurements by Cassini spacecraft have detected a tenuous atmosphere of oxygen and carbon dioxide, which appears to be sustained by chemical decomposition of the surface water ice under irradiation from Saturn's magnetospheric plasma (Teolis et al., 2010). Stephan et al. (2012) analyzed the spectral characteristics of Rhea and concluded that the major process that affects the surface properties of Rhea is the interaction between the surface material and the space environment which includes the impacts of energetic particles from the magnetospheric plasma. This exogenous process could be responsible for the erosion of small craters on the surface of Rhea. Plescia and Boyce (1982) analyzed different areas of Rhea's surface and found significantly different crater density. From our model, we found an age of 4.5 Gyrs old for the north polar region, this zone being densely cratered. Equatorial regions show subdued surface texture and absence of small craters, 24 4.5e+09 4e+09 3.5e+09 3e+09 2.5e+09 2e+09 1.5e+09 1e+09 5e+08 ] r y [ e g A 0 0 5 10 Rhea KS09 PB-M PB-UM 15 D [km] 20 25 30 Figure 8: Age of equatorial mantled (PB-M) and unmantled regions (PB-UM) from Voyager images and Cassini images (KS09) of Rhea with respect to D. suggesting that some areas have been mantled. Thereby, Plescia and Boyce (1982) separated them into mantled (M) and unmantled regions (UM). KS09 obtained crater counts from Cassini images with a resolution of 0.01 − 1 km/pixel for ∼ 25% of Rhea's surface, allowing them to reach very small diameters. We calculated ages for this Rhea's surface and plotted them together with the equatorial regions analyzed by Plescia and Boyce (1982) in Fig. 8. We can see that the mantled regions are younger than the other terrains, that is, ∼ 3.8 Gyrs old for D & 10 km, consistent with the age estimations by Plescia and Boyce (1985). However, although it is possible that this zone could be younger for smaller sizes, it cannot be guaranteed since there are no counts for craters smaller than 8 km in diameter for these zones. Then, in order to compare Rhea's surface age with the other satellites, we plot our results for KS09 counts in Fig. 3. Rhea has the oldest surface with respect to the other satellites, with the exception of Iapetus (Fig. 3). However, craters with D . 4.5 km could be young, and the slope of the age curve is steep, which causes age to strongly 25 decrease in direct proportion to crater size. Thus, the possible process that might have erased small craters has a strong dependence on size. This agrees with the above mentioned process that gradually erodes the surface of Rhea, thus erasing small craters without burying the greatest ones. In Fig. 2, for example, there are ∼ 40000 per 106 km2 craters with D & 2 that are erased from the surface of Rhea in the age of the Solar System. A peak near ∼ 0.5−km craters can be seen in the age curve. For D . 1 km, the number of observed craters on Rhea seems to have increased from the theoretical slope (see Fig. 7 in DZ13). Therefore, it is expected that our calculated age curve will change its behavior. As KS09 noted, for their observed impact crater distributions on the Saturnian satellites, D < 1−km craters are likely to be contaminated with secondary craters at some level, thus this "peak" in the age curve is probably due to this contamination on Rhea in the observed crater counts for small diameters. In relation to the observed cumulative crater SFD, KS09 and KS10 ob- tained cumulative slopes near −2 for all the diameter ranges. Then, although our model with s2 = 3.5, that produces a crater SFD with a power-law index of ∼ −3, fits better the number of observed craters, the power-law index of the distribution does not. Therefore, there exists the possibility that the surface of Rhea, is also saturated for the crater range observed. This would be consistent with the high density of craters observed and with the results of Squyres et al. (1997), who found that at least 25% of the craters on Rhea were destroyed by subsequent obliteration. 3.6. Comparison with other age estimations As mentioned before, KS09 obtained the number of craters for the mid- sized icy satellites, and from previous cratering rate estimations by Zahnle et al. (2003), they calculated surface ages for each satellite and for some crater diameters too. Zahnle et al. (2003) analyzed two cases of a population of impactors. In case A, they inverted crater counts on the Galilean satellites to obtain the size distribution of impactors on Jupiter. In case B, they inverted crater counts on Triton from Voyager images to obtain the size distribution of impactors. In Table 3, we show both our age estimations for Mimas, Tethys, Dione and Rhea, for some diameters, and the estimations by KS09. The tendency of an increasing age with diameter that we have found is only shown by the results of case B (in general). Our results lie, in general, between case A and B, but for small craters they are more similar to case B. This could be due to the fact that the production of craters in the Neptune 26 Table 3: Ages obtained from our results and those calculated by KS09 all in Gyrs for various diameter ranges. Satellite D [km] Age KS (A) KS (B) Mimas Tethys Dione-cp Dione-sp Rhea 5 10 20 1 2 5 10 20 1 2 5 10 20 5 10 20 1 2 5 10 20 3.1 4.2 4.5 1.1 1.6 3.5 4.2 4.4 1.2 2.5 4.4 4.5 4.5 4.2 4.5 4.5 1.6 3.4 4.5 4.5 4.5 4.4 4.4 3.3 4.6 4.6 4.6 4.5 4.3 4.6 4.6 4.6 4.6 4.5 4.6 4.4 3.3 4.6 4.6 4.6 4.6 4.5 0.8 1.3 0.9 1.6 1.2 1.7 2.1 2.7 0.9 1.4 2.6 3.3 3.1 2.0 2.0 1.4 1.7 2.2 3.1 3.7 3.9 System could be more directly related to SDOs just entering the Centaur zone, which is the case we are working on. However, the estimations by Zahnle et al. (2003) have an uncertainty factor of 4. According to our results, all surfaces analyzed up to now appear be old. However, we have noticed that for small sizes the observed number of craters is smaller than the calculated number from our model. The size limit of each satellite is presented in Table 1. Although this size limit is different for each satellite, a common explanation for this paucity of small craters could be crater saturation. In particular, crater saturation starts in small sizes and would gradually incorporate larger crater sizes. 27 3.7. Enceladus Enceladus is a very interesting and beautiful moon of Saturn. It has a diameter of 504 km and a density of 1.609gr/cm3 (Thomas, 2010). Cassini- Huygens images reveal a very active surface with unique features, which invites one to investigate them. The discovery of the plumes of water vapor and small icy particles in the south polar region by Cassini Huygens in 2005 (Porco et al., 2006) showed us a very active world and at the same time, it was very important for the study of icy satellites. High resolution images of the south polar terrains from Cassini reveal tectonic features, a region almost entirely free of impact craters but with several ∼ 130-km-long fractures called "tiger stripes", probably warmed by internal tidally generated heat (Porco et al., 2006). It was also found that this south polar region is 20 K hotter than expected from models, and a thermal emission of 3 to 7 GW was detected (Spencer et al. 2006). Gravity field studies by Iess et al. (2014) demonstrated that the structure of Enceladus is compatible with a differentiated body with, for example, a relatively low core density of ∼ 2.4gr/cm3 and an H2O mantle of density of 1gr/cm3 and thickness of 60 km. However, the local south polar heat fluxes, gravity and topography, is more consistent with a regional sub sea of ∼ 10 km thick in the south pole located beneath an ice crust ∼ 30 − 40 km thick (Iess et al., 2014). The surface of Enceladus is covered in almost pure water ice (Brown et al., 2006). Newman et al. (2008) mapped the distribution of crystalline and amorphous ices on the surface of Enceladus from photometric and spectral analysis of data from Cassini Visual and Infrared Mapping Spectrometer. They found a mostly crystalline ice surface with a higher degree of crystallinity in the "tiger-stripe"cracks, and amorphous ice between these stripes. "Ice blocks" were also observed on Enceladus' surface, like the regolites in asteroids and airless bodies of the Solar System (Porco et al., 2006). However, their origin seems to be different. Martens et al. (2015) analyzed their distribution and origin in detail and found that, although they are found in other regions, they are more concentrated within the south polar terrain, where they must be produced mainly by tectonic deformation. All this suggests a very active and evolving surface. Moreover, Enceladus displays different types of terrain: in particular, as one moves away from the "tiger stripes", more craters are found (KS09). This is due to the continuous resurfacing in the south polar zone. In fact, Porco et al. (2006) estimated ages within the south polar terrain possibly as young as 500000 years, or even younger. 28 (1982). As in the previous section, we estimate the age of Enceladus' surface comparing our calculations of the cratering production with the observed crater counts. There are crater counts on Enceladus from Cassini by KS09, but also previous counts from Voyager by Plescia and Boyce (1983) and Smith et al. Images from Voyager spacecrafts 1 and 2 showed a wide diversity of terrains on the surface of Enceladus. Different areas have been identified, according to the populations of craters and other landforms (Fig. 21 of Smith et al., 1982). Among these, the cratered terrains (ct1) and well preserved terrains (ct2) are distinguished. Another kind of terrain are the cratered plains (cp), whose principal landforms are bowl-shaped craters. In the equatorial region Voyager observed smooth plains sp1 and sp2. The grooves are the principal landforms in sp1, which have few craters of 2 km to 5 km in diameter, while sp2 and ridged plains (rp) are almost free of craters. The number of observed craters in Voyager images were taken from Fig. 1 and Table 1 in Plescia and Boyce (1983) for ct1 and ct2 (ECT) and cp (ECP), and from Fig. 23 in Smith et al. (1982) for sp1, sp2 and rp. Kirchoff and Schenk (2009) analyzed Cassini images for crater counts. They analyzed different types of terrain separately according to variations in crater density and geological features (see their Fig. 4). Those regions are cratered plains (cp-eq and cp-mid) and the ridged regions rp1, rp2, rp3, rp4, rp5, and rp6. The regions analyzed by Voyager and Cassini seem to be different. How- ever, through an inspection of cartographic base maps of Enceladus in both papers (Smith et al., 1982 and KS09), it seems that rp and sp2 and probably sp1 regions of Voyager are near the ridged plains (rp1-rp6) of KS09; ct1 and ct2 regions of Voyager are near mid-latitude cratered plains of KS09, and cp regions of Voyager are near equatorial cratered plains of KS09. It is worth noting, however, that the Cassini data have a higher spatial resolution than that from Voyager. The Cassini mission has been able to show craters with D < 2 km. Also, KS09 has made a subdivision of ridged plains (rp), resulting in better details. From our model developed in DZ13, we have got the cumulative number of craters produced by current Centaurs. In order to visualize all the observed crater counts and to compare them with our model, those numbers are shown in Fig. 9. In general, the observed curves are below the calculated curve in our model. Therefore, the observed number of craters on the studied terrains would be lower than the current expected contribution of Centaurs with the 29 2 m k 6 0 1 r e p ) D > ( c N 1e+08 1e+07 1e+06 100000 10000 1000 100 10 Enceladus model rp1 rp2 rp3 rp4 rp5 rp6 cp-eq cp-mid ECP ECT sp1 sp2-rp 1 10 D [km] Figure 9: Cumulative number of craters calculated from model by DZ13 (s2 = 3.5) and crater counts from Voyager (ECP, ECT, sp1, sp2-rp) and from Cassini by KS09. 30 ] r y [ e g A 4.5e+09 4e+09 3.5e+09 3e+09 2.5e+09 2e+09 1.5e+09 1e+09 5e+08 0 0 Enceladus cp-eq cp-mid rp6 ECP ECT 5 10 15 20 25 30 35 40 45 50 D [km] Figure 10: Age of the cp and rp6 surfaces of Enceladus with respect to D. exception of ECP and ECT, which present a greater number of observed craters than our model for D & 20 − 30 km respectively. We have calculated the age with respect to D from Eq. 5 and plotted the results in Figs. 10 and 11 for both observed crater counts (Voyager and Cassini) and regions. We have separated the results into two figures because of the wide age range. We can see that the observed areas of Enceladus are usually young, and that age increases as we move away from the "tiger stripe" cracks. Cratered plains in mid and high latitudes (cp-mid) are ∼ 4.3 Gyrs old, regions ct1 and ct2 of Voyager are ∼ 4.5 Gyrs old, but craters less than ∼ 6 km could be younger. On the contrary, cratered plains in equatorial latitudes (cp-eq) are younger than 3 Gyrs old for the crater sizes counted by KS09. ECP from Voyager seems to be old, but craters less than ∼ 20 km could be younger. All rp regions appear to be very young. The rp1 region, which is very near the "tiger stripes", a very active region, is the youngest region, and rp6 31 Enceladus rp1 rp2 rp3 rp4 rp5 rp6 sp1 sp2-rp 1.4e+09 1.2e+09 1e+09 8e+08 6e+08 4e+08 2e+08 ] r y [ e g A 0 1 2 3 4 5 D [km] 6 7 8 Figure 11: Age of the rp surfaces of Enceladus with respect to D. 32 is the oldest one. However, all sizes of craters in rp6 are young. In fact, there are craters with D . 16 km which are younger than ∼ 3.5 Gyrs old. In rp6, age increases with diameter like in the other mid-sized satellites. But in the rp1-rp5 zones, ages are similar for all the diameter ranges analyzed and they are all less than ∼ 500 Myrs old. The very young region rp1 is ∼ 70 Myrs old on average. Those ages of the ridged plains would be in agreement with estimations by Plescia and Boyce (1985) who estimated an age of no more than 100 Myrs. This behavior in the rp1-rp5 zones must be due to the plumes emanating from the south polar region and tectonic activity probably associated with "tiger stripes", whose production mechanisms strongly erase those craters. In comparison with previous research, our ages are more similar to those calculated by KS09 for their case B (see Table 3). Moreover, Enceladus' plumes are thought to be the source and mainte- nance of E-ring material, but a fraction of the plume particles returns to Enceladus hitting its surface (Kempf et al., 2010). The particle deposition on Enceladus surface is more frequent in the "tiger stripes" than in other regions, as for example mid latitudes (Kempf et al., 2010). Also, the largest water ice particles covering Enceladus' surface (∼ 0.2 mm) are concentrated in the "tiger stripes" zone and in general, the particle diameters are strongly correlated with the distance to this zone (Jaumann et al., 2008). Therefore, the plume material returning to Enceladus could be an "exogenous" source causing the erasure of craters preferentially near the tiger stripes. On the other hand, the heat emanating from the south polar terrain, detected by Cassini's Composite Infrared Spectrometer, suggests internal ge- ological activity (Spencer et al., 2006). The heat source for the production of the plumes, resurfacing and heat flow in the south polar region of Enceladus, has not yet been sufficiently understood. The estimated radiogenic and tidal heating resulting from the orbital eccentricity of Enceladus are not sufficient to account for the energy emanating from the south polar zone (Porco et al., 2006; Meyer and Wisdom, 2007). Besides, the absence of internal activity on Mimas, which has an orbital eccentricity greater than Enceladus', further constrain this as an unlikely source. Porco et al. (2006) analyzed other tidal heating mechanisms, such as mean motion resonances among the Saturnian moons, the 2:1 mean motion resonance with Dione being the most important, and secondary resonance heating, but those mechanisms are also inadequate. Squyres et al. (1983) suggested that the current rate of tidal heating might be sufficient to maintain a liquid sea under a crust of 10 km, but a previous 33 stronger heating would be required to initiate the melting. This previous process might have been that Enceladus had a larger eccentricity in the past (to obtain a greater tidal heating) or was captured into mean motion or secondary resonances with other satellites (Porco et al., 2006; Meyer and Wisdom, 2008). Meyer and Wisdom (2008) demonstrated that Enceladus is probably near its equilibrium eccentricity and that the current equilibrium heating rate is not enough to trigger the melting. Therefore, any previous stronger heating requires some non-equilibrium behavior. 4. Discussion and conclusions Based on both calculations of the cratering rate and the number of craters produced by Centaurs, and the comparison with observations, we have es- timated the surface age of each observed terrain on each mid-sized satellite of Saturn. The results are given for each satellite, but in general terms we can say that there are less observed small craters than calculated (except on Iapetus), which result in young ages calculated from our model. This could be interpreted as either efficient endogenous or exogenous process(es) erasing small craters or crater saturation. We have also found that the smaller craters are being preferentially re- moved. This agrees with a gradual process of erosion, either from a geological or exogenous origin or by cratering itself. This erasing process does not affect large craters, and the size limit depends on the satellite. For craters greater than this size limit, our results imply that there are primordial craters, and/or that another main source of craters might exist. We have also found a correlation between the limit size from which small craters are erased (and then age is below 4.5 Gyrs old) and the distance to Saturn. That is, the smaller the distance from Saturn, the higher the size limit (see Table 1). It is possible that there is not a general process to de- scribe this behavior since each satellite has its own peculiarities. However, a potential common process that seems to explain the paucity of small craters is crater saturation which could be connected to the correlation between dis- tances to Saturn and erasure size limit. Another possibility is that deposition of E-ring particles could be responsible for such a correlation. Since the E- ring has a density that generally decreases from Enceladus to Rhea (Hor´anyi et al., 2008), the E-ring flux on the mid-sized satellites could be lower as we move away from Enceladus. 34 The estimated ages for the terrains analyzed on the mid-sized satellites, with the exception of Enceladus, are ∼ 4.5 Gyrs old. Our ages are, within the ages calculated by KS09, but the fact that smaller craters are preferentially removed seems to be strong in our calculations. Almost all the Enceladus' terrains could be young, with the exception of the mid and high latitude plains. The regions near the south polar terrain could be as young as 50 Myrs old. On the contrary, all the surface of Iapetus is old and it likely records a primordial source of craters, smaller craters being likely approach- ing saturation until D . 2 km-craters, where saturation is complete. The ages calculated here are based on the current number of Centaur objects as contributors to craters. We have used the best known estimations from the literature for both this number and the size-frequency distribution. However, for small sizes, this SFD is uncertain and also unreachable with current direct observations. That is why the theoretical crater contribution, the comparison with observations and the determination of terrain ages in this paper should be carefully interpreted and not taken as absolute ages. They could serve, instead, to enhance both the understanding of the pro- cesses occurring on the satellites and the SFD for small sizes. Finally, the comparison of our age estimations with geological time scales on Saturnian satellites could help us determine what is still unknown today about those wonderful worlds. Acknowledgments: We would like to thank Gabinete de Ingl´es de la Facul- tad de Ciencias Astron´omicas y Geof´ısicas de la UNLP for a careful language revision. We are grateful to Eduardo Fern´andez Laj´us for valuable discussion suggestions and to Michelle Kirchoff and an anonymous referee for valuable comments and suggestions that helped us improve the manuscript. References [1] Brown, R.H., Clark, R.N., Buratti, B.J., et al. 2006. Composition and Physical Properties of Enceladus' Surface Science, 311, 1425. [2] Buratti B.J., Faulk, S.P., Mosher, J., et al. 2011. Search for and limits on plume activity on Mimas, Tethys, and Dione with the Cassini Visual Infrared Mapping Spectrometer (VIMS). Icarus, 214 534-540. [3] Buratti, B.J., Mosher, J.A. 1995. The dark side of Iapetus: Additional evidence for an exogenous origin. Icarus, 115, 219-227. 35 [4] Cruikshank, D.P., Bell, J.F., Gaffey, M.J., et al 1983. The dark side of Iapetus. Icarus, 53, 90-104. [5] Denk, T., Neukum, G., Roatsch, T., et al. 2010. Iapetus: Unique Surface Properties and a Global Color Dichotomy from Cassini Imaging. Science, 327, 435. [6] Di Sisto, R.P. and Brunini, A. 2007. The origin and distribution of the Centaur population. Icarus, 190, 224-235. [7] Di Sisto, R.P., Fern´andez J. A. and Brunini, A. 2009. On the popula- tion, physical decay and orbital distribution of Jupiter family comets: Numerical simulations. Icarus, 203, 140-154. [8] Di Sisto, R.P. and Brunini, A. 2011. Origin of craters on Phoebe: com- parison with Cassini's data. A&A, 534, A68, 6 pp. [9] Di Sisto, R.P. and Zanardi, M. 2013. The production of craters on the mid-sized Saturnian satellites by Centaur objects. A&A, 553, A79, 9 pp. [10] Dobrovolskis, A.R., Lissauer, J.J. 2005. The fate of ejecta from Hyper- ion. Icarus, 169, 462-473. [11] Dones, L., Chapman, C.R., McKinnon, W.B., et al. 2009. Icy Satellites of Saturn: Impact Cratering and Age Determination. In Saturn from Cassini-Huygens, M.K. Dougherty et al. (eds.) 613-635. [12] Emery, J.P., Burr, D.M., Cruikshank, D.P., et al. 2005. Near-infrared (0.8-4.0 µm) spectroscopy of Mimas, Enceladus, Tethys, and Rhea A&A, 435, 353-362. [13] Elliot, J.L., Kern, S.D., Clancy, K.B., et al. 2005. The Deep Ecliptic Survey: A Search for Kuiper Belt Objects and Centaurs. II. Dynamical Classification, the Kuiper Belt Plane, and the Core Population. AJ, 129, 1117-1162. [14] Gault, D.E. 1970. Saturation and equilibrium conditions for impact cra- tering on the lunar surface: Criteria and implications. Radio Science, 5, 273-291. [15] Hartmann, W. K. 1984. Does crater 'saturation equilibrium' occur in the solar system?. Icarus, 60, 56-74. 36 [16] Hamilton, D.P., Burns, J.A. 1994. Origin of Saturn's E Ring: Self- Sustained, Naturally. Science, 264, 550-553. [17] Hartmann, W.K., Gaskell, R.W. 1997. Planetary cratering 2: Studies of saturation equilibrium. Meteoritics, 32, 109-121. [18] Holman, M.J. and Wisdom, J. 1993. Dynamical stability in the outer Solar System and the delivery of short-period comets. AJ, 105, 19871999. [19] Hor´anyi, M., Juh´asz, A., Morfill, G. 2008. E Large-scale structure of Saturns E-ring. Geophysical Research Letters, 35, L04203. [20] Horedt, G.P., Neukum, G. 1984. Planetocentric versus heliocentric im- pacts in the Jovian and Saturnian satellite system. Journal of Geophys- ical Research, 89, 10405-10410. [21] Iess, L., Stevenson, D.J., Parisi, M. et al. 2014. The Gravity Field and Interior Structure of Enceladus. Science, 344, 78-80. [22] Ingersoll, A.P., Ewald, S.P. 2011. Total particulate mass in Enceladus plumes and mass of Saturn?s E ring inferred from Cassini ISS images. Icarus, 216, 492-506. [23] Ip, W.H. 2006. On a ring origin of the equatorial ridge of Iapetus. Geo- physical Research Letters, 33, CiteID L16203. [24] Jacobson, R.A., Antreasian, P.G., Bordi, J.J. et al. 2006. The Gravity Field of the Saturnian System from Satellite Observations and Space- craft Tracking Data. AJ, 132, 2520-2526. [25] Kempf, S., Beckmann, U., Schmidt, J. 2010. How the Enceladus dust plume feeds Saturn's E ring. Icarus, 206, 446-457. [26] Kirchoff, M.R., Schenk, P. 2009. Crater modification and geologic ac- tivity in Enceladus' heavily cratered plains: Evidence from the impact crater distribution. Icarus 202, 656-668. [27] Kirchoff, M.R. and Schenk, P. 2010. Impact cratering records of the mid-sized, icy saturnian satellites. Icarus 206, 485-497. 37 [28] Levison, H. F., Morbidelli, A., VanLaerhoven, C., Gomes, R., Tsiganis, K. 2008. Origin of the structure of the Kuiper belt during a dynamical instability in the orbits of Uranus and Neptune. Icarus, 196, 258-273. [29] Martens, H.R., Ingersoll, A.P., Ewald, S.P., Helfenstein, P., Giese, B. 2015. Spatial distribution of ice blocks on Enceladus and implications for their origin and emplacement. Icarus, 245, 162-176. [30] Meyer, J. and Wisdom, J. 2007. Tidal heating in Enceladus. Icarus, 188, 535-539. [31] Meyer, J. and Wisdom, J. 2008. Tidal evolution of Mimas, Enceladus and Dione. Icarus, 193, 213-223. [32] Morbidelli, A., Levison, H.F., and Gomes, R. 2008, in The solar sys- tem Beyond Neptune, ed. Barucci, M.A. et al. (Tucson, USA: Univ. of Arizona Press), 275. [33] Nesvorn´y, D., Alvarellos, J.L.A., Dones, L., et al. 2003. Orbital and Collisional Evolution of the Irregular Satellites. AJ, 126, 398-429. [34] Nesvorny, D., Vokrouhlicky, D., Morbidelli, A. 2007. Capture Of Irreg- ular Satellites During Planetary Encounters. American Astronomical Society, Bulletin of the American Astronomical Society, 39, 47. [35] Newman, S.F., Buratti, B.J., Brown, R.H. et al. 2008. Photometric and spectral analysis of the distribution of crystalline and amorphous ices on Enceladus as seen by Cassini. Icarus 193, 397-406. [36] Plescia, J.B. and Boyce, J.M. 1982. Crater densities and geological his- tories of Rhea, Dione, Mimas and Tethys. Nature, 295, 285-290. [37] Plescia, J.B. and Boyce, J.M. 1983. Crater numbers and geological histo- ries of Iapetus, Enceladus, Tethys and Hyperion. Nature, 301, 666-670. [38] Plescia, J.B. and Boyce, J.M. 1985. Impact Cratering History of the Saturnian Satellites. Journal of Geophysical Research, 90, 2029-2037. [39] Porco, C.C., Baker, E., Barbara, J., et al 2005. Cassini Imaging Science: Initial Results on Phoebe and Iapetus. Science, 307, 1237. 38 [40] Porco, C.C., Helfenstein, P., Thomas, P.C. et al. 2006. Cassini Observes the Active South Pole of Enceladus. Science 311, 1393. [41] Richardson, J.E. 2009. Cratering saturation and equilibrium: A new model looks at an old problem. Icarus, 204, 697-715. [42] Royer, E.M. and Hendrix, A.R., 2014. First far-ultraviolet disk- integrated phase curve analysis of Mimas, Tethys and Dione from the Cassini-UVIS data sets. Icarus 242, 158 - 171. [43] Schmedemann, N., Neukum, G. 2011. Impact Crater Size-Frequency Distribution (SFD) and Surface Ages on Mimas. 42nd LPSC, No. 1608, p.2772. [44] Schenk, P.M., and J.M. Moore. 2009. Eruptive Volcanism on Saturns Icy Moon Dione. LPSC, Abst. 2465. [45] Simon, S., Saur, J., Neubauer, F.M., Wennmacher, A., Dougherty, M.K. 2011. Magnetic signatures of a tenuous atmosphere at Dione. Geophys- ical Research Letters, 38 L15102. [46] Smith, B.A., Soderblom, L., Beebe, R.F. 1981. Encounter with Saturn - Voyager 1 imaging science results. Science, 212, 163-191. [47] Smith, B.A., Soderblom, L., Batson, R.M. 1982. A new look at the Saturn system - The Voyager 2 images. Science, 215, 504-537. [48] Spencer, J.R., Pearl, J.C., Segura, M., et al. 2006. Cassini Encounters Enceladus: Background and the Discovery of a South Polar Hot Spot. Science 311, 1401-1405. [49] Spencer, J.R., Denk, T. 2010. Formation of Iapetus' Extreme Albedo Dichotomy by Exogenically Triggered Thermal Ice Migration. Science, 327, 432. [50] Stephan, K., Jaumann, R., Wagner, R. et al. 2012. The Saturnian satel- lite Rhea as seenby Cassini VIMS. Planetary and Space Science, 61, 142 - 160. [51] Squyres, S.W., Reynolds, R.T., Cassen, P.M., Peale, S.J. 1983. The evolution of Enceladus. Icarus, 53, 319-331. 39 [52] Squyres, S.W., Howell, C., Liu, M.C., et al. 1997. Investigation of Crater "Saturation" Using Spatial Statistics. Icarus, 125, 67-82. [53] Tajeddine, R. Rambaux, N., Lainey, V. et al. 2014. Constraints on Mima's interior from Cassini ISS libration measurements. Science, 346, 322-324. [54] Teolis, B.D., Jones, G.H., Miles, P.F. et al. 2010. Cassini Finds an Oxygen-Carbon Dioxide Atmosphere at Saturns Icy Moon R-hea. Sci- ence 330, 181. [55] Thomas, P.C. 2010. Sizes, shapes, and derived properties of the satur- nian satellites after the Cassini nominal mission. Icarus 208, 395-401. [56] Tsiganis, K., Gomes, R., Morbidelli, A. and Levison, H.F. 2005. Ori- gin of the orbital architecture of the giant planets of the Solar System. Nature, 435, 459-461. [57] Wagner, R. J.; Neukum, G.; Giese, B.; et al. 2008. Geology of Saturn's Satellite Rhea on the Basis of the High-Resolution Images from the Targeted Flyby 049 on Aug. 30, 2007. In Lunar and Planetary Science XXXIX. 1391, p.1930. [58] Zahnle, K., Dones, L. and Levison, H.F., 1998. Cratering Rates on the Galilean Satellites. Icarus, 136, 202-222. [59] Zahnle, K., Schenk, P. Levison, H. and Dones, L., 2003. Cratering rates in the outer Solar System. Icarus, 163, 263-289. 40
1210.7066
1
1210
2012-10-26T08:26:05
The thermal reactivity of HCN and NH3 in interstellar ice analogues
[ "astro-ph.EP", "astro-ph.GA" ]
HCN is a molecule central to interstellar chemistry, since it is the simplest molecule containing a carbon-nitrogen bond and its solid state chemistry is rich. The aim of this work was to study the NH3 + HCN -> NH4+CN- thermal reaction in interstellar ice analogues. Laboratory experiments based on Fourier transform infrared spectroscopy and mass spectrometry were performed to characterise the NH4+CN- reaction product and its formation kinetics. This reaction is purely thermal and can occur at low temperatures in interstellar ices without requiring non-thermal processing by photons, electrons or cosmic rays. The reaction rate constant has a temperature dependence of k(T) = 0.016+0.010-0.006 s-1.exp((-2.7+-0.4 kJmol-1)/(RT)) when NH3 is much more abundant than HCN. When both reactants are diluted in water ice, the reaction is slowed down. We have estimated the CN- ion band strength to be A_CN- = 1.8+-1.5 x10-17 cm molec-1 at both 20 K and 140 K. NH4+CN- exhibits zeroth-order multilayer desorption kinetics with a rate of k_des(T) = 10^28 molecules cm-2 s-1.exp((-38.0+-1.4 kJmol-1)/(RT)). The NH3 + HCN -> NH4+CN- thermal reaction is of primary importance because (i) it decreases the amount of HCN available to be hydrogenated into CH2NH, (ii) the NH4+ and CN- ions react with species such as H2CO, or CH2NH to form complex molecules, and (iii) NH4+CN- is a reservoir of NH3 and HCN, which can be made available to a high temperature chemistry.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1–12 (2002) Printed 4 June 2018 (MN LATEX style file v2.2) The thermal reactivity of HCN and NH3 in interstellar ice analogues. J. A. Noble,⋆ P. Theule, F. Borget, G. Danger, M. Chomat, F. Duvernay, F. Mispelaer, and T. Chiavassa Aix-Marseille Universit´e, PIIM UMR-CNRS 7345, 13397, Marseille, France. Accepted 1988 December 15. Received 1988 December 14; in original form 1988 October 11 ABSTRACT HCN is a molecule central to interstellar chemistry, since it is the simplest molecule contain- ing a carbon-nitrogen bond and its solid state chemistry is rich. The aim of this work was to study the NH3 + HCN → NH+ 4 CN− thermal reaction in interstellar ice analogues. Laboratory experiments based on Fourier transform infrared spectroscopy and mass spectrometry were performed to characterise the NH+ 4 CN− reaction product and its formation kinetics. This reac- tion is purely thermal and can occur at low temperatures in interstellar ices without requiring non-thermal processing by photons, electrons or cosmic rays. The reaction rate constant has a temperature dependence of k(T ) = 0.016+0.010 ) when NH3 is much more abundant than HCN. When both reactants are diluted in water ice, the reaction is slowed down. We have estimated the CN− ion band strength to be ACN− = 1.8±1.5×10−17 cm molec−1 at both 20 K and 140 K. NH+ 4 CN− exhibits zeroth-order multilayer desorption kinetics with a rate of kdes(T ) = 1028 molecules cm−2 s−1 exp( −38.0±1.4 kJ mol−1 4 CN− thermal reaction is of primary importance because (i) it decreases the amount of HCN avail- able to be hydrogenated into CH2NH, (ii) the NH+ 4 and CN− ions react with species such as H2CO, or CH2NH to form complex molecules, and (iii) NH+ 4 CN− is a reservoir of NH3 and HCN, which can be made available to a high temperature chemistry. −0.006 s−1 exp( −2.7±0.4 kJ mol−1 RT RT ). The NH3 + HCN → NH+ Key words: astrochemistry – ISM: molecules – molecular processes – molecular data – methods: laboratory. 1 INTRODUCTION The HCN molecule plays a fundamental role in interstellar chem- istry as it is the simplest molecular species containing both C and N atoms (while the radical CN is the simplest interstellar species composed of C and N); reactions involving HCN can produce key species intermediate to the formation of amino acids, such as aminoacetonitrile (Danger et al. 2011). HCN has been observed in the gas phase in various en- vironments including molecular clouds (Snyder & Buhl 1971; Clark et al. 1974), circumstellar envelopes (Bieging et al. 1984), the comae of comets (Ip et al. 1990), the atmospheres of Nep- tune (Rosenqvist et al. 1992) and Titan (Tokunaga et al. 1981), and was recently discovered in the Murchison meteorite (Pizzarello 2012). Typical abundances with respect to H2 are 10−9 – 10−8 towards molecular clouds (Hirota et al. 1998) and massive proto- stars (Schreyer et al. 1997), rising to 10−6 in the hot envelope pre- ceding a hot core (Lahuis & van Dishoeck 2000; Boonman et al. 2001) due to a combination of gas phase chemistry and the sub- limation of ices from dust grain surfaces. As such, HCN has been shown to be a linear tracer of ongoing star formation and is used ⋆ Email: [email protected] c(cid:13) 2002 RAS to probe dense gas in star forming regions (Hirota et al. 1998). The gas phase ratios between isomers such as HNC and HCN indicate the extent of photochemical isomerisation that has occurred in as- trophysical environments. The HNC/HCN ratio is highly depen- dent on temperature, and has been observed to be ∼ 1 in prestel- lar cores (Padovani et al. 2011; Herbst et al. 2000), 0.6 – 0.9 in infrared dark clouds (Vasyunina et al. 2011), and less than 0.3 in Neptune's atmosphere (Moreno et al. 2011). The gas phase ratio CN/HCN can change by orders of magnitude between the edge of a molecular cloud and the densest regions of the core, as it is sensitive to photodissociation channels (Boger & Sternberg 2005). A greater knowledge of the grain surface chemistry network be- tween CN and HCN is needed to understand CN/HCN ratios in dense cores (Chapillon et al. 2012). CN− is the first diatomic anion to have been positively identified in an astrophysical environment, having been observed recently in the gas phase in the C-star enve- lope IRC +10216 (Ag´undez et al. 2010). One environment where HCN is of particular interest is in comets and their comae. The CN moiety (part) is key in the pro- duction of amino acids (Ferris & Hagan 1984), recently observed in comets for the first time by the Stardust spacecraft (Elsila et al. 2009). Both HCN and CN have been observed extensively in the gas phase on lines of sight towards cometary comae (e.g. 2 J. A. Noble et al. Huebner et al. 1974; Bockelee-Morvan et al. 1987; Woodney et al. 2002). HCN is believed to be the major source of the CN radi- cal (Paganini et al. 2010), although other sources, such as refrac- tory organic compounds, are necessary in order to produce the observed CN abundances (Fray et al. 2005). In the solid phase, HCN is unlikely to exist solely in its monomeric form, but also in the form of polymers and complex species which act as HCN reservoirs, such as hexamethyltetramine (HMT, Bernstein et al. 1995; Vinogradoff et al. 2011) or poly(methylene-imine) (PMI, Danger et al. 2011). Solid phase HCN has yet to be positively iden- tified in comets, but a tentative detection of solid phase HCN on Tri- ton was made recently using the AKARI telescope (Burgdorf et al. 2010). Experimentally, solid phase HCN has been produced under as- trophysically relevant conditions upon irradiation of CH4:N2 mix- tures with 5 keV electrons (Jamieson et al. 2009). HCN can re- act as an electrophile, as in the reaction of HCN with H atoms in the solid phase which produces methylamine (Theule et al. 2011a), or as a nucleophile, as in its reaction with H2CO to form hy- droxyacetonitrile HOCH2CN (Danger et al. 2012). The participa- tion of HCN in a low temperature solid phase Strecker synthe- sis (a series of reactions to synthesise an amino acid) yields aminoacetonitrile, NH2CH2CN, via thermal processing of an ice mixture of CH2NH, NH3 and HCN (Danger et al. 2011). In addi- tion, the irradiation of HCN in ice mixtures and aqueous solutions with photons or charged particles has been seen to yield complex ions (Gerakines et al. 2004), amino acids (Bernstein et al. 2002), amides, carboxylic acids and bases (Colin-Garc´ıa et al. 2009). HCN is also a weak acid, therefore it could react with bases to produce salts. Previously, acid-base reactions have been shown to yield salts under astrophysically relevant conditions. The re- action HNCO + NH3 produces NH+ 4 OCN− (Demyk et al. 1998; Raunier et al. 2003a; van Broekhuizen et al. 2004) with an activa- tion energy of 0.4 kJ mol−1 (Mispelaer et al. 2012); HNCO + H2O produces H3O+OCN− (Raunier et al. 2003b) with an activation en- ergy of 26 kJ mol−1 (Theule et al. 2011a), which then further reacts to form the isomerisation product HOCN with an activation en- ergy of 36 kJ mol−1 (Theule et al. 2011b); HCOOH + NH3 yields NH+ 4 HCOO− (Schutte et al. 1999). Such species have relatively high sublimation temperatures (greater than ∼ 150 K) compared with the simple molecules such as CO2, H2O and CO which com- pose the bulk of the icy mantles on dust grains (Raunier et al. 2004; Noble et al. 2012a). As such, it is likely that salts will remain on the grain surface following heating processes that sublime the majority of the ice mantle, ultimately forming part of the refractory organic residue observed on comet grains (Glavin et al. 2008). In this paper, we report the results of a study of the purely thermal reaction of the weak acid HCN with the base NH3: HCN + NH3 → NH+ 4 CN−. (1) NH3 has been observed in astrophysical ices at abundances of 2 – 15 % H2O (Lacy et al. 1998; Bottinelli et al. 2010), and is therefore a key basic, nucleophilic species available to react with acidic, electrophilic molecules in grain ice mantles. The 6.0 and 6.85 µm bands are observed in ices towards many astrophysical environments, including young stellar objects (YSOs, Keane et al. 2001; Boogert et al. 2008) and quiescent molecular clouds (Boogert et al. 2011), but have yet to be fully characterised. NH+ 4 is considered to be a likely candidate to account for some of the feature at 6.85 µm (Schutte & Khanna 2003). It has been sug- gested that most of the 6.0 µm excess, and at least some of the Table 1. Experiments performed in the current work. a ID i ii iii iv v vi vii viii Molecules Ratio at deposition HCN : NH3 : H2O : CO2 : CO 0 1 1 pure HCN 0 pure NH3 1 HCN & NH3 1.5 : HCN & NH3 1 : HCN & H2O HCN, NH3 & H2O : 1 HCN, NH3, H2O & CO2 2.5 : HCN, NH3, H2O & CO : 0 0 0 0 : : : : : : : : 20 : : : : 18 : : 30 : : 17 : : 2.5 : 1 0 3 6 1 2 0 0 0 0 0 0 1 0 : 0 : 0 : 0 : 0 : 0 : 0 : 0 : 2 a Ratios calculated from the experimental spectra. 6.85 µm band, is related to highly processed ices (Gibb & Whittet 2002). However, given that the bands are observed in quiescent re- gions, some low-energy formation route to their carriers must be possible. Laboratory studies have shown that NH+ 4 can be produced via acid-base reactions in ices at 10 K with no irradiation (e.g. Raunier et al. 2003a). Thus, reaction (1) is important as it involves the abundant ice species NH3, produces NH+ 4 , and also involves HCN, the simplest molecule containing a CN bond, which could be a prerequisite for the formation of amino acids (Danger et al. 2011). is characterised spectroscopically In this work, the NH+ employing FTIR spectroscopy. The 4 CN− product of the NH3 + HCN using Fourier reaction transform infrared (FTIR) spectroscopy. The kinetics of the NH3 + HCN reaction are investigated using isothermal ex- periments temperature dependence of the reaction rate is determined to be k(T) = 0.016+0.010 ). The band strength of CN− is determined to be ACN− = 1.8 ± 1.5 × 10−17 cm molec−1. Tem- perature programmed desorption (TPD) experiments using mass spectrometry allows calculation of the desorption rate as kdes(T ) = 1028 molecules cm−2 s−1 exp( −38.0±1.4 kJ mol−1 ) for a zeroth-order multilayer desorption. NH+ 4 CN− is therefore more refractory than H2O, and thus can participate in a warm, water-free chemistry. −0.006 s−1 exp( −2.7±0.4 kJ mol−1 RT RT 2 EXPERIMENTAL Experiments were performed using the RING experimental set-up, as described elsewhere (Theule et al. 2011b). Briefly, RING con- sists of a gold-plated copper surface within a high vacuum chamber (a few 10−9 mbar). The temperature of the surface is controlled us- ing a Lakeshore Model 336 temperature controller, a closed-cycle helium cryostat (ARS Cryo, model DE-204 SB, 4 K cryogenera- tor), and a heating resistor; the temperature is measured using a DTGS 670 Silicon diode, with a 0.1 K uncertainty. Molecular species in the form of room temperature gas mix- tures are dosed onto the gold surface by spraying at normal in- cidence via two injection lines. The infrared spectra of the con- densed mixtures are recorded by means of Fourier Transform Re- flection Absorption Infra Red Spectroscopy (FT-RAIRS) using a MCT detector in a Vertex 70 spectrometer. A typical spectrum has a 1 cm−1 resolution and is averaged over a few tens of interfero- grams. In the gas phase, molecules in the high vacuum chamber are measured by means of mass spectroscopy using a Hiden HAL VII c(cid:13) 2002 RAS, MNRAS 000, 1–12 Thermal reactivity of HCN and NH3 3 Figure 1. The infrared absorption spectra of pure HCN and NH3 deposited at 10 K: a) HCN at 10 K, b) HCN after heating to 100 K, c) NH3 at 10 K, d) NH3 after heating to 100 K. RGA quadrupole mass spectrometer (QMS) with a 70 eV impact electronic ionisation source. In the experiments presented here, a series of gas-phase mix- tures of HCN, NH3, H2O, CO and CO2 were prepared in the in- jection lines. Ammonia is commercially available as a 99.999 % pure gas from Air Liquide. The gas phase HCN monomer was syn- thesised via the thermal reaction of potassium cyanide, KCN, with an excess of stearic acid, CH3(CH2)16COOH, in a primary pumped vacuum line (Gerakines et al. 2004). A small quantity of CO2 was produced during HCN synthesis due to the decomposition of stearic acid. H2O vapour was obtained from deionised water purified by several freeze-pump-thaw cycles carried out under vacuum. A complete list of experiments performed in this study is pre- sented in Table 1. In all experiments, NH3 was dosed via one beam line, while the other species were dosed via a second line. Gas mix- tures were dosed onto the gold surface, held at 10 K. Except where noted in the text, the surface was subsequently subjected to a heat- ing ramp of 2 K min−1, and infrared spectra were recorded at inter- vals of 10 K. Molecules desorbing from the surface were monitored by mass spectroscopy. 3 RESULTS 3.1 Pure HCN and NH3 Figure 1 presents the infrared spectra of pure HCN (Experiment (i)) and pure NH3 (Experiment (ii)) deposited at 10 K. The spectra have absorption bands at characteristic frequencies, which are tabu- lated in Table 2. The spectra of NH3 (d'Hendecourt & Allamandola 1986; Sandford & Allamandola 1993) and HCN (Bernstein et al. 1997; Gerakines et al. 2004) are well documented in the literature; c(cid:13) 2002 RAS, MNRAS 000, 1–12 NH3 is characterised by four fundamental vibrational modes ν1 – ν4 with frequencies 3212, 1064, 3375, and 1625 cm−1, respectively, while HCN displays vibrational modes ν1 – ν3 with frequencies 3143, 822, and 2100 cm−1, respectively, as well as a prominent ab- sorption at 1617 cm−1 assigned as the first overtone of the HCN bending mode. The pure HCN and NH3 ices were heated from 10 – 150 K at a heating rate of 2 K min−1; the spectra of NH3 and HCN at 100 K are also presented in Figure 1. Between 70 – 80 K, crystallisation of NH3 results in a rapid change in position and form of the four main bands: ν1 shifts to lower frequency (3206 cm−1), ν2 and ν3 split into two peaks (1071 and 1093 cm−1, and 3378 and 3364 cm−1, respectively), and ν4 shifts to higher frequency (1649 cm−1). These changes are visible in the spectrum of pure NH3 at 100 K, shown in Figure 1, trace d. The crystallisation of pure HCN occurs gradually between 10 – 120 K, with no definitive change, unlike for NH3: ν1 shifts to lower frequency (3139 cm−1) at temperatures up to 60 K, before increasing to 3144 cm−1 by 120 K; ν2 and ν3 both shift to higher frequency above 80 K (2099 and 823 cm−1, respectively, likely due to crystallisation of the HCN (Danger et al. 2011)); the overtone at 1617 cm−1 shows similar temperature dependence as ν1, shifting to 1620 cm−1 by 60 K then returning to 1617 cm−1 by 120 K. Crystallisation is evinced by the difference between the low and high temperature spectra of Figure 1. The temperature programmed desorption of the pure species was measured by quadrupole mass spectrometry. The major frag- ments (relative intensity > 10 %) of NH3 are the parent fragment (m/z 17, 100 %) and the loss of one hydrogen at m/z 16 (86 %); for HCN the major fragments are m/z 27 (100 %) and m/z 26 (17 %). The resulting QMS spectra of the TPD of pure NH3 and pure HCN are plotted in the top panel of Figure 2. Thermal desorption is an activated process. The rate of des- 4 J. A. Noble et al. Table 2. Fundamental infrared band positions (cm−1), for the species HCN, NH3, H2O, CO and CO2 in various mixtures (as defined in Table 1). Unless otherwise stated, all values are for ices at 10 K. Vibration mode Assignment HCN/CN− NH3 NH3 CO2 CO NH+ 4 4 CN− c 4 CN− c NH+ NH+ ν3 ν3 ν2 ν3 ν1 ν4 ν4 ν1 CN stretch NH stretch umbrella CO stretch CO stretch NH bend NH bend CN stretch HCN (i) 2100 – – 2343b – – – – NH3 (ii) – 3375 1064 2342b – – – – HCN:NH3 (iv) (iii) 2077 3380 1098 2342b – 1479 1435 2092 2098 3398 1104 2343b – 1471 1435 2092 HCN:H2O HCN:NH3:H2O (v) 2091 – – 2342b – – – – (vi) 2090 3378 1101 2342b – – 1437 2092 HCN:NH3:H2O:X X = CO2 (vii)a X = CO (viii) 2086 3399 1112 2343b 2138 1498 n/ac n/ac 2086 3385 1113 2340 – 1482 1437 2092 a Deposition (and spectrum measured) at 15 K. b Present as a trace pollutant (produced during the HCN synthesis). c Isolated on the surface at high temperature, after the desorption of NH3 and HCN. The ice mixture containing CO (Experiment (viii)) was not heated to sufficiently high temperature to observe these bands. orption by unit surface, r, can be expressed by the Wigner-Polanyi equation (King 1975), where the desorption rate constant kdes is de- scribed in terms of an Arrhenius law: r = − dN dt = kdes Nn = A e−Edes/RT Nn, (2) where A is the pre-exponential factor, Edes is the energy of des- orption of a molecule from the surface (J mol−1), R is the gas con- stant (J K−1 mol−1), T is the temperature of the surface (K), N is the number of adsorbed molecules on the surface (molecules cm−2), and n is the order of the reaction. The units of A depend on n: molecules1−n cm−2+2n s−1. By rearranging Equation (2) as: r = − dN dT = A β e−Edes /RT Nn, (3) where β = dT dt = 2 K min−1 is the heating rate, the experimental data can be analysed. In order to calculate the energies of desorption of the isolated species HCN and NH3, zeroth order desorption kinetics are assumed, as is standard practise for the multilayer desorption of bulk material (see e.g. Collings et al. 2004; Burke & Brown 2010; Noble et al. 2012a). In this case, A has units of molecules cm−2 s−1. The energy of desorption, Edes was calculated using two meth- ods. Firstly (Method 1), assuming that the lattice vibrational fre- quency of the solid is 1013 s−1 and the number of molecules in a monolayer is approximately 1015 cm−2, the pre-exponential factor, A, can be fixed at a value of 1028 molecules cm−2 s−1, and the desorption energy calculated as the only free parameter in the fit of Equation 3 to the experimental data (Collings et al. 2003; Fuchs et al. 2006). Using this method, the calculated desorption energies were: Eads,HCN = 30 ± 1 kJ mol−1, and Eads,NH3 = 26 ± 1 kJ mol−1. In the second method of fitting (Method 2), as proposed by Hasegawa et al. (1992) the pre-exponential factor in Equation (3) was assumed to be a function of Eads approximated by: A = NML.ν = NMLr 2NMLEads π2M , (4) where M is and NML ∼ 1015 cm−2. The advantage of this method is that the adsorbate molecule, the mass of the fit requires only one variable, Edes, rather than assuming or fitting the pre-exponential factor and fitting Edes (Acharyya et al. 2007; Noble et al. 2012a). Using the second method, the de- rived desorption energies were: Eads,HCN = 28 ± 1 kJ mol−1, and Eads,NH3 = 25 ± 1 kJ mol−1. The pre-exponential factors were: AHCN =1.5 × 1027 ± 5.2 × 1023 molecules cm−2 s−1 and ANH3 = 1.7 × 1027 ± 2.0 × 1024 molecules cm−2 s−1. For both species, the values of the desorption energies cal- culated by the two methods fall within the limits of uncer- tainty. As discussed in Noble et al. (2012b), both methods are valid, since both pairs of parameters {A, Edes} give the same value for the quantity of physical importance, i.e. the kdes des- orption rate constant. We prefer to use the coupled solutions {A = 1028 ± 0 molecules cm−2 s−1, Eads,HCN = 30 ± 1 kJ mol−1}, and {A = 1028 ± 0 molecules cm−2 s−1, Eads,NH3 = 26 ± 1 kJ mol−1}. This latter value is consistent with the published value of 25.6 ± 0.2 kJ mol−1 for NH3 (Sandford & Allamandola 1993). 3.2 HCN and NH3 mixtures The HCN and NH3 gases were prepared in separate molecular beams and co-deposited on the gold surface held at 10 K, with an excess of NH3 in order to isolate the chemical reaction kinetics from the diffusion kinetics. Although this mixture (without H2O) is not directly astrophysically relevant, an excess of NH3 favours the complete reaction, gives HCN an homogeneous NH3 environment and thus produces reaction kinetics of (pseudo-) order one in HCN. 3.2.1 Identification of the NH3 + HCN reaction product The infrared spectra of Experiment (iii), a mixture of HCN and NH3, are presented in Figure 3. The bands attributed to HCN and NH3 are slightly shifted with respect to their values in pure HCN and NH3, as can be seen by comparing the values in Table 2. The presence of new bands in the spectrum at 10 K (c.f. Figure 1 and Table 2) also suggests that the species have undergone a reaction in the gas phase at 300 K during the deposition, before thermalisation to 10 K. It has previously been suggested that this reaction could also occur as a result of the release of energy during condensation (Gerakines et al. 2004). The wide band at 1479 cm−1 is attributed to the NH+ 4 ion, which has been extensively studied in astrophysically relevant ices c(cid:13) 2002 RAS, MNRAS 000, 1–12 Thermal reactivity of HCN and NH3 5 NH+ We identify the product of the NH3 + HCN reaction to be the 4 CN− species with the following four arguments: • the identification of the NH+ 4 CN− absorption bands in the IR spectrum of the HCN:NH3 mixture upon heating to 150 K (Fig- ure 3, left hand panel), • in the TPD spectrum of the HCN:NH3 mixture (shown for m/z 27 in the lower panel of Figure 2, and discussed further in § 3.2.6), part of the product desorbs when dominant NH3 desorbs around 135 K. The mass spectrum of the pure product is obtained between 140 K and 160 K, as shown in Figure 2. Mass features for the prod- uct are obtained at m/z 27 (100 %), 26 (17 %), 17 (20 %) and 16 (15 %) at 150 K. The observed mass features of the product are a combination of the pure HCN and NH3 desorption features. These features thus correspond to a proton transfer during the reverse re- action NH+ 4 + CN− → NH3 + HCN and the subsequent rapid des- orption of HCN and NH3 at high temperature (140 – 160 K). Within the observational limitations of our experiments, no m/z 44 feature corresponding to the molecular ion NH+ 4 CN− is observed, • the similarity of this system to the low temperature proton 4 + OCN− (Demyk et al. transfer reaction NH3 + HNCO → NH+ 1998; Raunier et al. 2003a; van Broekhuizen et al. 2004), • the previous literature evidence of the spontaneous reaction between HCN and NH3, represented in Equation (1), upon deposi- tion (Gerakines et al. 2004; Clutter & Thompson 1969). We can therefore assign the product of the NH3 + HCN reac- tion to the NH+ 4 CN− species. A number of values can be derived from experiments on HCN:NH3 mixtures: the band strength of the CN− ion, ACN, the activation energy of the HCN + NH3 reaction, and the desorption energy of the salt NH+ 4 CN− from the surface. 3.2.2 The HCN bending mode at 848 cm−1. In a mixture of HCN:NH3, the only isolated HCN band (which does not overlap with bands of NH3 or CN−) is the HCN bend at 848 cm−1. Figure 4 (right panel) illustrates the effect of the envi- ronment on this band; spectra are plotted for Experiments (i) pure HCN, (iv) low NH3 concentration, and (iii) high NH3 concentration at 10 K. The band widens and shifts redwards with increasing NH3 concentration. The presence of the broad bending mode absorption feature at 800 cm−1 in the spectrum of H2O prevents the study of the HCN bending mode in mixtures containing H2O. In a sample of pure NH+ 4 CN− at 20 K (Figure 4, trace d), prepared by mixing gaseous HCN and an excess of NH3 in the injection lines before dosing, the HCN bending mode is not seen, as all HCN is converted to CN− before dosing onto the surface (Danger et al. 2011). In all of the experiments presented here (without H2O) the HCN bending mode is clearly observed, indicating that in all experiments some HCN remains unreacted on the surface at 10 K. Conversely, the ν2 and ν3 bands of NH3 both increase in fre- quency with the addition of higher concentrations of HCN. The ν2 band shifts from 1064 cm−1 in pure NH3 (Experiment (ii)) to 1098 cm−1 with a low HCN concentration (Experiment (iii)) to 1104 cm−1 with a higher HCN concentration (Experiment (iv)), while the ν2 band shifts from 3375 cm−1 to 3380 cm−1 to 3398 cm−1 in the same experiments. The absorption strength of the HCN bending mode in pure HCN at 10 K was calculated from the data in Experiment (i) to be Aδ HCN = 1.9 ± 0.7 × 10−18 cm molec−1, using the area under the HCN bending mode and the HCN CN stretching mode peaks, and Figure 2. The temperature programmed desorption spectra of a) pure NH3, 4 CN−. The experimental desorption spectra are plot- b) pure HCN, c) NH+ ted as solid lines (m/z 27 in black and m/z 17 in grey). Overplotted on these peaks are the best fitting zeroth order desorption kinetics for each molecular species; the fixed A method is plotted as a dotted line, while the parame- terised A method is plotted as a dashed line. The fitting approach is fully described in the text, where the two methods are denoted as Method 1 and Method 2, respectively. (Raunier et al. 2003b, 2004; Gerakines et al. 2004; G´alvez et al. 2010). In addition to the broad band centred at 1479 cm−1, which is assigned to the ν4 NH bending mode, NH+ 4 exhibits a broad fea- ture in the 3000 cm−1 region, which is the superposition of the ν3 NH stretching mode, the 2ν4 overtone, and the ν2 + ν4 combina- tion mode (Wagner & Hornig 1950; Clutter & Thompson 1969). In the 3000 cm−1 region, at low temperatures (before NH3 and HCN desorption), the NH+ 4 contribution is difficult to deconvolve from the NH3 contribution. At 150 K, when NH3 and HCN have des- orbed from the surface, the IR spectrum of the product is isolated, and the NH+ 4 feature is revealed as a three-peaked band with max- ima at 3168 (ν2 + ν4), 3014 (ν3), and 2845 cm−1 (2ν4). We ob- tain the same NH+ 4 CN− IR spectrum as Gerakines et al. (2004) and Clutter & Thompson (1969). The band at 1479 cm−1 grows and shifts redwards as the mix- ture of HCN and NH3 is heated from 10 K to 150 K, as illustrated in the right hand panel of Figure 3. Upon the desorption of both reactants, HCN and NH3, between 110 and 150 K (as described in Figure 2) the band shifts to 1435 cm−1. This shift is due to both a change in environment and to crystallisation. The production of NH+ 4 necessitates the presence of an anion in the ice. We positively identify the counter ion as CN− from its IR band at 2092 cm−1 (Figure 3, middle panel). As seen for the 4 band at 1479 cm−1, the HCN/CN− band at 2077 cm−1 shifts NH+ during heating (although it shifts bluewards), and between 110 and 150 K the band narrows and shifts to 2092 cm−1 due to a changing environment and crystallisation of the NH+ 4 CN− product. c(cid:13) 2002 RAS, MNRAS 000, 1–12 6 J. A. Noble et al. Figure 3. IR spectra of Experiment (iii), a mixture of HCN and NH3 with an excess of NH3. Left panel: Upon deposition at 10 K, and after heating to 150 K. Middle panel: evolution of the ν3 CN stretching mode of HCN/CN− as a function of temperature. Right panel: evolution of the ν4 NH bending mode band of NH+ 4 as a function of temperature. 3.2.3 The CN stretching mode at 2100 cm−1. It is particularly interesting to note the effect of the environment on the CN stretch at 2100 cm−1, as the band is very complex, and the contributions of the species HCN and CN− are largely irresolvable. The spectra of Experiments (i) pure HCN, (iv) low NH3 concentra- tion, and (iii) high NH3 concentration at 10 K are plotted in Figure 4 (left panel), as for the HCN bending mode discussed in the previous section. The band at 2100 cm−1 behaves very similarly to the HCN bending mode at 848 cm−1, widening and shifting redwards with in- creasing NH3 concentration. It is evident from Figure 4 that in the different experiments, different components contribute to the over- all profile of the band. By comparison to trace d in the same figure (pure NH+ 4 CN−), it can be inferred that an increasing concentration of NH3 results in an increased conversion of HCN to CN− in our mixtures at 10 K. However, due to the complexities introduced by the effects on the band position and profile of concentration and, as previously discussed with respect to Figure 2), temperature, it has not been possible to resolve the contributions of the species HCN and CN− to the CN stretching mode in a reliable and reproducible manner across all of our spectra. 3.2.4 Determination of the CN− ion band strength. In order to determine the band strength of the CN− ion, it is neces- sary to calculate the CN− ion column density in our ice mixtures. Due to the difficulty of deconvolving the contributions of HCN and CN− at 2092 – 2099 cm−1 discussed above, it is not possible to evaluate directly the number of CN− ions in the ice from its ab- sorption band at low temperature. At temperatures above 140 K, all HCN has desorbed from the surface, and therefore the fea- c(cid:13) 2002 RAS, MNRAS 000, 1–12 Figure 4. A comparison of the CN stretching mode (left panel) and HCN bending mode (right panel) absorption bands at 10 K for a) pure HCN (Ex- periment (i)), b) HCN in a low concentration of NH3 (Experiment (iv)), and c) HCN in an excess of NH3 (Experiment (iii)). Trace d is a spectrum of pure NH+ 4 CN− at 20 K taken from (Danger et al. 2011). The dotted lines mark the positions of the band maxima in Experiment (i), and are included to guide the eye. the absorption strength of the CN stretching mode (Bernstein et al. 1997). Thermal reactivity of HCN and NH3 7 ture at 2092 cm−1 (4.78 µm) derives solely from absorption by the CN− ion. The assumption is made that Equation 1 holds and therefore the number of CN− ions is equivalent to the number of NH+ 4 ions. Thus, using the area under the absorption feature at 2092 cm−1 and the number of NH+ 4 ions in the ice (derived from ten HCN + NH3 experiments) the average absorption strength was calculated as ACN− ,140 K = 1.8 ± 1.5 × 10−17 cm molec−1. The uncer- tainty is calculated from the spread of values in the ten experiments, and is significant, reflecting the complexity of deriving such values from laboratory data. the trace d) absorption Using a spectrum of pure NH+ from Danger et al. strength 4 CN− at 20 K (shown (2011), we cal- in Figure 4, culated 20 K as ACN−,20 K = 1.8 ± 0.1 × 10−17 cm molec−1. This result is identical to the value calculated at high temperature, suggesting that the CN− band in pure NH+ 4 CN− is not highly susceptible to temperature changes in the range of 20 – 140 K. of CN− at The band strength of the CN− ion is comparable to the band strength of most other interstellar molecules, unlike the OCN− ion, which exhibits a high band strength (1.3 × 10−16 cm molec−1 van Broekhuizen et al. 2004) that compensates for its low abun- dance and makes its detection possible in the IR spectra of inter- stellar ices. Further calculations were performed on data from Experi- ments (iii) and (iv) in an attempt to deconvolve the contributions of HCN and CN− to the CN stretch absorption at 2100 cm−1. Firstly, the number of HCN molecules in Experiment (iv) was calculated from its absorption band at 848 cm−1 (Figure 4, trace b), using the absorption strength calculated in § 3.2.2 (assuming that the magni- tude does not change with environment), to be ∼ 9.3 × 1017. Sec- ondly, using the absorption strength calculated above for ACN− ,20 K at 2100 cm−1, it was possible to calculate the number of CN− radicals in Experiment (iv), and thus infer the number of HCN molecules as ∼ 2.5 × 1018, approximately 1.7 times higher than that calculated by the first method. The large uncertainty on this value most likely derives from the assumptions made in the calculations: that the absorption strengths of the CN− and HCN CN stretching mode absorptions, and the HCN bending mode absorption do not change with environment. As has been previously demonstrated by Borget et al. (2012), the CN stretch in nitriles is very sensitive to environment, with the absorption strength of CN in aminoacetoni- trile changing by a factor of two between the pure molecule and a mixture of aminoacetonitrile and H2O (1:3). Thus we conclude that, while we can accurately determine the absorption strength of CN− at low to intermediate temperatures in a pure salt, it is not possible from these data to reliably deconvolve HCN and CN− contributions to the CN stretching mode absorption at 2100 cm−1. The implication of this for future observations is that in the infrared spectra of ices where both HCN and CN− species are present, their bands are likely to be superimposed upon one an- other. Given the large difference in absorption strengths of the two species – 5.1 × 10−18 cm molec−1 for HCN (Bernstein et al. 1997) and 1.8 × 10−17 cm molec−1 for CN− – it is important that the band not be treated as one species, but that efforts are made to deconvolve the individual contributions of HCN and CN−. Further experimen- tal studies are required to address this complex problem. 3.2.5 Determination of the reaction rate constant to of order determine the In NH3 + HCN → NH+ 4 CN− reaction, isothermal kinetic exper- iments were performed, following Bossa et al. (2009). A series activation energy the c(cid:13) 2002 RAS, MNRAS 000, 1–12 Figure 5. Time evolution of the normalised abundances of the NH3 reactant 4 product at different fixed temperatures. Top panel: NH3 evo- and the NH+ lution at 80 K (diamond) and 90 K (triangle). Bottom panel: NH+ 4 evolution at 60 K (square) and 80 K (cross). Overplotted on the experimental data are the fits of Equation 5. Table 3. Experimental rate constant, k, as a function of temperature. Temperature (K) k (× 10−4 s−1) 60 70 80 90 100 105 0.54 ± 0.11 0.91 ± 0.27 2.64 ± 0.88 4.35 ± 1.00 4.11 ± 0.93 7.09 ± 0.48 of isothermal experiments was carried out in which mixtures of HCN:NH3 with large excesses of NH3 (the ratios were typically 1:15) were deposited at 15 K, heated as quickly as possible (∼ 10 K min−1) to a specific temperature, and held at that fixed temperature for a number of hours, in order to study the reaction kinetics. The temperatures chosen were 60, 70, 80, 90, 100, and 105 K. Above 105 K, desorption of the reactant molecules becomes important, and below 60 K the reaction proceeds so slowly that the experiments become prohibitively long to carry out. The abundances of NH3 and NH+ 4 in the ice mixtures were monitored via their characteristic absorption features in the IR spectra at 3375 and ∼ 1440 cm−1, respectively. The time evolution of the species' molar fractions for four experiments are plotted in Figure 5. The other experimental data is omitted for clarity. It is evident from this figure that an increase in temperature increases the rate of the reaction, particularly in the first few hours of each experiment. As NH3 was in excess with respect to HCN, each HCN molecule was surrounded by a homogeneous NH3 environment. Thus, we can express the rates of destruction of NH3 and forma- tion of NH+ 4 with first-order reaction kinetics, in terms of the molar fraction of the reactants: 8 J. A. Noble et al. 3.2.6 Determination of the NH+ 4 CN− desorption rate constant A TPD spectrum of NH+ 4 CN− (m/z 27 and m/z 17) is shown in the lower panel of Figure 2. It is evident in this TPD trace that some of the product (or unreacted HCN and NH3) co-desorbs with the dom- inant species between ∼ 120 – 140 K (similarly to the pure HCN in the upper panel) and that the NH+ 4 CN− desorbs at higher tem- perature. It is not possible from the experiments presented here to determine at what temperature NH+ 4 CN− first starts to desorb from 4 CN− present the surface, as in none of our experiments was the NH+ on the surface with no HCN. Thus in all TPD traces, the signals of the desorption of HCN and NH+ 4 CN− overlap. The dominant mass fraction for NH+ 4 CN− is m/z 27, but as mentioned in § 3.2.1 above, other masses observed desorbing at this temperature include m/z 26 (17 %), 17 (20 %), 16 (15 %), and 15 (2 %). Under current ex- perimental conditions there is no evidence of the desorption of m/z 44 or 43, and thus it is concluded that NH+ 4 CN− desorbs from the surface entirely as HCN and NH3, following the reverse reaction NH+ 4 + CN− → NH3 + HCN, rather than as NH+ The zeroth-order shape of the m/z 27 NH+ 4 CN−. 4 CN− curve in Fig- ure 2 is less marked since only a few monolayers of NH+ 4 CN− are desorbing. Using the two methods described in § 3.1 above (fitting to the m/z 27 desorption peak only, and taking the average of val- ues calculated from five desorption spectra) the energy of desorp- 4 CN− was calculated as Eads,NH4CN = 38.0 ± 1.4 kJ mol−1 tion of NH+ (Method 1) and Eads,NH4CN = 35.4 ± 1.3 kJ mol−1 (Method 2, with A = 1.3 × 1027 ± 6.0 × 1023 molecules cm−2 s−1). When consider- ing the energies derived for all three molecular species, the values calculated using the fixed A method are consistently 6 – 7 % higher than those calculated using the second method (A as a function of Eads). 3.3 HCN and NH3 mixtures diluted in H2O In order to investigate the reaction of HCN and NH3 under con- ditions relevant to ices in the interstellar medium, H2O must be included in the mixtures. H2O is the most abundant solid phase in- terstellar molecule, with abundances of 10−4 with respect to H2. 3.3.1 HCN:H2O ice mixture Experiment (v) investigates the effect of dilution of HCN in a water ice at a ratio of 1:18. It is clear from the band positions tabulated in Table 2 that the peak position of the CN stretching absorption in HCN is shifted redward (with respect to that of pure HCN) when diluted in a H2O ice (to 2091 cm−1), although the shift is not as marked as that seen for the HCN/CN− band in a HCN:NH3 mixture (2077 cm−1, Experiment (iii)). As expected, no reaction is seen for a mixture of HCN:H2O, neither at 10 K nor during heating to 180 K, where both HCN and H2O have desorbed. 3.3.2 HCN:NH3:H2O ice mixture Experiment (vi) is a mixture of HCN:NH3:H2O with concentration ratio 1:3:30. This experiment investigates the HCN + NH3 reaction when heavily diluted in H2O, as would be expected in interstellar ices. The IR spectrum of this mixture at 10 K is shown in Figure 7, trace a. The addition of H2O shifts the HCN/CN− and NH3 bands with respect to their positions in HCN:NH3 mixtures. The HCN/CN− ν3 band is observed at 2090 cm−1, while the NH3 ν2 and ν3 bands are centred at 1101 and 3378 cm−1, respectively. c(cid:13) 2002 RAS, MNRAS 000, 1–12 Figure 6. Experimentally determined ln(k) as a function of the inverse of the temperature, and the fit against an Arrhenius law (dot-dashed line). dt d(HCN) d(NH3) 4 CN−) d(NH+ dt dt = −k(T ).(HCN) = −k(T ).(HCN) = k(T ).(HCN) (5)  The first-order kinetics in HCN implies that the molar fraction of NH3 is constant. Although not necessary in the first-order case, it is convenient to express the kinetic equations in terms of unitless molar fractions, which allows the calculation of a reaction rate con- stant k in s−1. Resolving Equation 5 for the isothermal experimental curves gives a reaction rate constant at a fixed temperature. The fits of four of the experiments are overplotted on the experimental data in Figure 5. The reaction rate constants, determined at the chosen 4 band at ∼ 1440 cm−1 only, are temperatures by analysis of the NH+ displayed in Table 3. The temperature dependence of the reaction rate constant is plotted in Figure 6, in the form of ln(k) as a function of the in- verse of the temperature. Assuming k(T ) follows an Arrhenius law, a least-squares straight line fit to the experimental data al- lows the calculation of the activation energy of the reaction HCN + NH3. We derive an activation energy of 2.7 ± 0.4 kJ mol−1 with a pre-exponential factor of 0.016+0.010 −0.006 s−1. The uncertainty on the calculated values includes the dispersion of the results measured at 80 K and 90 K, where two experiments were per- formed at each temperature. The value for the activation en- ergy is consistent with values measured previously for different thermal reaction systems (Bossa et al. 2009; Theule et al. 2011b; Mispelaer et al. 2012). Similarly, as observed previously, the pre- exponential factor is low, which may be due to the solvent cage effect in the solid solutions, i.e. the need for the molecules to orientate within the solid into favourable steric configuration be- fore reaction. Therefore the reaction rate constant has a k(T ) = 0.016+0.010 ) temperature dependence within the small temperature interval (60 – 105 K) where we have been able to measure it. The question of the extrapolation to lower tem- peratures is critical to interstellar ice chemistry, but difficult to ad- dress experimentally. −0.006 s−1 exp( −2.7±0.4 kJ mol−1 RT Thermal reactivity of HCN and NH3 9 Figure 7. Left panel: Infrared spectra of a) a HCN:NH3:H2O mixture (Experiment (vi), 10 K), b) a HCN:NH3:H2O:CO2 mixture (Experiment (vii), 15 K), c) a HCN:NH3:H2O:CO mixture (Experiment (viii), 10 K). Right panel: shift of the HCN/CN− band as a function of the ice mixture. The dotted line marks the position of the CN stretching mode absorption band in pure HCN. For this mixture, the NH+ 4 band at 1460 cm−1 is not observed as a distinct peak immediately after deposition at 10 K, unlike in mixtures of HCN:NH3 only. There is perhaps a trace amount of NH+ 4 CN− initially present in the mixture, but it is likely that the dilution of HCN and NH3 in a H2O matrix prevents their direct reaction in the gas phase during the deposition. During the heat- ing of the ice mixture, a wide and shallow NH+ 4 band, centered at ∼ 1490 cm−1 develops in the spectra. At temperatures of 160 – 170 K (when almost all of the H2O has desorbed from the sur- face), this band shifts redwards to 1437 cm−1. The HCN/CN− band at 2090 cm−1 shifts slightly redwards (6 – 7 cm−1, to around 2086 cm−1) during the heating of the ice mixture, due to the H2O environ- ment, the conversion of HCN to CN−, the crystallisation of CN−, or a combination of these effects. Upon desorption of the H2O, when 4 CN− is the only species remaining on the surface, the peak of NH+ the band shifts to 2092 cm−1. 3.3.3 HCN:NH3:H2O:CO2 and HCN:NH3:H2O:CO ice mixtures Further experiments were performed on astrophysically relevant ice mixtures (Experiments (vii) and (viii)) to investigate the effect of including CO2 and CO in the ice on the position of the CN− absorption band. We wanted to investigate the possibility that the CN− ion could contribute to the band at 2041 cm−1 (4.90 µm), as- signed to OCS (Gibb et al. 2004; Dartois 2005), or to the so-called "XCN" band at 2165 cm−1 (4.62 µm) observed towards many as- trophysical environments, including high and low mass YSOs (e.g. Soifer et al. 1979; Gibb et al. 2000; van Broekhuizen et al. 2005; Aikawa et al. 2012). It has been previously observed that the ice absorption features of ionic species can be very sensitive to their c(cid:13) 2002 RAS, MNRAS 000, 1–12 Figure 8. The normalised abundance of NH+ 4 produced during the temper- ature ramps in Experiments (iii) and (vi). The reaction between HCN and NH3 without H2O (Experiment (iii), diamonds) occurs more rapidly than the reaction in a H2O matrix (Experiment (vi), triangles). Values are nor- malised to the abundance at 100 K, before the desorption of NH3 and HCN from the surface begins. chemical environment (Schutte & Khanna 2003) and could shift by a few wavenumbers. The ice mixture containing CO2 at 15 K is shown in Fig- ure 7 (left panel, trace b), while the ice mixture containing CO at 10 K is shown in trace c of the same figure. The HCN/CN− and 10 J. A. Noble et al. NH3 absorption bands are shifted by the presence of these two species in the ice, as tabulated in Table 2. Figure 7 (right panel) shows the spectral region containing the HCN/CN− stretching ab- sorption for all three astrophysically relevant ice mixtures studied. The traces are labelled as in Figure 7 (trace a) HCN:NH3:H2O, trace b) HCN:NH3:H2O:CO2, trace c) HCN:NH3:H2O:CO). The CN absorption feature of HCN/CN− is observed at 2100 cm−1 in pure HCN, at 2091 cm−1 when diluted into H2O, and at 2086 cm−1 in HCN:H2O:CO2 and HCN:H2O:CO mixtures, is around 45 – 60 cm−1 from the 2041 cm−1 band and around 65 – 80 cm−1 from the 2165 cm−1 band. By comparison of the absorption features in Figure 7, and the band positions in Table 2, it is clear that the addition of H2O, CO2 and CO to ice mixtures containing HCN and NH3 is not sufficient to induce a significant shift in the position of the HCN or CN− fea- tures. The presence of CO and CO2 in the mixtures shifts the peak to slightly lower frequency (2086 cm−1). In the spectra presented here, it is not possible to deconvolve the contributions of HCN and CN− to the absorption feature centred at ∼ 2090 cm−1. No absorp- tion features are observed in the region of 2165 – 2175 cm−1. We conclude that CN− can not contribute to either the 2041 cm−1 band, assigned to OCS, or to the "XCN" band. If present in ices, it would likely be observed at or very close to 2092 cm−1. Regarding the kinetics of the reaction between HCN and NH3, it is evident that dilution in a H2O matrix slows the formation of NH+ 4 CN−, as illustrated in Figure 8. Quantification of the effect of diffusion on the kinetics of a thermal reaction is extremely compli- cated, and will be the subject of a forthcoming article (Mispelaer et al., in preparation). In this article, we have not attempted to further study the kinetics of the HCN + NH3 reaction in mixtures contain- ing H2O. 4 ASTROPHYSICAL IMPLICATIONS In low mass dark molecular clouds, the average abundance of gas phase HCN is around 1 – 1.5 times higher than that of HNCO (see references in Vasyunina et al. 2011), and thus if OCN− is present in interstellar ices, it is reasonable to assume that HCN and CN− could equally be present. Gas phase HCN is present in molecu- lar clouds at abundances (with respect to H2) of ∼ 10−7 – 10−8, while solid phase H2O, the most abundant solid phase species, is present at abundances of ∼ 10−4. We can thus assume that the abun- dance of HCN in the solid phase is approximately 0.1 % H2O. With values of 5.1 × 10−18 cm molecule−1 (Bernstein et al. 1997) and 1.8 × 10−17 cm molecule−1 (this work) for the CN stretch of HCN and CN−, respectively, the band strengths of these molecules are not high enough to compensate for their expected low inter- stellar abundances. Thus, given the sensitivity of current infrared telescopes, these two molecules are unlikely to be detected in ice mantles. An HCN detection has been reported in Titan, where at- mospheric chemistry has resulted in a higher abundance of solid phase HCN (Burgdorf et al. 2010). Moreover, neither the HCN nor the CN− band position in different ice environments matches the 4.90 µm (2041 cm−1) band currently assigned to OCS, nor can these species contribute to the 4.42 µm "XCN" band. However, NH+ 4 CN− may contribute to the 6.85 µm band through its NH+ 4 bending ab- sorption band, as discussed above. The chemistry of HCN is of primary importance in interstel- lar ices, since HCN is the simplest molecular species with a CN bond and, as such, HCN is probably at the origin of some prebiotic molecules. An illustration of the chemical network centred upon HCN and NH+ 4 CN− is shown in Figure 9. HCN can be produced by the hydrogenation of the CN radical, and then hydrogenated into CH2NH and CH3NH2 (Theule et al. 2011a). As the CN− moiety of the NH+ 4 CN− species is more nucleophilic than HCN, it could be the source of a rich ice chemistry. Its recent detection in the gas phase highlights its potential importance, and validates stud- ies into the formation routes of interstellar ions (Ag´undez et al. 2010). NH+ 4 CN− reacts with H2CO to form hydroxyacetonitrile HOCH2CN (Danger et al. 2012). NH+ 4 CN− can also react with the HCN hydrogenation product CH2NH to give aminoacetonitrile NH2CH2CN, which can go on to form glycine according to the Strecker synthesis. However, the competing reaction of NH+ 4 CN− with CH2NH can also form the poly(methylene-imine) CN-(CH2- NH)n-H (CN-PMI, Danger et al. 2011). The reaction product de- pends on the starting concentrations of NH+ 4 CN− and CH2NH; an excess of CH2NH promotes the polymerisation of CH2NH and the production of CN-PMI, while an excess of CN− promotes the pro- duction of aminoacetonitrile (and ultimately glycine). It is likely that CN− will react with other electrophiles in interstellar ices. The reaction HCN + NH3 → NH+ 4 CN− is the preliminary step neces- sary to the reactions discussed above and, as such, it is important to quantify the kinetics of this rate limiting reaction. Indeed, without H2O, HCN does not react with CH2O (Woon 2001). It is worth emphasising that the reactions in Figure 9 are ther- mally activated, meaning that they do not need any non-thermal process in order to occur at low temperature in astrophysical ices. The competition and the resulting branching ratios for these ther- mal reactions are entirely determined by the temperature depen- dence of their rate constants. As the reaction NH3 + HCN is the first step in the network, its kinetics are limiting to the overall net- work. The measured barrier to this reaction (2.7 ± 0.4 kJ mol−1) 4 OCN− is a little high compared to the formation of the salt NH+ (0.4 kJ mol−1, Mispelaer et al. 2012), but lower than the forma- tion of more complex species such as aminomethanol (4.5 kJ mol−1, Bossa et al. 2009). This simplified network is not complete; in or- der to fully explain the kinetics surrounding NH+ 4 CN−, it is nec- essary to include all reactions, in particular those originating from NH3, such as the NH3 + CO2 reaction, which can be either ther- mally activated (Bossa et al. 2008) or activated by low-energy electrons (Bertin et al. 2009). In addition to constraining its reaction network, it is impor- tant to understand how much HCN can be stored in the NH+ 4 CN− form, whether in molecular clouds, or in higher temperature en- vironments such as comets. When the HCN hydrogenation prod- uct CH2NH reacts with HCOOH, rather than with NH+ 4 CN−, the complex molecules hexamethyltetramine (HMT, C6H12N4) or a second form of poly(methylene-imine) (HCOO-(CH2-NH)n-H) are formed. In particular, HMT is known to decompose under VUV irradiation, producing HCN molecules and CN radicals, and there- fore is a possible extended source of the observed HCN and CN in comets (Bernstein et al. 1995; Vinogradoff et al. 2011). reverse temperature the higher At 4 CN− → NH3 + HCN/HNC can occur, reaction NH+ releasing NH3 and HCN (and possibly HNC) into the gas phase. This type of isomerisation has previously been shown for the HNCO/HOCN system in water ice (Theule et al. 2011b). The formation of NH+ 4 CN− allows HCN to be stored within a less volatile species and made available to a high temperature chemistry, when the species fragments and eventually desorbs. As a follow up to studying its formation kinetics, it would be interesting to study the NH+ 4 CN− photodissociation under a VUV field. Does this sort of irradiation release the simple molecules c(cid:13) 2002 RAS, MNRAS 000, 1–12 Thermal reactivity of HCN and NH3 11 REFERENCES Acharyya, K., Fuchs, G. W., Fraser, H. J., van Dishoeck, E. F., & Linnartz, H. 2007, A&A, 466, 1005 Ag´undez, M., Cernicharo, J., Gu´elin, M., et al. 2010, A&A, 517, L2 Aikawa, Y., Kamuro, D., Sakon, I., et al. 2012, A&A, 538, A57 Bertin, M., Martin, I., Duvernay, F., et al. 2009, Phys. Chem. Chem. Phys., 11, 1838 Bernstein, M. P., Sandford, S. A., Allamandola, L. J., Chang, S., & Scharberg, M. A. 1995, ApJ, 454, 327 Bernstein, M. P., Sandford, S. A., & Allamandola, L. J. 1997, ApJ, 476, 932 Bernstein, M.P., Dworkin, J.P., Sandford, S.A., Cooper, G.W., & Allamandola, L. J. 2002, Nature, 416, 401 Bieging, J. H., Chapman, B., & Welch, W. J. 1984, ApJ, 285, 656 Bockelee-Morvan, D., Crovisier, J., Despois, D., et al. 1987, A&A, 180, 253 Boger, G. I., & Sternberg, A. 2005, ApJ, 632, 302 Boogert, A. C. A., Pontoppidan, K. M., Knez, C., et al. 2008, ApJ, 678, 985 Boogert, A. C. A., Huard, T. L., Cook, A. M., et al. 2011, ApJ, 729, 92 Boonman, A. M. S., Stark, R., van der Tak, F. F. S., et al. 2001, ApJL, 553, L63 Borget, F., Danger, G., Duvernay, F., et al. 2012, A&A, 541, A114 Bossa, J. B., Theule, P., Duvernay, F., Borget, F., & Chiavassa, T. 2008, A&A, 492, 719 Bossa, J. B., Theule, P., Duvernay, F., & Chiavassa, T. 2009, ApJ, 707, 1524 Bottinelli, S., Boogert, A. C. A., Bouwman, J., et al. 2010, ApJ, 718, 1100 Burgdorf, M., Cruikshank, D. P., Dalle Ore, C. M. et al. 2010, ApJL, 718, L53 Burke, D. J., & Brown, W. A. 2010, PCCP, 12, 5947 Chapillon, E., Guilloteau, S., Dutrey, A., Pi´etu, V., & Gu´elin, M. 2012, A&A, 537, A60 Clark, F. O., Buhl, D., & Snyder, L. E. 1974, ApJ, 190, 545 Clutter, D. R., & Thompson, W. E. 1969, J. Chem. Phys., 51, 153 Colin-Garc´ıa, M., Negr´on-Mendoza, A., & Ramos-Bernal, S. 2009, Bioastronomy 2007: Molecules, Microbes and Extrater- restrial Life, 420, 175 Collings, M. P., Dever, J. W., Fraser, H. J., & McCoustra, M. R. S. 2003, Ap&SS, 285, 633 Collings, M. P., Anderson, M. A., Chen, R., et al. 2004, MNRAS, 354, 1133 Danger, G., Borget, F., Chomat, M., et al. 2011, A&A, 535, A47 Danger, G., Duvernay, F., Theul´e, P., Borget, F., & Chiavassa, T. 2012, ApJ, 756, 11 Dartois, E. 2005, Space Sci. Rev., 119, 293 Demyk, K., Dartois, E., d'Hendecourt, L., et al. 1998, A&A, 339, 553 Elsila, J. E., Glavin, D. P., & Dworkin, J. P. 2009, Meteoritics and Planetary Science, 44, 1323 Ferris, J. P., & Hagan, W. J. 1984, Tetrahedron, 40, 1093 Fray, N., B´enilan, Y., Cottin, H., Gazeau, M.-C., & Crovisier, J. 2005, Planet. Space Sci., 53, 1243 Fuchs, G. W., Acharyya, K., Bisschop, S. E., et al. 2006, Faraday Discussions, 133, 331 G´alvez, O., Mat´e, B., Herrero, V. J., & Escribano, R. 2010, ApJ, 724, 539 Gerakines, P. A., Moore, M. H., & Hudson, R. L. 2004, Icarus, Figure 9. A schematic of the simplified chemical network surrounding the NH3 + HCN → NH+ 4 CN− thermal reaction. The final products are a) hydroxyacetonitrile (HOCH2CN), b) aminoacetonitrile (NH2CH2CN), c) poly(methylene-imine) (PMI, R-(CH2-NH)n-H, where R = HCOO or CN), and d) hexamethylenetetramine (HMT, C6H12N4). HCN and NH3, as upon heating the salt, or rather yield a more complex chemistry? Moreover, the HCN/HNC ratio produced by the reverse reaction could be measured in the gas phase using gas phase spectroscopy. This type of experiment could aid interpreta- tion of gas phase HCN/HNC ratios in interstellar environments. The inventory of the thermal reactions involving the most abundant solid state species must be continued, and the kinetics of the rate limiting reactions should be investigated, in order to achieve a complete network of interstellar grain reactions. The competition of this series of coupled reactions must be quantified as a function of temperature and of initial abundance ratios in order to understand interstellar ice chemistry. 5 CONCLUSION In this work, we have shown experimentally that the purely ther- 4 CN− in inter- mal reaction between HCN and NH3 produces NH+ stellar ice analogues at low temperature. NH+ 4 CN− has a charac- teristic CN− absorption band at 2092 cm−1 (4.78 µm) that can be slightly redshifted by the ice environment. However, the environ- mental effect is not sufficient that the CN− band can match the 2041 cm−1 (4.90 µm) band observed in the IR spectra of interstel- lar ices. In addition, at low temperatures, the contributions of HCN and CN− to the observed CN stretching mode absorption are diffi- cult to differentiate. We have estimated the CN− ion band strength to be ACN− = 1.8 ± 1.5 × 10−17 cm molec−1 at 20 K and at 140 K; when coupled with its expected low abundance in interstellar ices, it is unlikely that solid phase CN− could be detected by current IR telescopes. We have measured the kinetics of the pure HCN and NH3 reaction with an excess of NH3, and we have derived the tem- perature dependence for the reaction rate constant to be k(T ) = 0.016+0.010 ). When the reactants are diluted in water ice, the reaction is slowed down. We have also measured a desorption energy of Edes,NH4CN = 38.0 ± 1.4 kJ mol−1, assuming a pre-exponential factor of 1028 molecules cm−2 s−1. −0.006 s−1 exp( −2.7±0.4 kJ mol−1 RT ACKNOWLEDGMENTS This work has been funded by the French national program Physique Chimie du Milieu Interstellaire (PCMI) and the Centre National d'Etudes Spatiales (CNES). J.A.N. is a Royal Commis- sion for the Exhibition of 1851 Research Fellow. The authors would like to thank Prof. H. Cottin for the HCN synthesis protocol. c(cid:13) 2002 RAS, MNRAS 000, 1–12 12 J. A. Noble et al. 170, 202 Gibb, E. L., Whittet, D. C. B., Schutte, W. A., et al. 2000, ApJ, 536, 347 Gibb, E. L., & Whittet, D. C. B. 2002, ApJL, 566, L113 Gibb, E. L., Whittet, D. C. B., Boogert, A. C. A., & Tielens, A. G. G. M. 2004, ApJS, 151, 35 Glavin, D. P., Dworkin, J. P., & Sandford, S. A. 2008, Meteoritics and Planetary Science, 43, 399 Hasegawa, T. I., Herbst, E., & Leung, C. M. 1992, ApJS, 82, 167 d'Hendecourt, L. B., & Allamandola, L. J. 1986, A&AS, 64, 453 Herbst, E., Terzieva, R., & Talbi, D. 2000, MNRAS, 311, 869 Hirota, T., Yamamoto, S., Mikami, H., & Ohishi, M. 1998, ApJ, 503, 717 Huebner, W. F., Buhl, D., & Snyder, L. E. 1974, Icarus, 23, 580 Ip, W.-H., Balsiger, H., Geiss, J., et al. 1990, Annales Geophysi- cae, 8, 319 Jamieson, C. S., Chang, A. H. H., & Kaiser, R. I. 2009, Advances in Space Research, 43, 1446 Keane, J. V., Tielens, A. G. G. M., Boogert, A. C. A., Schutte, W. A., & Whittet, D. C. B. 2001, A&A, 376, 254 King, D. A. 1975, Surf. Sci., 47, 384 Lacy, J. H., Faraji, H., Sandford, S. A., & Allamandola, L. J. 1998, ApJL, 501, L105 Lahuis, F., & van Dishoeck, E. F. 2000, A&A, 355, 699 Mispelaer, F., Theule, P., Duvernay, F., Roubin, P., & Chiavassa, T. 2012, A&A, 540, A40 Moreno, R., Lellouch, E., Lara, L. M., et al. 2011, A&A, 536, L12 Noble, J. A., Congiu, E., Dulieu, F., & Fraser, H. J. 2012a, MN- RAS, 421, 768 Noble, J. A., Theule, P., Mispelaer, F., et al. 2012b, A&A, 543, A5 Padovani, M., Walmsley, C. M., Tafalla, M., Hily-Blant, P., & Pineau Des Forets, G. 2011, A&A, 534, A77 Paganini, L., Villanueva, G. L., Lara, L. M., et al. 2010, ApJ, 715, 1258 Pizzarello, S. 2012, ApJL, 754, L27 Raunier, S., Chiavassa, T., Marinelli, F., Allouche, A., & Aycard, J.-P. 2003a, Chem. Phys. Lett., 368, 594 Raunier, S., Chiavassa, T., Allouche, A., Marinelli, F., & Aycard, J.-P. 2003b, Chem. Phys., 288, 197 Raunier, S., Chiavassa, T., Marinelli, F., & Aycard, J.-P. 2004, Chem. Phys., 302, 259 Rosenqvist, J., Lellouch, E., Romani, P. N., Paubert, G., & Encre- naz, T. 1992, ApJL, 392, L99 Sandford, S. A., & Allamandola, L. J. 1993, ApJ, 417, 815 Schreyer, K., Helmich, F. P., van Dishoeck, E. F., & Henning, T. 1997, A&A, 326, 347 Schutte, W. A., Boogert, A. C. A., Tielens, A. G. G. M., et al. 1999, A&A, 343, 966 Schutte, W. A., & Khanna, R. K. 2003, A&A, 398, 1049 Snyder, L. E., & Buhl, D. 1971, ApJL, 163, L47 Soifer, B. T., Puetter, R. C., Russell, R. W. et al. 1979, ApJL, 232, L53 Theule, P., Borget, F., Mispelaer, F. et al. 2011a, A&A, 534, A64 Theule, P., Duvernay, F., Ilmane, A., Coussan, S., & Chiavassa, T. 2011b, A&A, 530, A96 Tokunaga, A. T., Beck, S. C., Geballe, T. R., Lacy, J. H., & Ser- abyn, G. 1981, BAAS, 13, 701 van Broekhuizen, F. A., Keane, J. V., & Schutte, W. A. 2004, A&A, 415, 425 van Broekhuizen, F. A., Pontoppidan, K. M., Fraser, H. J., & van Dishoeck, E. F. 2005, A&A, 441, 249 Vasyunina, T., Linz, H., Henning, T., et al. 2011, A&A, 527, A88 Vinogradoff, V., Duvernay, F., Danger, G., Theule, P., & Chi- avassa, T. 2011, A&A, 530, A128 Wagner, E. L., & Hornig, D. F. 1950, J. Chem. Phys., 18, 296 Woodney, L. M., A'Hearn, M. F., Schleicher, D. G., et al. 2002, Icarus, 157, 193 Woon, D. E. 2001, Icarus, 149, 277 This paper has been typeset from a TEX/ LATEX file prepared by the author. c(cid:13) 2002 RAS, MNRAS 000, 1–12
1510.02776
1
1510
2015-10-09T19:19:58
The Laplace resonance in the Kepler-60 system
[ "astro-ph.EP" ]
We investigate the dynamical stability of the Kepler-60 planetary system with three super-Earths. We first determine their orbital elements and masses by Transit Timing Variation (TTV) data spanning quarters Q1-Q16 of the KEPLER mission. The system is dynamically active but the TTV data constrain masses to ~4 Earth masses and orbits in safely wide stable zones. The observations prefer two types of solutions. The true three-body Laplace MMR exhibits the critical angle librating around 45 degrees and aligned apsides of the inner and outer pair of planets. In the Laplace MMR formed through a chain of two-planet 5:4 and 4:3 MMRs, all critical angles librate with small amplitudes of ~30 degrees and apsidal lines in planet's pairs are anti-aligned. The system is simultaneously locked in a three-body MMR with librations amplitude of ~10 degrees. The true Laplace MMR can evolve towards a chain of two-body MMRs in the presence of planetary migration. Therefore the three-body MMR formed in this way seems to be more likely state of the system. However, the true three-body MMR cannot be disregarded a priori and it remains a puzzling configuration that may challenge the planet formation theory.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 5 (2012) Printed 5 October 2018 (MN LATEX style file v2.2) The Laplace resonance in the Kepler-60 planetary system K. Go´zdziewski1, C. Migaszewski1,2, F. Panichi2, and E. Szuszkiewicz2 1Centre for Astronomy, Faculty of Physics, Astronomy and Informatics, Nicolaus Copernicus University, Grudziadzka 5, 87-100 Toru´n, Poland 2Institute of Physics and CASA*, Faculty of Mathematics and Physics, University of Szczecin, Wielkopolska 15, 70-451 Szczecin, Poland 5 October 2018 ABSTRACT We investigate the dynamical stability of the Kepler-60 planetary system with three super- Earths. We determine their orbital elements and masses by Transit Timing Variation (TTV) data spanning quarters Q1-Q16 of the KEPLER mission. The system is dynamically active but the TTV data constrain masses to ∼ 4 m⊕ and orbits in safely wide stable zones. The observations prefer two types of solutions. The true three-body Laplace MMR exhibits the critical angle librating around (cid:39) 45◦ and aligned apsides of the inner and outer pair of planets. In the Laplace MMR formed through a chain of two-planet 5:4 and 4:3 MMRs, all critical angles librate with small amplitudes ∼ 30◦ and apsidal lines in planet's pairs are anti-aligned. The system is simultaneously locked in a three-body MMR with librations amplitude (cid:39) 10o. The true Laplace MMR can evolve towards a chain of two-body MMRs in the presence of planetary migration. Therefore the three-body MMR formed in this way seems to be more likely state of the system. However, the true three-body MMR cannot be disregarded a priori and it remains a puzzling configuration that may challenge the planet formation theory. Key words: extrasolar planets -- N-body problem -- photometry -- star: Kepler-60 1 INTRODUCTION 2 THREE-BODY MEAN MOTION RESONANCE The KEPLER mission has lead to the discovery of a few hundred multiple planetary systems with super-Earth planets. Such systems may be involved in low-order mean motion resonances (MMRs) or close to the MMRs (e.g. Lee et al. 2013). The Transit Timing Variation (Agol et al. 2005) or the photodynamical N-body method (Carter et al. 2011) are common approaches making it possible to determine planetary masses and orbital architectures of their parent systems. A large sample of KEPLER multiple systems has been an- alyzed in (Rowe et al. 2015; Mullally et al. 2015). They provided a rich catalogue of basic photometric models of such systems, in- cluding TTV measurements for a number of them. In this sample, the Kepler-60 system of three super-Earth plan- ets (Steffen et al. 2012) with yet unconstrained masses reveals the mean orbital period ratios ∼ 1.250 and ∼ 1.334 very close to relatively prime integers. This is suggestive for a chain of two- body MMRs. One pair of planets is near the 5:4 MMR and the second one is near the 4:3 MMR. Moreover, the mean-motions of Kepler-60b (n1), Kepler-60c (n2) and Kepler-60d (n3) satisfy n1 − 2n2 + n3 ≈ −0.002◦d−1. This relation may be understood as the generalized Laplace resonance (Papaloizou 2015). Our preliminary dynamical analysis has revealed that the mu- tual interactions between the Kepler-60 planets are strong and the dynamical constraints may be decisive for maintaining the long- term stability of this system. Therefore, besides a characterization of the putative resonance, and determining the masses, we aim to perform a comprehensive dynamical study of the Kepler-60 on the basis of the TTV measurements in (Rowe et al. 2015). c(cid:13) 2012 RAS Three-body MMR's may be considered as the most important case of multiple MMRs, after two-body MMRs, which may actively shape short-term and long-term dynamics of multi-planet config- urations. The overlap of two-body and three-body MMRs has been identified as a source of deterministic chaos in the Outer Solar sys- tem and in the asteroid belt (e.g. Murray & Holman 1999; Guzzo 2006; Smirnov & Shevchenko 2013). The best-studied and known example of the three-body MMR is the Laplace resonance of the Galilean moons. The mean-motions of Io (nIo), Europa (nE) and Ganymede (nG) satisfy nIo − 2nE = nE − 2nG ≈ 0.74◦d−1 that im- plies nIo − 3nE + 2nG (cid:39) 0◦d−1 (e.g. Sinclair 1975). Following Papaloizou (2015), we consider a subset of five types of the Laplace MMRs. In a coplanar system, a combina- tion of the first-order MMRs' librating -- nonlibrating critical angles φ1, j = pλ1 − (p + 1)λ2 + ϖ j, φ2, j = qλ2 − (q + 1)λ3 + ϖ j+1, for relatively prime integers p,q > 0, where j = 1,2, λi, ϖi are the mean longitudes and pericenter longitudes (i = 1,2,3) for subse- quent planets, results in a chain of two-body MMRs of the first order. Librations of the critical angles in this chain naturally imply librations of φL = pλ1 − (p + q + 1)λ2 + (q + 1)λ3 (type I). A sce- nario when only φL librates but all two-body MMR critical angles circulate will be called the "true" or "pure" three-body MMR (type II). We do not consider here two more types related to three or two out of four two-planet MMR critical angles librating. The first extrasolar system exhibiting the three-body Laplace MMR has been discovered around Gliese 876 (Marcy et al. 2001; Rivera et al. 2010). In this case two-body MMR critical angles of K. Go´zdziewski, C. Migaszewski, F. Panichi & E. Szuszkiewicz 2 the first pair of planets librate around 0◦, one ciritcal angle librates around 0◦ for the outer pair and φL librates around 0◦ with ampli- tude ∼ 40◦(Mart´ı et al. 2013). Another intriguing example of two instances of the Laplace resonance in one planetary system is likely the HR 8799 system of four massive ∼ 10 mJup planets in wide or- bits ∼ 100 au (Marois et al. 2010; Go´zdziewski & Migaszewski 2014). Given a putative Laplace resonance in the Kepler-60 sys- tem, this MMR may form in different environments, spanning wide mass-ranges, from tiny moons of Pluto (Showalter & Hamilton 2015), the Mercury-like satellites of Jupiter, super-Earths and Jo- vian planets, to (and perhaps beyond) the brown dwarf mass limit. a posteriori Lomb-Scargle periodograms of the residuals of best- fitting models did not show significant frequencies. Because the values of L are non-intuitive for comparing so- lutions, a rescaled value logL = log0.2420− logL/N (in days) is more suitable to express the quality of fits. Since the best-fitting models should provide χ2/N ∼ 1, then L ∼ (cid:104)σ(cid:105) measures a scat- ter of measurements around the best-fitting models, similar with an r.m.s -- smaller L means better fit (e.g., Baluev 2009). To perform the MCMC analysis we used the affine-invariant ensemble sampler (Goodman & Weare 2010; Foreman-Mackey et al. 2013). As a different method, we applied a few kinds of ge- netic and evolutionary algorithms (GEA from hereafter, Charbon- neau 1995; Ruci´nski et al. 2010) to independently maximize the logL function. We set the same parameter bounds in the GEA runs as in the MCMC experiments. 3.1 The best-fitting configurations We first performed an extensive search with the GEA, collecting ∼ 106 solutions in each multi-CPU run. We found that the best- fitting models with L (cid:39) 0.029 d have well determined minima for the orbital periods Pi and transit epochs Ti, similar with the er- ror term σ f ∼ 0.02 days. As expected, eccentricities are not con- strained and we found equally good models in whole (xi,yi)-planes. The results are consistent with the MCMC experiments illus- trated in a figure in supplementary material. It shows one -- and two -- dimensional projections of the posterior probability distribution. The posterior is multi-modal and complex with a dominant peak localized roughly around masses of ∼ 4 m⊕ and a weaker peak around ∼ 10 m⊕ for all planets. There are also two relatively weak, quasi-symmetric peaks for all xi ∼ ±0.25 and yi ∼ 0.25. Strong linear correlation between all pairs of (xi,x j) and (yi,y j), i (cid:54)= j is present. Given close commensurabilities of the orbital periods, it is an additional indication of the MMRs in the system. The linear cor- relations mean a tight alignment of apsidal lines which is a typical feature of the low-order MMRs. We searched for other signatures of the MMRs by computing amplitudes of the critical angles of the two-body and three-body MMRs in the GEA runs, by integrating numerically the N-body equations of motion for all models with L < 0.050 d. We found that basically all the solutions exhibit librations of φL = λ1 − 2λ2 + λ3 around (cid:39) 45◦ with amplitudes as small as 10◦. Most of these mod- els do not show librations of any of the two-body MMRs, hence they represent the true three-body MMRs. For small eccentrici- ties we found similar small L ∼ 0.029 d, low-amplitude φL solu- tions with all two-body MMR critical angles librating. That means a chain of two-body MMRs. Elements of two representative low- eccentricity models with small L (cid:39) 0.029 d are given in Table 1 and illustrated in Fig. 1. These qualitatively different modes of the Laplace MMR (marked in red and blue, respectively) can be hardly distinguished by the fit quality. We note that within uncertainties these solutions have the same Pi, Ti and mi. Regular evolutions of the critical angles as well as apsidal angles ∆ϖi, j = ϖi − ϖ j for 2 Myrs are shown in Figs. 2 and 3, respectively. 3.2 The long-term stability of resonant configurations To investigate the dynamical stability of the system, we applied the Mean Exponential Growth factor of Nearby Orbits (MEGNO, (cid:104)Y(cid:105) Cincotta et al. 2003) which is a CPU-efficient variant of the Maximal Lyapunov Exponent (MLE) useful for analysing stabil- c(cid:13) 2012 RAS, MNRAS 000, 1 -- 5 3 THE TTV DATA MODEL AND OPTIMIZATION The TTV dataset D for Kepler-60 consists of 419 measurements spanning quarters Q1-Q16, with a reported mean uncertainty (cid:104)σ(cid:105) ∼ 0.024 days (Rowe et al. 2015). Due to this large scatter which com- pares to the magnitude of the TTV data, we introduce a simplified orbital architecture. We assume that the system is coplanar or close to coplanar and then the inclinations Ii = 90◦ and nodal longitudes Ωi = 0◦, i = 1,2,3. Orbits in compact KEPLER systems are usually quasi-circular, hence to get rid of weakly constrained longitudes of pericenter ϖi when eccentricities ei ∼ 0, we introduce osculating, astrocentric elements {Pi,xi,yi,Ti} instead of(cid:8)ai,ei,ωi,Mi (cid:9): (cid:17) (cid:115) (cid:16) a3 i Pi = 2π k2 (m0 + mi) , Ti = t0 + Pi 2π i − Mi M (t) , and xi = ei cosϖi, yi = ei sinϖi where k is the Gauss constant, M (t) is the mean anomaly at the epoch of the first transit Ti, and Mi, Pi, ai are for the mean anomaly, the orbital period and semi-major-axis, at the osculating epoch t0 for each planet, respectively. We computed the transits moments with a code by Deck et al. (2014) and with our own code for an independent check. i The TTV technique is plagued by non-unique and/or un- constrained solutions. Therefore, we applied two different meth- ods of quasi-global optimization in parallel, to explore the space of free parameters θθθ in the model including geometric elements (Pi,Ti,xi,yi) and masses mi, for i = 1,2,3 marking subsequent plan- ets, as well as the measurements uncertainty correction term σ f (see below). The Markov Chain Monte Carlo (MCMC) technique is widely used by the photometric community to determine the posterior probability distribution P (θθθD) of model parameters θθθ, given the data set D: P (θθθD) ∝ P (θθθ)P (Dθθθ), where P (θθθ) is the prior, and the sampling data distribution P (Dθθθ) ≡ logL(θθθ,D). We choose all priors as flat (or uniform improper) by placing limits on model parameters, i.e., Pi > 0 d, Ti > 0 d, xi,yi ∈ (−0.5,0.5), mi ∈ [0.1,30] m⊕, σ f > 0 d (i = 1,2,3). ν ∼ 1.7 and a large scatter of residuals suggestive for underestimated uncertainties, we opti- mized the maximum likelihood function L: Since our preliminary fits revealed χ2 logL = − 1 2 ∑ i,t (O-C)2 i,t σ2 i,t − 1 2 ∑ i,t logσ2 i,t − 1 2 N log2π, (1) where (O-C)i,t is the (O-C) deviation of the observed t-th transit moment of an i-th planet from its N-body ephemeris, and the TTV uncertainty σ2 f with σ f parameter scaling the raw un- certainty σi,t and N is the total number of TTV observations. We as- sume that the uncertainties are Gaussian and independent. Indeed, i,t → σ2 i,t + σ2 The Laplace resonance in the Kepler-60 system 3 Table 1. Orbital parameters of representative models of the Kepler-60 sys- tem for the true three-body Laplace resonance (Fit I) and for a chain of two- body MMRs (Fit II). The osculating epoch is BKJD+65.0. Both systems are coplanar with the inclination i = 90◦ and nodal longitudes Ω = 0◦. Mass of the star is 1.105m(cid:12) (Rowe et al. 2015). Elements xi,yi and ei,ϖi are weakly limited with the TTV, but are constrained by the dynamical stability. planet Kepler-60 b Kepler-60 c Kepler-60 d Fit I (three-body true Laplace MMR) mp [m⊕] P[d] ecosϖ esinϖ T [d] a [au] e ω[deg] M [deg] σ f [d] L [d] 4.0±0.7 7.1320±0.0005 0.0152 -0.0354 69.032±0.009 0.07497 0.0386 -66.74 134.92 5.0±1.0 8.9179±0.0006 0.0512 -0.0244 73.075±0.009 0.08701 0.0567 -25.47 -24.75 0.018 ± 0.002 0.029 3.6±1.1 11.903±0.001 0.0077 -0.0204 66.267±0.012 0.10548 0.0218 -69.44 301.99 Fit II (two-body MMR chain with Laplace MMR) mp [m⊕] P[d] ecosϖ esinϖ T [d] a [au] e ω[deg] M [deg] σ f [d] L [d] 4.1±0.7 7.1320 ±0.0005 -0.0108 0.0008 69.035±0.009 0.07497 0.0108 175.56 -110.47 4.8±1.0 8.9177±0.0006 0.0282 0.0073 73.075± 0.009 0.08701 0.0291 14.61 -67.35 0.018 ± 0.002 0.029 3.8±1.1 11.903±0.001 -0.0129 0.0080 66.273±0.012 0.10548 0.0151 148.19 81.84 ity of planetary systems with strongly interacting planets (e.g. Go´zdziewski & Migaszewski 2014). We computed 2-dim dynamical maps in the neighborhood of the best-fitting solutions to show the MMRs structure of the phase- space. The top left-hand panel of Fig. 4 shows the dynamical map in the (a1,a2) -- plane for the best-fitting true Laplace MMR (Tab. 1). (A map for the second, MMR chain mode looks similar). All other orbital elements are kept at their best-fitting values. For each initial condition at the grid, the equations of motion where integrated up to 32,000yr. This corresponds to ∼ 106 × P3, which is sufficient to detect short-term chaotic motions for the MMRs instability time- scale (e.g. Go´zdziewski & Migaszewski 2014). As the dynamical maps show, the observational uncertainties of a1 and a2 (cid:39) 10−5 au are much smaller than the width of stable islands. For the true Laplace MMR, we found narrower islands of stable motions even for e ∼ 0.22 and larger (the bottom-right panel in Fig. 4) consistent with the posterior (online material). However, such large eccentricities would be difficult to explain in the frame- work of the present state of the planet formation theory (Sect. 4). We note that stable solutions found for masses ∼ 4 m⊕ concur with an empirical mass-radius relation mp/m⊕ = (Rp/R⊕)2.06 by Fabrycky et al. (2014). It provides m1 (cid:39) 4.0m⊕, m2 (cid:39) 4.7m⊕ and m3 (cid:39) 3.2m⊕. For second type of the best-fitting models the masses ∼ 10 m⊕ and eccentricities ∼ 0.2 -- we did not find long-term sta- ble solutions. Yet these models provide L ∼ 0.030 d, which means worse solutions than derived for the smaller mass range. A dynamical character of the system is illustrated in the (a1,a2)-map in the bottom-left panel of Fig. 4 which has been Figure 1. Synthetic curves of best-fitting low-eccentricity models (Tab. 1): Fit I (red circles) and Fit II (blue circles) over-plotted on the TTV data. Figure 2. Evolution of the Laplace MMR critical angle φL and apsidal angles ∆ϖi, j for Fit I (Tab. 1). All two-body MMR critical angles rotate (not shown). c(cid:13) 2012 RAS, MNRAS 000, 1 -- 5 -200-150-100-50 0 50 100 150 0 200 400 600 800 1000 1200 1400 1600TTV (O-C) [min]transit epoch [+BKJD 2,454,833]Kepler-60b-150-100-50 0 50 100 150 0 200 400 600 800 1000 1200 1400 1600TTV (O-C) [min]transit epoch [+BKJD 2,454,833]Kepler-60c-200-150-100-50 0 50 100 150 200 0 200 400 600 800 1000 1200 1400 1600TTV (O-C) [min]transit epoch [+BKJD 2,454,833]Kepler-60d 4 K. Go´zdziewski, C. Migaszewski, F. Panichi & E. Szuszkiewicz 4 THE TRUE THREE-BODY MMR VIA MIGRATION? A formation of the three-body MMRs has been recently studied by e.g. Libert & Tsiganis (2011); Quillen (2011); Quillen & French (2014); Batygin et al. (2015). It is widely accepted that short-period planets form in protoplanetary discs at distances wider than the ones observed now and then migrate inwards due to planet-disc interactions. It is known that convergent migration of a few planets leads to formation of chains of MMRs (e.g., Papaloizou & Terquem 2010). From this perspective the true three-body resonance seems to be unexpected. A lack of two-body MMRs is not the only sur- prising feature of this configuration. Also mutual orientations of the apsidal lines (apsides of subsequent orbits are aligned) differ from a typical outcome of the migration of three-planet systems, i.e., ap- sides anti-aligned in a low eccentricity regime (Papaloizou 2015). Although the orientations may be different if merging and scatter- ing of initially larger number of planets (or protoplanetary cores) are considered (Terquem & Papaloizou 2007). Explaining how the true three-body MMR could be formed via migration is well beyond the scope of this Letter, as many different disc parameters and initial orbits should be tested. Instead, we tried to verify if the conditio sine qua non of such scenario is fulfilled. We tested if chosen representative configurations are stable against migration, i.e., whether or not after adding forces mimicking the migration and circularization to the equations of motion (Moore et al. 2013) the system stays in the true three-body MMR or leaves it (possibly moving towards a chain of two-body MMRs). We checked that regardless of whether the migration is conver- gent or divergent the system leaves the resonance. The parameter K which is the ratio of the migration and circularization time-scales was being changed in a range of [0.1,100]. For K (cid:38) 1 the systems tend towards a chain of two-body MMRs with anti-aligned apsides of subsequent orbits. Naturally, for the divergent migration the pe- riod ratios increase and the systems leave the chain. For K (cid:46) 1 the systems self-disrupt. Those tests suggest that the true three-body MMR will be a challenge for the resonances formation theory. Figure 3. Evolution of the critical angles for Fit II (Tab. 1) correspond- ing to a chain of two-body MMRs. Apsidal angles ∆ϖ1,2 and ∆ϖ2,3 librate around 180◦ with semi-amplitudes ∼ 30◦ (not shown here). 5 CONCLUSIONS computed for only 80 years. Already for this time-span a strongly chaotic structure appears, revealing dominant, wide strips of two- body and three-body MMRs that intersect in a central island of sta- ble configurations that corresponds to the Laplace MMR. Such a structure may be interpreted as the Arnold web emerging due to MMRs overlap. A build-up of this structure is illustrated in the (a1,a2)-maps (the left column in Fig. 4). After a sufficient satu- ration time (∼ 4,000 yr and longer), only narrow stable regions re- main in the top left-hand map that indicates the Chirikov regime of the chaotic dynamics (Froeschl´e et al. 2000; Guzzo 2006). In that case a chaotic diffusion may lead to a random-walk "wandering" of solutions along the resonances. This effect leads to relatively fast and significant changes of the dynamical actions (semi-major axes). Chaotic models are unstable in the Lagrangian, geometrical sense, particularly in regions of moderate and large eccentricities. The Arnold web is one more feature of a strongly resonant system. Rigorously stable best-fitting solutions found in this paper are pe- culiar, since as we confirm here, recent studies (Wang et al. 2012; Showalter & Hamilton 2015) indicate that three-body resonances exhibit very short Lyapunov times that may cause a rapid instabil- ity. The orbital periods of the Kepler-60 planetary system exhibit close commensurabilities which may be interpreted as a chain of two- body, first-order MMRs or the true three-body MMR with none of the two-body MMRs critical angles librating. We found a strong observational indication that the zero-th order three-body MMR 1 : −2 : 1 is present. Regardless of its type, it could be considered as the generalized Laplace resonance (Papaloizou 2015). It is charac- terized by ∼ 10◦ -- amplitude libration around ∼ 45◦ libration center. The TTV series imply a very complex structure of the phase space of the system. Strong mutual perturbations between ∼ 4m⊕ super-Earths lead to the Chirikov regime of the dynamics which is governed by chaotic diffusion due to the overlap of two- and three- body MMRs. Long-term stable orbital configurations are confined to isolated islands associated with the MMRs. A past putative mi- gration probably rules out the true three-body MMR since the res- onance is not robust against the migration. The most likely state of Kepler-60 system is a chain of two-body MMRs with all criti- cal angles librating with small amplitudes. The lightcurve has low signal-to-noise ratio, similar with many other multiple-systems in the KEPLER sample. Therefore constraining a particular type of the Laplace MMR in the Kepler-60 system with the present data is not likely. It remains an open problem whose solution may shed more light on the formation of this intriguing system. c(cid:13) 2012 RAS, MNRAS 000, 1 -- 5 The Laplace resonance in the Kepler-60 system 5 Figure 4. Dynamical maps for the best-fitting three-body MMRs models in Tab. 1. The MEGNO (cid:104)Y(cid:105) ∼ 2 indicates a regular (long-term stable) solution marked with black/dark blue colour, (cid:104)Y(cid:105) much larger than 2, up to (cid:38) 5 indicates a chaotic solution (yellow). The integration time of each initial condition is 32,000 yr (∼ 106 × P3) besides the bottom-left panel where the integration time is 80 yr. The asterisk symbol means the position of the nominal model. The top-row: dynamical maps for Fit I in Tab. 1. Subsequent maps are for the (a1,a2) -- , (a1,e1) -- and (ϖ2,ϖ3) -- planes. The bottom-row: (a1,a2) -- map for Fit I with reduced integration time, (a1,e1) -- plane for Fit II, and (a1,e1) -- map for the true Laplace resonance in moderate -- eccentricity region. See the text for details. 6 ACKNOWLEDGEMENTS We thank the anonymous reviewer for comments that improved this paper. This work has been supported by Polish National Science Centre MAESTRO grant DEC-2012/06/A/ST9/00276. K.G. thanks the Pozna´n Supercomputer and Network Centre (PCSS, Poland) for computing resources (grant No. 195) and a generous support. REFERENCES Agol E., Steffen J., Sari R., Clarkson W., 2005, MNRAS, 359, 567 Baluev R. V., 2009, MNRAS, 393, 969 Batygin K., Deck K. M., Holman M. J., 2015, AJ, 149, 167 Carter J. A., Fabrycky D. C., Ragozzine D., Holman M. J., Quinn S. N., et al. 2011, Science, 331, 562 Charbonneau P., 1995, ApJS, 101, 309 Cincotta P. M., Giordano C. M., Sim´o C., 2003, Physica D Non- linear Phenomena, 182, 151 Deck K. M., Agol E., Holman M. J., Nesvorn´y D., 2014, ApJ, 787, 132 Fabrycky D. C., Lissauer J. J., Ragozzine D., Rowe J. F., Steffen J. H., Agol E., Barclay T., et al. 2014, ApJ, 790, 146 Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013, PASP, 125, 306 Froeschl´e C., Guzzo M., Lega E., 2000, Science, 289, 2108 Goodman J., Weare J., 2010, Comm. Apl. Math and Comp. Sci., 1, 65 Go´zdziewski K., Migaszewski C., 2014, MNRAS, 440, 3140 Guzzo M., 2006, Icarus, 181, 475 Lee M. H., Fabrycky D., Lin D. N. C., 2013, ApJ, 774, 52 Libert A.-S., Tsiganis K., 2011, Celestial Mechanics and Dynam- ical Astronomy, 111, 201 c(cid:13) 2012 RAS, MNRAS 000, 1 -- 5 Marcy G. W., Butler R. P., Fischer D., Vogt S. S., Lissauer J. J., Rivera E. J., 2001, ApJ, 556, 296 Marois C., Zuckerman B., Konopacky Q. M., Macintosh B., Bar- man T., 2010, Nature, 468, 1080 Mart´ı J. G., Giuppone C. A., Beaug´e C., 2013, MNRAS, 433, 928 Moore A., Hasan I., Quillen A. C., 2013, MNRAS, 432, 1196 Mullally et al. 2015, ApJS, 217, 31 Murray N., Holman M., 1999, Science, 283, 1877 Papaloizou J. C. B., 2015, International Journal of Astrobiology, 14, 291 Papaloizou J. C. B., Terquem C., 2010, MNRAS, 405, 573 Quillen A. C., 2011, MNRAS, 418, 1043 Quillen A. C., French R. S., 2014, MNRAS, 445, 3959 Rivera E. J., Laughlin G., Butler R. P., Vogt S. S., Haghighipour N., Meschiari S., 2010, ApJ, 719, 890 Rowe J. F., Coughlin J. L., Antoci V., Barclay T., Batalha N. M., Borucki W. J., Burke C. J., et al. 2015, ApJS, 217, 16 Ruci´nski M., Izzo D., Biscani F., 2010, Parallel Computing, 10, 555 Showalter M. R., Hamilton D. P., 2015, Nature, 522, 45 Sinclair A. T., 1975, MNRAS, 171, 59 Smirnov E. A., Shevchenko I. I., 2013, Icarus, 222, 220 Steffen J. H., Fabrycky D. C., Ford E. B., Carter J. A., D´esert J.- M., et al. 2012, MNRAS, 421, 2342 Terquem C., Papaloizou J. C. B., 2007, ApJ, 654, 1110 Wang S., Ji J., Zhou J.-L., 2012, ApJ, 753, 170 SUPPLEMENTARY ONLINE MATERIAL 6 K. Go´zdziewski, C. Migaszewski, F. Panichi & E. Szuszkiewicz Figure 5. One -- and two -- dimensional projections of the posterior probability distribution for all free parameters of the TTV model. The MCMC chain length is 1,024,000 iterations in each of 512 different instances selected in a small ball around a solution found with the genetic algorithms. First transit epochs Tb,c,d and orbital periods Pb,c,d are expressed in days, planetary masses mb,c,d are expressed in Earth masses, and the uncertainty correction term σ f is given in minutes (b ≡ 1,c ≡ 2,d ≡ 3 are indices of subsequent planets). c(cid:13) 2012 RAS, MNRAS 000, 1 -- 5 0.250.000.25xb0.250.000.25yb0.0300.0550.0800.105Tb+6.897e10.00300.00550.00800.0105Pc+8.9120.40.20.00.20.4xc0.250.000.25yc0.0300.0550.0800.105Tc+7.302e10.0100.0140.0180.022Pd+1.189e10.20.00.2xd0.20.00.20.4yd66.2466.2866.32Td481216mb481216mc481216md0.00250.00500.0075Pb+7.125243240σf0.250.000.25xb0.250.000.25yb0.0300.0550.0800.105Tb+6.897e10.00300.00550.00800.0105Pc+8.9120.40.20.00.20.4xc0.250.000.25yc0.0300.0550.0800.105Tc+7.302e10.0100.0140.0180.022Pd+1.189e10.20.00.2xd0.20.00.20.4yd66.2466.2866.32Td481216mb481216mc481216md243240σf
1002.2392
1
1002
2010-02-11T19:02:51
Irregular satellites of Jupiter: Capture configurations of binary-asteroids
[ "astro-ph.EP" ]
The origins of irregular satellites of the giant planets are an important piece of the giant "puzzle" that is the theory of Solar System formation. It is well established that they are not "in situ" formation objects, around the planet, as are believed to be the regular ones. Then, the most plausible hypothesis to explain their origins is that they formed elsewhere and were captured by the planet. However, captures under restricted three-body problem dynamics have temporary feature, which makes necessary the action of an auxiliary capture mechanism. Nevertheless, there not exist one well established capture mechanism. In this work, we tried to understand which aspects of a binary-asteroid capture mechanism could favor the permanent capture of one member of a binary asteroid. We performed more than eight thousand numerical simulations of capture trajectories considering the four-body dynamical system Sun, Jupiter, Binary-asteroid. We restricted the problem to the circular planar prograde case, and time of integration to 10^4 years. With respect to the binary features, we noted that 1) tighter binaries are much more susceptible to produce permanent captures than the large separation-ones. We also found that 2) the permanent capture probability of the minor member of the binary is much more expressive than the major body permanent capture probability. On the other hand, among the aspects of capture-disruption process, 4) a pseudo eastern-quadrature was noted to be a very likely capture angular configuration at the instant of binary disruptions. In addition, we also found that the 5) capture probability is higher for binary asteroids which disrupt in an inferior-conjunction with Jupiter. These results show that the Sun plays a very important role on the capture dynamic of binary asteroids.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1–12 (2009) Printed 2 March 2018 (MN LATEX style file v2.2) Irregular satellites of Jupiter: Capture configurations of binary-asteroids H. S. Gaspar⋆, O. C. Winter⋆ and E. Vieira Neto⋆ UNESP Univ Estadual Paulista, Grupo de Dinamica Orbital e Planetologia, CEP 12.516-410,Guaratinguet´a, SP - Brazil ABSTRACT The origins of irregular satellites of the giant planets are an important piece of the giant "puzzle" that is the theory of Solar System formation. It is well established that they are not in situ formation objects, around the planet, as are believed to be the regular ones. Then, the most plausible hypothesis to explain their origins is that they formed elsewhere and were captured by the planet. However, captures under restricted three- body problem dynamics have temporary feature, which makes necessary the action of an auxiliary capture mechanism. Nevertheless, there not exist one well established capture mechanism.In this work, we tried to understand which aspects of a binary- asteroid capture mechanism could favour the permanent capture of one member of a binary asteroid.We performed more than eight thousand numerical simulations of capture trajectories considering the four-body dynamical system Sun, Jupiter, Binary- asteroid. We restricted the problem to the circular planar prograde case, and time of integration to 104 years. With respect to the binary features, we noted that 1) tighter binaries are much more susceptible to produce permanent captures than the large separation-ones. We also found that 2) the permanent capture probability of the minor member of the binary is much more expressive than the major body permanent capture probability. On the other hand, among the aspects of capture-disruption process, 4) a pseudo eastern-quadrature was noted to be a very likely capture angular configuration at the instant of binary disruptions. In addition, we also found that the 5) capture probability is higher for binary asteroids which disrupt in an inferior-conjunction with Jupiter. These results show that the Sun plays a very important role on the capture dynamic of binary asteroids. Key words: planets and satellites: formation – minor planets, asteroids – Solar system: formation 0 1 0 2 b e F 1 1 . ] P E h p - o r t s a [ 1 v 2 9 3 2 . 2 0 0 1 : v i X r a 1 INTRODUCTION The existence of more than 350 natural satellites is known, from which approximately 50 % are planetary ones. An in- teresting point about this number, is that before 1997 just a tenth of such objects was known, i.e., the new "CCD ob- servational era" allowed this number to increase by an or- der of magnitude within just a half decade (Gladman et al. 1998, 2000, 2001; Sheppard & Jewitt 2003; Holman et al. 2004; Kavelaars et al. 2004; Sheppard et al. 2005, 2006). The planetary satellites can be distinguished into two char- acteristic groups: regulars and irregulars (Kuiper 1956; Peale 1999). The first group, is characterized by small values of semi-major axis, eccentricities and inclinations. in situ- These characteristics are a strong signature of ⋆ E-mail: [email protected] (OCW); [email protected] (EVN) [email protected] (HSG); ocwin- irregular satellites have retrograde orbital formation through matter accretion from the circumplane- tary disc (Lunine & Stevenson 1982; Vieira Neto & Winter 2001; Canup & Ward 2002, 2006; Mosqueira & Estrada 2003; Sheppard & Jewitt 2003). In contrast, the irregu- lar satellites have large values of semi-major axis (Burns 1986), often high eccentricities and inclinations. A large part of incli- nations higher than 90 degrees (Jewitt & Haghighipour 2007). Another important characteristic of the irregular ones are the family groups, i.e., satellite groups charac- terized by similar orbital elements (Gladman et al. 2001; Kavelaars et al. 2004). These peculiar characteristics are in- compatible with the in situ-formation model through mat- ter accretion (Kuiper 1956), and since they are the ma- jority group of planetary satellites in the solar system, there exists a large scientific interest about their origin. Then, the most plausible hypothesis to explain their ori- gins is that they formed elsewhere and were captured 2 H. S. Gaspar, O. C. Winter & E. Vieira Neto by the planet (Kuiper 1956; Heppenheimer & Porco 1977; Pollack et al. 1979; Colombo & Franklin 1971). However, many studies have shown that gravitational captures un- der three-body-dynamics are temporary (Everhart 1973; Heppenheimer & Porco 1977; Carusi & Valsecchi 1979; Benner & Mckinnon 1995; Vieira Neto & Winter 2001; Winter & Vieira Neto 2001). This fact has induced re- searchers to propose some auxiliary capture mechanism. Among others, we point out four mostly well known: (i) Gas drag capture (Pollack et al. 1979; ´Cuk & Burns 2004): A temporarily captured asteroid becomes perma- nently captured through kinetic energy decrease due to gas drag inside the circumplanetary disk of gas and dust; (ii) Pull-Down capture (Heppenheimer & Porco 1977; Vieira Neto et al. 2004; Oliveira et al. 2007): A temporary captured asteroid becomes permanently captured due to an increase of the Hill's radius of the planet. This increase in Hill's radius occurs due to either planet's mass growth or planet's migration away from the Sun; interaction (iii) Close-approach captures (Colombo & Franklin 1971; Tsui 1999, 2000; Astakhov et al. 2003; Nesvorn´y et al. 2003; Funato et al. 2004): A tem- porarily captured asteroid becomes permanently captured through energy and angular momentum exchanges with an existing satellite; (iv) Capture of binary-asteroids (Agnor & Hamilton 2006; Vokrouhlick´y et al. 2008): One member of a binary- asteroid becomes permanently captured when the binary approaches the planet and disrupts. The capture mechanism of binary-asteroids is very in- teresting since the present observations have shown an in- creasing number of such systems in the main populations of such objects as the Kuiper Belt, Main Belt and Near Earth Asteroids (Noll 2006). Agnor & Hamilton (2006) presented numerical simulations of close encounters between Neptune and a binary-asteroid, where they considered one asteroid comparable to Triton and a secondary, with equal mass or one order of magnitude lower. Their results show that is pos- sible to disrupt the binary when the close approach happens inside a spherical region whose radius, called tidal radius, is given by: rtd = aB (cid:18) 3MP m1 + m2(cid:19)1/3 (1) where aB, MP , m1 and m2 are the binary semi-major axis, planet mass, primary and secondary asteroid masses, respec- tively. A possible outcome after disruption is the capture of one member of the primordial binary-asteroid. The increasing number of binary-asteroid discoveries (Noll 2006) along with the Agnor & Hamilton (2006) results, have motivated us to study the binary-asteroid capture pro- cess in the context of four-body dynamics, where we consid- ered Sun, Jupiter and a pair of asteroids. The purpose of this work is to identify the most important orbital characteristics inherent to the binary-asteroid capture/disruption process that produce the permanent capture of at least one mem- ber. Compared to existing works (Agnor & Hamilton 2006; Vokrouhlick´y et al. 2008), our study considers the inclusion of solar perturbation. We found that the Sun's presence has a crucial influence on the binary capture/disruption process, at least in the planar case. This paper is built with the following structure: Section 2 describes the model we use in our study as well the adopted numerical approach. Section 3 presents the results with an analysis of them. Finally, the last section summarizes our conclusions. 2 CAPTURE MODEL Given the temporary feature of captures in the three-body problem, we have chosen to perform a study under four-body dynamics using the Sun and Jupiter as primary bodies and a binary-asteroid. We basically propose a capture model in which a binary-asteroid first becomes temporarily captured by Jupiter, and then disrupts and has one of its member per- manently captured by Jupiter while the other one escapes. The present paper addresses the early results of a more gen- eral work which is under development. It should be noted that the main goal of the present work is not to reproduce the actual configuration of Jupiter's irregular satellites, but rather, to comprehend how specific configurations can lead an asteroid member of a primordial binary to a permanent capture. As a first stage, we have considered only the pla- nar prograde case. Given this capture model, our task is to search for the initial conditions which yields asteroid tempo- rary captures by Jupiter. The temporary captures, as well as the close encounters, are intrinsic features of solar system formation theories (Pollack et al. 1996; Hahn & Malhotra 2005; Tsiganis et al. 2005; Gomes et al. 2005). Furthermore, temporary captures seem to be a more efficient mechanism to accomplish asteroid captures because there is a longer in- teraction time between the binary-asteroid and the planet rather than a single passage with a very short time of inter- action. 2.1 Procedure In this work, as a first study, we considered only the copla- nar four-body dynamics. Furthermore, we set Jupiter's ec- centricity to zero for all the simulations. In order to perform the numeric studies we used an integrator based on Gauss- Radau spacing (Everhart 1985). In order to verify the inte- grator's accuracy we checked whether the system's total en- ergy holds throughout the integration. We found the energy variation was lower than 10−11. In addition, we have checked the value of the Jacobi constant of the binary-asteroid cen- ter of mass before binary disruptions for all the trajectories. We found that the Jacobi constant variation was lower than 10−9 The adopted procedure basically followed three phases: i) Firstly, we performed a capture time analysis of the sys- tem Sun-Jupiter-particle through which we obtain the, from now on designate, suitable initial conditions, i.e., initial con- ditions which result in a particle's temporary capture by Jupiter. ii) Given the suitable initial conditions, we replace the individual particle by a pair of bodies in order to set up a binary-asteroid, which will be called initial conditions. iii) Finally, we study the binary-asteroid capture through simulations of the system Sun, Jupiter, binary-asteroid, by considering a set of different initial conditions derived from each one of the suitable initial conditions. ) U A ( Y 1.5 1.0 0.5 0.0 -0.5 Initial position Final position -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 X (AU) Figure 1. Example of escape trajectory in backward integration. The particle is in clockwise trajectory under Jupiter fixed frame of referenceThe initial point indicated by blue 'x' corresponds to the primary initial conditions (around Jupiter). The final point indicated by red '+' corresponds to the suitable initial condition, which will be replaced by the binary-asteroid. 2.2 Suitable Initial Conditions In order to obtain the suitable initial conditions, we per- formed a primary capture time study following the steps of Vieira Neto & Winter (2001). It consists of the integra- tion of a particle trajectory using a negative time step un- der the three-body dynamics considering Sun and Jupiter as primaries. By setting the particle to start orbiting around Jupiter we have three possible outcomes depending on par- ticle's initial condition: i) The particle collides with Jupiter. ii) The particle remains orbiting the vicinity of Jupiter up to the final time of integration (104 years), in such cases, the particle's initial conditions are stable ones. iii) The particle escapes from Jupiter's vicinity and begins to orbit the Sun. In order to make easy the comprehension, we finally define these particle's orbital elements around Jupiter as primary initial conditions. Summarizing: Primary initial conditions: Three-body problem initial conditions which are backward integrated in time in order to obtain the suitable initial conditions; Suitable initial conditions: Three-body problem initial conditions which lead the particle to temporary captures by Jupiter. By replacing the particle for a binary-asteroid one derives the initial conditions of the binary-asteroids; Initial conditions: The real initial conditions used to study the binary-asteroid capture dynamics, in which we consider Sun, Jupiter and a pair of asteroids; We are particularly interested on the data relative to escape trajectories, which are capture ones when it is considered time forward. The particle's and Jupiter's orbital elements with respect to the Sun at the instant when the particle escapes1 are taken as the suitable initial conditions, as illus- trates the example in Figure 1. The main results of our three-body dynamics simula- tions, Sun-Jupiter-particle, is shown in the capture time map 1 We check the particle's two-body energy every time interval of one year during the numerical simulations. We consider that the escape occurs when the particle's two-body energy with respect to Jupiter becomes positive. Capture configurations of binary-asteroids 3 4 10 3 10 2 10 1 10 0 10 ) s r a e y ( e m i T (a) 0.2 0.4 0.6 0.8 1 Semi-major axis (r H) 1.0 0.8 y t i c i r t n e c c E 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 y t i c i r t n e c c E 0.2 0.4 0.6 0.8 1 Semi-major axis (r H) (b) Figure 2. (a) Capture time mapping. The color scale correspond to the capture time (in years) for such pair of initial semi-major axis and eccentricity. The green '+' indicates the initial conditions which produce collisions.Initials longitude of pericenter and true anomaly of the particle were set to zero ( = f = 0). (b) Primary initial conditions used to obtain the suitable initial conditions. The green and black 'x' indicate the cases whose capture time are shorter and longer than a thousand of years, respectively. The pink dotted line is given by Equation 2. of Figure 2(a). This is an a×e diagram whose color scale de- notes particle's escape time (in years) in the backward's in- tegration. In this plot, the particle's semi-major axis is given in terms of Hill's radius of Jupiter. This map has been gen- erated by setting the particle's initial longitude of pericenter () and initial true anomaly (f ) with respect to Jupiter as zero. In plot (a) of Figure 2, the yellow region corresponds to the primary initial conditions which do not result in escapes from Jupiter in 104 years. Considering the satellite eccentric- ity up to 0.5, Domingos et al. (2006) obtained an expression 4 H. S. Gaspar, O. C. Winter & E. Vieira Neto for this stable region as a critical semi-major axis inside which the satellites would remain stable. This expression is given by aE ≈ 0.4895(1.0000 − 1.0305eP − 0.2738esat) (2) where eP and esat are the planet's and satellite's eccentrici- ties, respectively. Particularly, in this work, the second term between parenthesis on the right hand side of Equation 2 vanishes since we have taken eP = 0. In order to obtain this expression Domingos et al. (2006) followed the same proce- dure used by Vieira Neto & Winter (2001) and also set the particle's initial longitude of pericenter () and initial true anomaly (f ) as zero. In section 3.3 we will show that the stable region can be more extensive for initial values of and f different from zero. Finally, the non-yellow region on the map (a) of Fig- ure 2 indicates the primary initial conditions that resulted in escape. These are the cases from which we can take the suitable initial conditions. However, it is not feasible to use all the data obtained from the capture time analysis. By an- alyzing the capture times of escape cases, we found only 45 cases in which the capture time exceed one thousand years. It is feasible to take all these long time capture cases to com- pose the set of suitable initial conditions. Among the cases in which the capture times are shorter than a thousand of years, we took an uniformly-spaced grid of points in a × e space in order to complete the set of suitable initial condi- tions with a representative set of the whole data. Plot (b) in Figure 2 summarizes the set of primary initial conditions we used to obtain the suitable initial conditions. The a × e diagram of Figure 3 shows the suitable initial conditions which was obtained from the primary ones. The plot shows the semi-major axis and eccentricities of the main asteroid in the heliocentric frame. Furthermore, all the trajectories are direct with respect to the Sun. It is also plotted, in Fig- ure 3, the Tisserand relations for T = 2.996 and T = 3.036. The Tisserand relation (Tisserand 1896), as one can find in Murray & Dermott (1999), is given by: T = 1 2a +pa(1 − e2) cos(I) ≈ constant where a, e and I are the object heliocentric semi-major axis, eccentricity and Inclination. (3) 2.3 Model for the Binary Asteroid The term binary-asteroid can refer to either a system in which a pair of asteroids of similar masses orbit their common barycenter or an asteroid that has a small satellite (Noll 2006). In order to compose a binary-asteroid by using the suitable initial conditions, we added a second body orbiting the first one. From now on, the main and the secondary asteroids will be called P1 and P2. For simplicity, we model a binary-asteroid by setting P2's orbital elements with respect to P1. In other words, asteroid P2 initially orbits asteroid P1. P2's orbital elements with respect to P1 will be referred as binary's elements. Given our particular interest in Jupiter's irregular satellites, we set the P2's mass as m2 = 1019 kg based on Himalia's mass, the largest irregular satellite of Jupiter (Emelyanov 2005). Using the same mass ratio as Agnor & Hamilton (2006), we set P1's mass m1 = 10 m2 = 1020 kg. y t i c i r t n e c c E 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 Jupiter 4.0 5.0 6.0 7.0 8.0 Semi-major axis (A.U.) Figure 3. Suitable initial conditions a × e heliocentric diagram. The green "x" and black "+" indicate suitable initial conditions which were obtained from primary initial conditions whose cap- ture time are shorter and longer than a thousand years, respec- tively. Red and blue curves are Tisserand relations T = 2.996 and T = 3.036, respectively, which encompass all the suitable initial conditions. As already stated, the main goal of this paper is to iden- tify the most appropriate configurations that would gener- ate permanent capture of one asteroid from a binary system. Thus, since we have a huge range of possibilities, is this pa- per we made some restrictions in order to be able to explore a significant part of the initial conditions space. Among the restrictions, we considered that P2 is always initially in pro- grade circular orbit around P1, (e = I = 0). From each of the 81 suitable initial conditions we de- rived 108 new initial conditions. We vary the initial binary's true anomaly fB from 0 up to 330o in steps of 30o, and the initial binary's semi-major axis aB from 0.1 r h up to 0.5 r h in steps of 0.05 r h. Here r h, with 'h' written in low- ercase, is P1's Hill's radius with respect to the Sun calcu- lated for each one of the suitable initial conditions. Let's make clear that we did not make use of the tidal radius given by Equation 1.The Hill's radius, as one can find in Murray & Dermott (1999), is defined by: rHill = (cid:16) µ 3(cid:17)1/3 a, (4) where µ and a are the mass ratio and the semi-major axis, respectively. The chosen upper semi-major axis limit of 0.5 r h is a well-established limit of stability for prograde systems (Hamilton & Burns 1991, 1992; Domingos et al. 2006). The lower semi-major axis limit was arbitrarily chosen, though tighter binaries do exist. By calculating the Hill's radius of P1 r h with respect to the Sun (Equation 4), one finds that these values vary from 1 × 10−3AU to 2 × 10−3AU., and P2's initial orbital velocity with respect to P1 vary from 7m/s to 22m/s. 2.4 Binary asteroid's capture simulations Since we derived 108 initial conditions from each one of 81 suitable initial conditions, we performed a total of 8 748 binary-asteroid capture trajectory simulations. We set the trajectory integration time to 104 years and the output time step to 10 hours. In order to avoid a large amount of data storage we designed an algorithm which identified the inte- gration main stages. We labeled each instant of these main stages, as follow: T 1: instant when the binary-asteroid is first captured by Jupiter; T 2: instant when the binary-asteroid disrupts; T 3: instant when only one member of the disrupted binary-asteroid escapes from Jupiter; Once the algorithm identifies each one of the three in- stants, it stores the instantaneous system configuration data, as well as the integration instant, in three distinct files. By taking the difference between T 1 and either T 3 or the fi- nal time of integration, we can compute the capture time for each case. This identification algorithm basically con- sists on a two-body energy check-up, every integration step, described as follow: The binary-asteroid approaches Jupiter in a quasi- parabolic trajectory, given that it initially orbits the Sun. For all the cases considered in this study at least one aster- oid two-body energy with respect to Jupiter was positive. At the instant we found that both P1's and P2's two-body en- ergies with respect to Jupiter became negative the instant T 1 is identified. Although, those energies do not become negative at the same instant. Similarly, P2's two-body energy with respect to P1 is initially negative given that P2 initially orbits P1. Thus, if P2's two-body energy with respect to P1 becomes positive, instant T 2 is identified. Otherwise we do not identify neither instant T 2 nor T 3. Rather it could happen a double capture, a mutual collision or a double escape. However, in all the cases considered in this study the binary disrupted or collide with each other. Therefore, after the binary-asteroid's capture by Jupiter, both P1's and P2's individual two-body energies with respect to Jupiter are negative. Furthermore, after the binary-asteroid's disruption, interactions between P1 and P2 become negligible, allowing either P1 or P2 to individually escape from Jupiter. Finally, at the instant in which either P1's or P2's two-body energy with respect to Jupiter became positive the instant T 3 is identified. Three possible outcomes succeed T 3: (i) The remaining asteroid collides with Jupiter, which characterizes a collision; (ii) The remaining asteroid escapes from Jupiter, which characterizes a double escape; (iii) The remaining asteroid stays bound throughout the rest of the integration time, which we characterize as a per- manent capture; 3 RESULTS We show in this section some plots built with the data stored at instants T 1, T 2 and T 3 in three respective subsections. We will discuss some statistical results in the fourth subsec- tion, and show some examples of capture trajectories in the last subsection. Capture configurations of binary-asteroids 5 0.04 0.03 0.02 ) h 0.01 r ( B a ∆ 0.00 -0.01 -0.02 -0.03 -0.04 3 10 2 10 1 10 0 10 -1 10 m a r g o t s i H 0 0.1 0.2 0.3 0.4 0.5 0.6 initial a B (r h) Figure 4. Binary's initial semi-major axis versus variation of the binary's semi-major axis at instant T 1 for the cases which resulted in permanent captures of either P1 or P2, in red and blue triangles, respectively. Red and blue filled lines are P1's and P2's permanent capture histograms, respectively. 3.1 Analysis at the instant of binary capture (T 1) The plot in Figure 4 compares the binary-asteroid separation at instant T 1 with its initial separation, through binary's semi-major axis aB at instant T 1 and initial binary's semi- major axis aB, respectively. Both axis in this plot are mea- sured in units of initial Hill's radius of P1 r h. This plot al- lows us to comprehend the binary-asteroid's evolution from the beginning of the integration to instant T 1. The histograms reveal that i) tighter binary-asteroids are more susceptible to permanent captures than binary- asteroids with large separations and that ii) the capture probability of the binary's smaller member is much higher than the major companion's capture probability. The blue histogram shows a roughly negative exponential behavior of permanent capture probability, with respect to the ini- tial binary separation, which becomes very low for initial aB ' 0.35 rh. Furthermore, beyond initial aB = 0.35 rh one can observe that the permanent capture occurs more of- ten with binaries whose semi-major axis decreased. Con- sequently, these results point out to a limit aB ≈ 0.4 rh be- yond which permanent capture plausibility is negligible. Ev- idently, ∆aB dispersion increases as the initial separations increases since it causes the binary bound to be weaker, con- sequently more susceptible to secular variations due to Sun and Jupiter perturbations. As weaker bounded binaries disrupt more easily, one could expect that they would more easily generate perma- nent captures. However, our results show that tighter bina- ries have higher probability to generate permanent captures. This apparent paradox can be understood in terms of the energy exchange needed to turn a temporary capture into a permanent one. Based on results by Tsui (1999, 2000), which show that it is possible for an asteroid to be kept captured by a planet due to exchange reactions with a local satel- lite, it is possible to explain the binary capture mechanism through energy exchanges in four steps: (i) Lets consider a binary-asteroid, initially orbiting the Sun, which will be captured by Jupiter. Since the binary- asteroid is primordially orbiting the Sun, each one of its members' individual energy is higher than the escape energy 6 H. S. Gaspar, O. C. Winter & E. Vieira Neto ε0, i.e., the minimum necessary energy to allow each asteroid individually escape from Jupiter. (ii) However, once the two asteroids orbits closely their common barycenter, angular momentum and energy ex- changes occurs constantly. Furthermore, after the binary- asteroid be temporarily captured (T 1), Jupiter starts to dis- turb the binary binding. Therefore, the exchange reactions become more intense; (iii) As a consequence of Jupiter's perturbation the binary-asteroid disrupts (T 2). However, before the binary disruption the energy exchanges between the asteroids pro- vides some energy states in which one member's energy is lower than the escape energy ε0, which will not allow it to escape from Jupiter; (iv) Finally, after the binary disruption the interactions between the asteroids become negligible, so that, the aster- oid whose energy decreased to values lower than ε0 remains captured by Jupiter while the other whose energy increased will escape from Jupiter after some time (T 3). Note that these ideas agree well with our results: (i) The rupture of tighter binaries imply on larger energy exchange, implying on higher probability of permanent cap- ture; (ii) The smaller body of the binary is the one that suffers larger energy exchange and consequently is the one with higher probability of permanent capture; (iii) Finally, the the existence of a separation limit agrees well with the conclusions since the binary separation is pro- portional to the binding energy, which, in its turn, corre- sponds to the maximum energy that the asteroids can ex- change. In other words, larger separation-binaries can not provide enough energy exchange between its members in or- der to allow one of them to became permanently captured. Plot (b) of Figure 5 illustrates this energy exchange pro- cess for the trajectory shown in plot (a). Plot (c) shows the time evolution of the Jacobi constant value for each individ- ual asteroid. It illustrates that shortly after instant T 2 the interaction between P1 and P2 becomes negligible. Note also that, the Jacobi constant value of P2 higher than Jacobi con- stant value of Lagrangian point L1. Therefore, shows that P2 will never escape Jupiter's vicinity Nevertheless, we could expect this increasing fraction of captured low-semimajor axis binaries to have a maximum where it must roll over since very tighter binaries should reach a limit in which the binary-asteroids can be consid- ered as a particle and would never disrupt. Furthermore, we should also expect an increasing fraction of mutual collisions inasmuch binary semi-major axis decreases. 3.2 Analysis at the instant of binary disruption (T 2) As defined before, the instant T 2 is characterized by the binary-asteroid disruption. So, the graphics in this subsec- tion refer to elements of each asteroid individually with re- spect to Jupiter. Plot (a) in the Figure 6 is an a×e diagram, at the instant T 2, of the asteroids that remained permanently captured by Jupiter. Plot (b) shows the a×e diagram of the last asteroid to escape from Jupiter for double escape cases. By compar- ing these two diagrams, one finds a region on a×e space in which an asteroid remains permanently captured if its semi- major axis and eccentricity are enclosed within it at instant T 2. It must be clear that Figure 6 shows T 2 instantaneous diagram of osculating elements which must vary in time due to solar perturbation. Furthermore, we should expect some weak interaction between the pair so far as they recede suf- ficiently away from each other. By checking the variation of Jacobi constant value, of each individual asteroid, we could estimate how long it takes to this mutual interaction become negligible. By considering a variation of the order of 10−8 in Jacobi constant value, we found that for the worst case it took about half orbital period about Jupiter (∼ 170 days) to satisfy the condition. By fitting an expression that bounds this region in the a×e space we found a limit on semi-major axis given in terms of eccentricity, as follow: a∗(e) = 0.4500(1.0000 − 0.2046e) (5) The condition a < a∗(e) at instant T 2 can be thought as sufficient but not necessary capture condition. In other words, if the captured object obeys a < a∗(e), then our sim- ulations show that the temporary capture always becomes permanent. However, we also found large number of perma- nent captures that had a slightly greater semi-major axis in a region where there is a mix of captures and double escapes A second graphic at instant T 2 is as an angular illus- trative histogram. One can better understand the angular configuration analyzing the phase angles θ1 and θ2 shown in Figure 7. The angle θ1 gives the P1 phase angle with re- spect to the Sun-Jupiter direction, while, θ2 gives P2 phase angle with respect to the Jupiter-P1 directions. Note that in this system P1 orbits counterclockwise Jupiter and P2 orbits counterclockwise P1. Analysing the angular histograms of Figure 8 we see that: i) the disruption preferentially occurs when Jupiter, P1 and P2 are approximately aligned, i.e., θ2 ≈ 0 or θ2 ≈ 180o. The smaller probability for θ1 ≈ 180o in capture histogram (a), plus the higher probability observed for θ1 ≈ 180o in es- cape histogram (b) indicate that ii) disruptions which occurs when binary-asteroid is located aligned between Jupiter and Sun most likely result in double escapes. On the other hand, histogram (a) shows that iii) permanent capture of P2 most likely results from binary-asteroids which disrupt at a angu- lar position approximately 90o after it cross the Sun-Jupiter line (θ1 ≈ 270o). Finally, from the capture histogram (a) one also notes that permanent captures of P2 succeed from disruptions which occurred when P2 were at inferior con- junction with P1, as seen from Jupiter (θ2 ≈ 180o). The preference for captures with the smaller asteroid unbinding when closer to Jupiter (θ2 ≈ 180o) is simply un- derstood as being due to the velocity vector about its center- of-mass motion being opposite to the planetocentric velocity in that geometry. Consequently, asteroid P2 gets its individ- ual speed, with respect to Jupiter, reduced to a minimum. The preference for captures when θ1 ≈ 270o is under- stood as being due to the solar perturbation, i.e., the Sun's gravity tends to increase the velocity of binary center of mass about Jupiter when θ1 ≈ 90o while it tends to de- crease this velocity when θ1 ≈ 270o. These characteristic angular positions with respect to the Sun-Jupiter line, rein- P1 P2 T T T 1 2 3 1.5 1.0 0.5 ) U A ( Y Capture configurations of binary-asteroids 7 Zoom 0.1 0.0 0.0 Jupiter -0.1 -0.5 -1.0 -0.3 -0.2 -0.1 0.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 X (AU) y g r e n e y d o b - o w t s ' d i o r e t s A r e t i p u J o t t c e p s e r h t i w 2.0x10 1.5x10 1.0x10 5.0x10 -2 -2 -2 -3 0.0x10 0 -5.0x10 -1.0x10 -1.5x10 -2.0x10 -3 -2 -2 -2 P1 P2 Binary T 1 T 2 T 3 -7 9x10 -7 6x10 -7 3x10 0 0x10 -3x10 -7 -6x10 -7 -9x10 -7 y g r e n e s ' y r a n i B 0 2 4 6 8 10 12 14 (a) (b) C C J(P2)-C J(P1)-C J(L1) J(L1) Time (years) T 2 0.025 0.020 0.015 J C 0.010 0.005 0.000 -0.005 T 1 T 2 T 3 0 2 4 6 8 10 12 14 Time (Years) (c) Figure 5. A first example of capture process. In plot (a), red '+' and blue 'x' show the trajectories of P1 and P2 in Jupiter's (black circle) planetocentric non-rotating frame of reference, respectively. The output frequency is 10 days. Dashed grey line and the black arrow, in the zoom window, indicate Sun direction at the disruption instant (T 2). Tracking binary-asteroid trajectory, one sees that it becomes temporarily captured (light blue triangle) by Jupiter, disrupts (green square inside zoom box) and finally has its minor member permanently captured by Jupiter while its major member escapes from Jupiter's vicinity (pink down triangle). Plots (b) and (c) show the time evolution of energy and CJ for the trajectory of plot (a), respectively. In plot (b), red '+' and blue 'x' are the two-body energies of P1 and P2 with respect to Jupiter, respectively, and the pink filled line is the binary's two-body energy, i.e., P2's two-body energy with respect to P1. In plot (c), red '+' is the difference of Cj values calculated for asteroid P1 and L1 Lagrangian point, and similarly , blue 'x' difference for P2 and L1. 8 H. S. Gaspar, O. C. Winter & E. Vieira Neto e e 1.00 0.80 0.60 0.40 0.20 0.00 0.1 0.2 0.3 0.4 0.5 0.6 a (r H) (a) 1.00 0.80 0.60 0.40 0.20 0.00 0.1 0.2 0.3 0.4 0.5 0.6 a (r H) (b) Figure 6. Diagrams of semi-major axis versus eccentricity of the asteroids at the instant T 2. Plot (a) shows the cases which re- sulted in permanent captures of either P1 or P2, in red and blue, respectively. Plot (b) shows the cases which resulted in double escapes. The green '+' corresponds to the orbital elements of the asteroids that escape after instant T 3. Pink dotted straight line is the fitted limit semi-major axis given by Equation 5. force the importance of the solar presence on the dynamics of binary-asteroid captures: 3.3 Analysis at the escape instant of one asteroid (T 3) The data stored at instant T 3 shows the final configuration of the captured asteroid, given that the captured asteroid will not suffer any interaction with its primordial partner. Figure 9 is an a × e diagram of captured asteroid at instant T 3. The pink filled contour is an extended border of a more general region of stability. In fact, we firstly worried about the points located beyond critical semi-major axis found by Domingos et al. (2006) since they represent the permanently captured asteroids. Nevertheless, by performing a more gen- eral capture time analysis we found the extended stability border. We performed this more general study following the same procedure described at section 2.2, but considering dis- tinct initial values for longitude of pericenter 0 and true anomaly f0. From the results, shown in the Figure 10, we found that the pairs of initial values (0 = 0, f0 = 180o) Figure 7. Sketch of angular configuration between the bodies. One can infers a Sun-Jupiter-P1 alignment if θ1=0 or θ1=180oand a Jupiter-P1-P2 alignment when θ2=0 or θ2=180o. (a) (b) Figure 8. Angular histograms of instant T 2. The red circles indicate the angular position of P1 with respect to Sun-Jupiter direction given by angle θ1. The angular sections that surround each red circle indicate the angular position of P2 with respect to Jupiter-P1 direction, given by θ2 (see Figure 7). The color of the angular sections in histogram (a) correspond to the percentages of the total number of P2's permanent captures (972; see Table 1), and the color of angular sections in histogram (b) correspond to the cases which resulted in double escapes (7446; Table 1), given by the respective color scales. White color indicates absolute null number of events. The bodies orbit counterclockwise. Table 1. Percentages of captures and collisions. Description Short capture times Long capture times Simulations Captures Captures of P1 Captures of P2 Double capturesa Collisions a P1 and P2 captured 3 888 7 1 6 0 180 4 860 972 5 966 1 143 and (0 = 180o, f0 = 180o) yield capture time maps in which the stability regions are much larger than for initial values (0 = 0, f0 = 0), shown in Figure 2. By combining the borders of both maps, in such manner we obtain the larger region of stability, we built the referred more general stability edge shown, as a pink filled line, in Figure 9. The points in the plot of Figure 9 shows that the final orbital elements of permanently captured asteroids cover a wide region in the a × e space. Most of the captured objects are very far from the planet (aB & 0.4 rH) and will probably be removed due to perturbations not included in our study. The good candidates to survive are those closer to the planet (aB . 0.35 rH). Since in this study we considered only the prograde planar case, we are not able to compare our results with the currently known Jupiter's irregular satellites. How- ever, we included them in the plot of Figure 9 just to have an idea of the orbital shape of captured objects. In order to make a fair comparison it will be needed to make a study in the 3-D space considering the inclinations. This is a study in progress 3.4 Binary captures in numbers Among 8 748 simulated trajectories, 4 860 are from long cap- ture time primary initial conditions and 3 888 from short capture time ones. Table 1 presents the quantities of perma- nent captures and collisions. Second and third columns show the values with respect to the short and long capture time cases, respectively. As it is shown in Table 1, though perma- nent capture probability of the cases derived from short-time primary initial conditions are low (0.18 %), the permanent capture probability of cases derived from long time condi- tions are much larger (20 %). The collision probabilities for both long and short times derived from primary initial con- ditions have the same order and are not negligible. 3.5 Sample of capture trajectories This section presents some examples of capture trajecto- ries from three distinct cases. Firstly, we already showed in Figure 5(a), a typical example where a binary-asteroid ap- proached Jupiter, became captured, disrupted after a while and had its minor member permanently captured by Jupiter while its major member escaped Jupiter's vicinity. Fig- ure 11(a), shows rare example where the major asteroid re- mains captured by Jupiter after disruption of the primordial binary-asteroid. Furthermore, another peculiarity in this ex- ample is T 2=T 3, i.e., the escape of P2 happens at the in- Capture configurations of binary-asteroids 9 stant of binary disruption. Finally, Figure 11(b) presents our single case among > 8000 performed where both asteroids remained captured by Jupiter up to the end of integration, even after the disruption of the primordial binary-asteroid. In both examples of the Figure 11 we simulated the tra- jectories for 104 years, and for both cases the instant T 2 happens before 102 years of integration. Nevertheless, by checking the Jacobi constant value for each one of the bod- ies of the double capture example (Figure 11b), we found that their values are smaller than the L1 Jacobi constant value. Therefore, these bodies may eventually escape from the planet. 4 CONCLUSIONS In this work we have studied the capture dynamics of binary- asteroids by looking for favorable conditions of capture. Our results present new perspectives about the problem of binary-asteroid captures emphasizing the importance of Sun's role in the dynamics. The Sun is not necessary to pro- duce a binary rupture since Jupiter alone can do it. However, the Sun plays a key role in the disruption process in order to make the capture to become permanent. The results allow us to comprehend about both binary-asteroid's features as well as intrinsic features of capture process' main stages. The observed characteristics at the first main stage, T 1, have revealed that: i) Tighter binary-asteroids are more sus- ceptible to permanent captures than binaries with larger separation. In fact, the permanent capture probability be- haves inversely proportional to the binary's semi-major axis. This results indicate that binary's energy exchange allows the asteroid to become permanently captured. That is, since tighter binaries interactions are more intense, its members can exchange higher amount of energy. In such a way one asteroid can have its energy sufficiently decreased in order to not be able to escape from Jupiter. Through this conclusion, we could argue that binary-asteroids with high eccentricities would disrupt more easily, but would not exchange the nec- essary amount of energy to result in a permanent capture.As mentioned, there must exist a lower binary-separation limit below which the binary never disrupts and consequently there is no capture. The observed characteristics at the second main stage, T 2, tell about process' features. It was shown that the an- gular position of bodies at disruption instant (T 2) are re- lated with the permanent capture probability. Summarizing: ii) Disruption preferentially occurs when both asteroids are approximately aligned with Jupiter; Nevertheless, iii) dis- ruptions which occurs when P2 is located between Jupiter and P1 result more often in permanent capture; Finally, we found that iv) the permanent capture probability is higher when the binary-asteroid disrupts in an eastern quadrature i.e., an angular position approximately 90o after the binary cross the Sun-Jupiter direction. The permanent capture probability for an specific set of initial conditions, such derived from long time primary initial conditions, was shown to reach 20 %. Therefore, the good candidates are those derived from long capture time primary initial conditions. As a note, we give here a reference of a similar work sub- mitted to Icarus journal, which also considers the solar per- 10 H. S. Gaspar, O. C. Winter & E. Vieira Neto 1.0 0.8 0.6 0.4 0.2 e 0.0 0.1 0.2 0.3 0.4 0.5 0.6 a (r H) Figure 9. Same as plot (a) in Figure 6 for instant T 3, though. Red '+' and blue 'x' represent the cases which resulted in permanent captures of either P1 or P2, respectively. Black circumferences represent the orbital elements of Jupiter's real prograde irregular satellites. The pink dotted line is the same as in plot (a) of Figure 6. The filled pink contour is a stability edge we found numerically as discussed in section 3.3. y t i c i r t n e c c E 1.0 0.8 0.6 0.4 0.2 0.0 4 10 1.0 3 10 0.8 2 10 1 10 ) s r a e Y ( s e m i T y t i c i r t n e c c E 0.6 0.4 0 10 0.2 -1 10 0.0 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 Semi-major axis (r H) Semi-major axis (r H) (a) Figure 10. Same as Figure 2 for (a) ω0 = 0 and λ0 = 180o and (b) ω0 = 180o and λ0 = 180o. 4 10 3 10 2 10 1 10 0 10 ) s r a e Y ( s e m i T -1 10 (b) turbation. This work is available on astro-ph (Philpott et al. 2009). Finally, as the main goal of this paper was to address the favorable conditions which makes the permanent cap- ture plausible, we have chosen a procedure to get initial conditions without taking into account where the incoming objects came from. It means that, in this work we have just tried the model plausibility without taking into account how it could reproduce the currently observed objects. So in or- der to get a more realistic probability, it must be considered several aspects as inclinations, mass ratios, binary eccentric- Capture configurations of binary-asteroids 11 Zoom 0.2 0.1 0.0 -0.1 ) U A ( Y 0.0 0.0 -0.0 -0.0 -0.1 -0.2 -0.1 -0.1 -0.1 -0.0 0.0 0.0 Jupiter -0.3 -0.4 -0.5 P1 P2 T 1 2=T 3 T -0.6 -0.4 -0.2 0.0 0.2 X (AU) (a) Zoom Jupiter 0.0 -0.1 -0.2 -0.1 -0.1 0.0 0.1 0.2 0.1 0.0 ) U A ( Y -0.1 -0.2 -0.3 -0.4 -0.5 P1 P2 T T 1 2 -0.6 -0.4 -0.2 X (AU) 0.0 0.2 (b) Figure 11. Two examples of capture pathways. Red '+' and blue 'x' are the coordinates of P1 and P2, respectively. Light blue triangle and green square denote T 1and T 2, respectively. (a) The asteroid P1 remains permanently captured by Jupiter. (b) Both asteroids remain captured by Jupiter even after disruption. ities as well as to study how realistic are the trajectories of incoming objects. ACKNOWLEDGMENTS The comments and questions of an anonymous referee helped to significantly improve this paper. The authors gratefully acknowledge CNPq, FAPESP and CAPES, which 12 H. S. Gaspar, O. C. Winter & E. Vieira Neto have funded this work. This paper has been typeset from a Murray C. D., Dermott S. F., 1999, Solar System Dynam- TEX/ LATEX file prepared by the author. REFERENCES Agnor C. B., Hamilton D. P., 2006, Nature, 441, 192 Astakhov S. A., Burbanks A. D., Wiggins S., Farrelly D., 2003, Nature, 423, 264 Benner L. A., Mckinnon W. B., 1995, Icarus, 118, 155 Burns J. A., 1986, The Evolution of Satellite Orbits, 1 edn. ics, 1 edn. Cambridge University Press Nesvorn´y D., Alvarellos J. L., Dones L., Levison H. F., 2003, AJ, 126, 398 Noll K. S., 2006, in Lazzaro D., Ferraz-Mello S., Fern´andez J. A., eds, Proc. of IAU: Symp. No. 229 Solar system binaries. Cambridge University Press, pp 301–318 Oliveira D. S., Winter O. C., Vieira Neto E., de Felipe G., 2007, EM&P, 100, 233 Peale S. J., 1999, Ann. Rev. A&A, 37, 533 Philpott C., Hamilton D. P., Agnor C. B., , 2009, Three- Body Capture of Irregular Satellites: Application to Jupiter, arXiv:0911.1369v1 The University of Arizona Press, Tucson, pp 117–158 Pollack J. B., Burns J. A., Tauber M. E., 1979, Icarus, 37, Canup R. M., Ward W. R., 2002, AJ, 124, 3404 Canup R. M., Ward W. R., 2006, Nature, 441, 834 Carusi A., Valsecchi G., 1979, Numerical Simulations of Close Encounters Between Jupiter and Minor Bodies, 2 edn. The University of Arizona Press, Tucson, pp 391– 416 Colombo G., Franklin F. A., 1971, Icarus, 15, 186 ´Cuk M., Burns J. A., 2004, Icarus, 167, 369 Domingos R. C., Winter O. C., Yokoyama T., 2006, MN- 587 Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62 Sheppard S. S., Jewitt D., Kleyna J., 2005, AJ, 129, 518 Sheppard S. S., Jewitt D., Kleyna J., 2006, AJ, 132, 171 Sheppard S. S., Jewitt D. C., 2003, Nature, 423, 261 Tisserand F. F., 1896, Trait´e de M´ecanique C´eleste IV, 1 edn. Gauthier-Villars Tsiganis K., Gomes R., Morbidelli A., Levison H. F., 2005, RAS, 373, 1227 Nature, 435, 459 Tsui K., 1999, Planet. Space Sci., 47, 917 Tsui K., 2000, Icarus, 148, 139 Vieira Neto E., Winter O. C., 2001, AJ, 122, 440 Vieira Neto E., Winter O. C., Yokoyama T., 2004, A&A, 414, 727 Vokrouhlick´y D., Nesvorn´y D., Levison H. F., 2008, AJ, 136, 1463 Winter O. C., Vieira Neto E., 2001, A&A, 377, 1119 Emelyanov N., 2005, A&A, 438, L33 Everhart E., 1973, AJ, 78, 316 Everhart E., 1985, in Carusi A., Valsecchi G. B., eds, Dy- namics of Comets: Their Origin and Evolution, Proceed- ings of IAU Colloq. 83, held in Rome, Italy, June 11-15, 1984. Edited by Andrea Carusi and Giovanni B. Valsec- chi. Dordrecht: Reidel, Astrophysics and Space Science Library. Volume 115, 1985,, p.185 An efficient integrator that uses Gauss-Radau spacings. pp 185–+ Funato Y., Makino J., Hut P., Kokubo E., Kinoshita D., 2004, Nature, 427, 518 Gladman B., Kavelaars J., Holman M., Petit J., Scholl H., Nicholson P., Burns J. A., 2000, Icarus, 147, 320 Gladman B., Kavelaars J. J., Holman M., Nicholson P. D., Burns J. A., Hergenrother C. W., Petit J.-M., Marsden B. G., Jacobson R., Gray W., Grav T., 2001, Nature, 412, 163 Gladman B. J., Nicholson P. D., Burns J. A., Kavelaars J., Marsden B. G., Williams G. V., Offutt W. B., 1998, Nature, 392, 897 Gomes R., Levison H. F., Tsiganis K., Morbidelli A., 2005, Nature, 435, 466 Hahn J. M., Malhotra R., 2005, AJ, 130, 2392 Hamilton D. P., Burns J. A., 1991, Icarus, 92, 118 Hamilton D. P., Burns J. A., 1992, Icarus, 96, 43 Heppenheimer T. A., Porco C., 1977, Icarus, 30, 385 Holman M. J., Kavelaars J. J., Grav T., Gladman B. J., Fraser W. C., Milisavljevic D., Nicholson P. D., Burns J. A., Carruba V., Petit J., Rousselot P., Mousis O., Mars- den B. G., Jacobson R. A., 2004, Nature, 430, 865 Jewitt D. C., Haghighipour N., 2007, Ann. Rev. A&A, 45, 261 Kavelaars J. J., Holman M. J., Grav T., Milisavljevic D., Fraser W., Gladman B. J., Petit J., Rousselot P., Mousis O., Nicholson P. D., 2004, Icarus, 169, 474 Kuiper G., 1956, Vistas in Astronomy, 2, 1631 Lunine J. I., Stevenson D. J., 1982, Icarus, 52, 14 Mosqueira I., Estrada P. R., 2003, Icarus, 163, 198
1804.01997
1
1804
2018-04-05T18:00:04
Transiting Disintegrating Planetary Debris around WD 1145+017
[ "astro-ph.EP" ]
More than a decade after astronomers realized that disrupted planetary material likely pollutes the surfaces of many white dwarf stars, the discovery of transiting debris orbiting the white dwarf WD 1145+017 has opened the door to new explorations of this process. We describe the observational evidence for transiting planetary material and the current theoretical understanding (and in some cases lack thereof) of the phenomenon.
astro-ph.EP
astro-ph
Transiting Disintegrating Planetary Debris around WD 1145+017 Andrew Vanderburg and Saul A. Rappaport Abstract More than a decade after astronomers realized that disrupted planetary material likely pollutes the surfaces of many white dwarf stars, the discovery of transiting debris orbiting the white dwarf WD 1145+017 has opened the door to new explorations of this process. We describe the observational evidence for transiting planetary material and the current theoretical understanding (and in some cases lack thereof) of the phenomenon. Overview A bit more than a decade after astronomers first began to suspect that white dwarf stars occasionally disrupt and accrete asteroids and small planets from their primordial planetary systems (Debes and Sigurdsson 2002; Jura 2003), an impressive body of evidence had emerged in support of this scenario. In early 2015, it was known that (a) a large fraction of white dwarf stars (between 25% and 50%) are "polluted" with trace amounts of elements like silicon, iron, calcium and magnesium in their atmospheres (Zuckerman et al. 2010; Koester et al. 2014), (b) many of these polluted white dwarfs also showed evidence of warm rocky material orbiting the star in a debris disk (Barber et al. 2012), and (c) the abundance ratios of heavy elements in the atmospheres of white dwarfs very closely matched the abundance patterns in rocky bodies in the solar system (Zuckerman et al. 2007; Farihi et al. 2013). The generally accepted explanation for these observations was that these polluted white dwarfs host planetary systems which at least partially survived the white dwarf progenitor's evolution off the main sequence. The host star's evolution was not without ill effects: as the host shed its outer layers and began to cool and contract into a white dwarf, the star's mass loss caused changes to the planetary system's dynamics. Numerical simulations have shown that planetary systems whose host stars have undergone this type of mass loss can occasionally perturb small planets or asteroids into highly eccentric orbits which can occasionally have periastron passages close enough to the host star (by this time a white dwarf) to be tidally disrupted. The planetary, lunar (Payne et al. 2016), or asteroidal remnants would then be pulverized into a fine dust (causing the infrared excesses observed around many polluted white dwarf stars) and slowly accreted onto the white dwarf's surface, where the constituent elements would manifest themselves by the presence of spectral lines. However, the evidence for this scenario was entirely circumstantial, relying on analysis of the after-effects of plan- etary disruption. There were occasional detections of transient events indicating possible tidal disruptions in progress (Del Santo et al. 2014; Xu and Jura 2014), but no unambiguous detections of disintegrating rocky material until the discovery of transits around a polluted white dwarf called WD 1145+017. For the first time, astronomers had defini- tively observed the transient process of a white dwarf tidally disrupting a large rocky body (planet, moon, or asteroid), Andrew Vanderburg Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138 Current address: Department of Astronomy, The University of Texas at Austin, Austin, TX 78712, USA e-mail: [email protected] Saul A. Rappaport Massachusetts Institute of Technology, Cambridge, MA 02139, e-mail: [email protected] 1 2 Andrew Vanderburg and Saul A. Rappaport in real time. WD 1145+017 has provided the strongest evidence yet that white dwarfs disrupt their planetary systems, and has given astronomers new ways to study and constrain models of the process. Discovery Observations WD 1145-017 was first identified as a somewhat anonymous white dwarf in 1991 by Berg et al. (1992), who used low- resolution spectroscopy to classify it as a helium-envelope white dwarf. This classification was confirmed by Friedrich et al. (2000), although neither groups' classification spectra were strong enough nor taken at high enough spectral resolution to detect any features besides the deep and broad helium lines. Little attention was given to WD 1145+017 until 2014, when it fell in one of the fields of view of NASA's Kepler Observatory in its extended K2 mission. During the interval between June and August 2014, K2 observed the field containing WD 1145+017 (along with 20,000 other stars) for a bit less than 80 days. Once the K2 data were publicly released, Vanderburg et al. (2015) searched through the data looking for planetary transits. The majority of the stars observed by K2 were main-sequence stars, typically the mass of the sun or lower. However, a handful of other objects, including about 150 white dwarf stars, were proposed by various groups and were observed by K2 as well. A periodogram search of the K2 data revealed about 80 'ordinary' transiting planet candidates around main se- quence stars, as well as a strong transit-like signal at a period of 4.5 hours from WD 1145+017. The transit profile was broad (occupying nearly 30% of an orbital cycle), had a profile that was unlike that of a typical hard-body transit, and was variable in depth. The periodogram and fold about the 4.4989-hour period are shown in Fig. 1. A closer analysis of the K2 light curve also revealed evidence for five other significant periodicities with periods between 4.55 and 4.86 hours (designated Periods B through F), suggesting multiple bodies in close orbits. These other periodicities can be seen in the Lomb-Scargle periodogram in Fig. 1, and the corresponding folded transit profiles are shown in the panels below the periodogram. A significant limitation of the K2 data was its time sampling. For bandwidth reasons, Kepler only can store photo- metric data for the majority of the targets it observes, including WD 1145+017, after an integration time of 30 minutes. Because the transit signal around WD 1145+017 had a period of 4.5 hours, K2 only recorded nine photometric mea- surements per orbit. This limitation is particularly unfortunate for studying objects transiting or eclipsing white dwarfs, because the host star's small size means transits and eclipses happen on timescales of about 1 minute, rather than the hour to day timescales for transits of main sequence stars. Such rapid transit features are totally smeared out by the 30-minute K2 integrations. Motivated by the very short orbital periods found in the K2 data, Vanderburg et al. (2015) began photometric observations of WD 1145+017 using small ground-based telescopes at rapid cadence to confirm and better resolve the transits. After a few nights of high-cadence photometric observations, several things became clear. WD 1145+017 was indeed being transited by objects in a (cid:39)4.5 hour orbital period, but the transits were not always present. The transits were deep, up to 40% of the star's flux, and lasted about 5 minutes – too deep and short in duration to be a transit of anything other than a white dwarf, but too long to be the transit of a small solid body across the white dwarf star (see Fig. 2). The transits were asymmetric with a fast ingress time and slow egress time – similar to transits of disintegrating planets observed around main sequence stars by Kepler (see see van Lieshout and Rappaport, "Disintegrating Rocky Exoplanets"; this Handbook). Finally, the transits did not always occur at the same phase of a 4.5 hour orbit. On two different nights, Vanderburg et al. (2015) observed two convincing transits separated by 4.5 hours, but the pairs of transits observed on each of these nights happened almost 180 degrees out of phase, with respect to a 4.5 hour orbit, further suggesting the possibility of multiple objects in orbit. Medium resolution spectroscopy from the MMT Observatory yielded two additional important clues. First, there was no evidence for radial velocity variations over the 4.5 hour orbital periods (with a limit of ∼500 m s−1), confirming that any bodies transiting WD 1145+017 had to be of planetary mass or less. More importantly, the MMT spectrum revealed that WD 1145+017 exhibited absorption lines from elements heavier than hydrogen and helium, including magnesium, aluminum, silicon, calcium, iron, and nickel. WD 1145+017 is therefore a "polluted" white dwarf – a member of the class of white dwarfs believed to have accreted disrupted planetary material. Data from NASA's WISE spacecraft show evidence for infrared excess emission, another hallmark of polluted white dwarfs and evidence for disrupted planetary material in orbit of the white dwarf. Transiting Disintegrating Planetary Debris around WD 1145+017 3 Vanderburg et al. (2015) interpreted the transits of WD 1145+017 as being produced by dust clouds. This inference was made by analogy with a similar phenomenon of so-called 'disintegrating' planets that are found transiting main- sequence stars and which appear to exhibit dusty tails ("Disintegrating Rocky Exoplanets"; this Handbook). These objects, including KIC 12557548 (Rappaport et al. 2012), KOI 2700 (Rappaport et al. 2014), and K2-22 (Sanchis- Ojeda et al. 2015), show similar features to the objects transiting WD 1145+017, including asymmetric transits with rapidly varying transit depths. The disintegrating planets are all in short-period orbits around their hosts stars, and are highly irradiated, to the point where rocky minerals would likely sublimate rapidly. Perez-Becker and Chiang (2013) showed that a small, sub-Mercury sized object orbiting close to a host main-sequence star could undergo rapid mass loss as rocky material sublimates due to the high irradiation environment, and flows away from the planet in a Parker-type thermal wind. While the idea of small disintegrating rocky objects with dusty effluents causing the transits is at least partially and qualitatively successful, Vanderburg et al. (2015) freely admitted that the observations of WD 1145+017 do not support the exact scenario seen in the systems orbiting main-sequence stars. In particular, there have been no detections of phase shifts in the transits of disintegrating planets around main-sequence stars – when the transits of these canonical disintegrating planets appear, they always happen at the same orbital phase. The cause of the phase shifts between the different sets of transits seen by Vanderburg et al. (2015) was clarified by Croll et al. (2015) who, acting on news of the detection of transits of WD 1145+017, obtained a considerably larger set of ground-based follow-up data than Vanderburg et al. (2015) in May 2015. Over the course of 32 hours of observations, Croll et al. (2015) detected nine transits. A blind periodicity search for these transits revealed that many of the transits seemed to recur with an orbital period about 30 seconds shorter than the dominant period detected by Vanderburg et al. (2015) in K2 data, but that the recurring transits appeared to be scattered over a range of orbital phases. Croll et al. (2015) interpreted these transits as being caused by many different discrete bodies in almost identical 4.5 hour orbits, explaining the phase shifts seen between the sets of transits detected by Vanderburg et al. (2015). Fig. 1 K2 discovery observations of WD 1145+017 (Vanderburg et al. 2015). Top panel: Lomb-Scargle periodogram with amplitudes of the first 3 harmonics summed. Six significant and distinct peaks are identified. Bottom panel: Folded lightcurve for each of the detected periodicities. The corresponding fold period is written above each panel. Meanwhile, an independent group led by Siyi Xu also found their attention drawn to WD 1145+017. Xu et al. (2016) were conducting a spectroscopic survey of white dwarfs with excess infrared emission at high spectral resolution with the goal of detecting heavy elements in white dwarf spectra to learn about the compositions of small extrasolar Harmonic-summed Lomb-Scargle Periodogram4.04.24.44.64.85.0Period [hours]0.00.20.40.60.81.01.21.4Signal AmplitudeP = 10-4ABCDEFPeriod A: 4.49888 hours-0.4-0.20.00.20.4Orbital Phase0.9850.9900.9951.0001.005Relative BrightnessPeriod B: 4.60530 hours-0.4-0.20.00.20.4Orbital Phase0.9940.9960.9981.0001.0021.004Relative BrightnessPeriod C: 4.78283 hours-0.4-0.20.00.20.4Orbital Phase0.9940.9960.9981.0001.0021.004Relative BrightnessPeriod D: 4.55000 hours-0.4-0.20.00.20.4Orbital Phase0.9940.9960.9981.0001.0021.004Relative BrightnessPeriod E: 4.82336 hours-0.4-0.20.00.20.4Orbital Phase0.9960.9981.0001.0021.004Relative BrightnessPeriod F: 4.85848 hours-0.4-0.20.00.20.4Orbital Phase0.9940.9960.9981.0001.0021.004Relative Brightnessaaaabbbbccccddddeeeeffffgggg 4 Andrew Vanderburg and Saul A. Rappaport Fig. 2 Ground-based observations of two transits by Vanderburg et al. (2015), compared to a model transit of a solid body (red) and a dust cloud (blue). We note that the transit shapes do not always have the short ingress-long egress asymmetry. Follow-up observations have shown that while there is some preference for this shape, symmetric transits and long ingress-short egress asymmetries are common as well. asteroids or planets. On 15 April 2015, only a few hours after Vanderburg et al. (2015) detected the first set of transits in their ground-based follow-up, Xu et al. (2016) obtained a high signal-to-noise spectrum of WD 1145+017 with the High Resolution Echelle Spectrometer (HIRES) at Keck Observatory. As Xu et al. (2016) expected from their detection of near infrared excess emission, the spectrum revealed the presence of numerous absorption features corresponding to elements like iron, silicon, nickel, magnesium, and calcium. Xu et al. (2016) were surprised, however, to find broad absorption features near many of the detected metal absorption lines, presumably caused by circumstellar gas orbiting WD 1145+017. Illustrative composite spectra (i.e., summed over 5 different metal lines) are shown in Fig. 3 for the 2015 April observations as well as from a follow-up observation from 2016 February. The circumstellar features are broad (with line widths up to 300 km s−1) and deep (obscuring up to 30% of the star's flux at those wavelengths), and the features are visible in lines from ionized states of iron, magnesium, chromium, titanium, calcium, manganese, and nickel. Fig. 3 Illustrative composite spectra (i.e., summed over 5 different metal lines) for the 2015 April observations as well as from a follow-up observation from 2016 February (Siyi Xu, 2016, private communication.). Note the broad line widths of ∼300 km s−1. The offset of the dashed vertical line from zero indicates the gravitational redshift from the surface of the white dwarf. -10-50510Time from mid-event (minutes)0.40.60.81.01.2Relative BrightnessDust cloud transitSolid-body transit Transiting Disintegrating Planetary Debris around WD 1145+017 5 The circumstellar absorption features found by Xu et al. (2016) are unique among all other polluted white dwarfs. Some white dwarfs show circumstellar gas in emission (Gansicke et al. 2006; Manser et al. 2016) indicating hot gas disks (not necessarily viewed edge-on), and other white dwarfs show weak circumstellar absorption features in ultra- violet spectra and optical spectra (Debes et al. 2012a; Gansicke et al. 2012), but none shows circumstellar absorption at the strengths seen around WD 1145+017. The discovery of circumstellar gas absorption at WD 1145+017 is inde- pendent confirmation of the presence of material orbiting close to the white dwarf between the star and our vantage point on Earth. The same favorable orbital inclination of the WD 1145+017 disrupted planetary system, i.e., near 90◦, is required to see both the transits and the circumstellar gas absorption. Ground-based follow-up observations Ground-based photometric monitoring The first extensive and systematic ground-based photometric observations of WD 1145+017 commenced in 2015 November (Gansicke et al. 2016; Rappaport et al. 2016). Gansicke et al. (2016) photometrically observed WD 1145+017 during 15 nights in 2015 November using 2.4 m and 1 m telescopes. The results are shown in Fig. 4. Several distinct dips can be seen over the course of a single orbit, with some as deep as 50%. Averaged over the orbit, as much as ∼11% of the flux is removed by the dips. It is quite apparent from this result that the source was much more 'ac- tive', in the sense of having more and deeper dips, during this period than during the K2 and the initial ground-based followup observations (Vanderburg et al. 2015; Croll et al. 2015). A number of the dips seen in Fig. 4 could be tracked from night to night, thereby allowing for more precise periods to be derived. The periods found by Gansicke et al. (2016) range from 4.491 to 4.495 hours. These differ by between 0.1% and 0.2% from the K2 'A' period of 4.4989 hours. No sign of the K2 'B' through 'F' periods was found in the Gansicke et al. (2016) data. However, if the dips at these latter periods had remained at the same depths found in the K2 observations (fractions of a percent) they could not have been detected. Thus, it appears that WD 1145+017 was mostly active near the 'A' period in 2015 November, but with much greater dip depths. Fig. 4 Photometric monitoring of WD 1145+017 for an interval of a month using a 2.4-m telescope (from Gansicke et al. 2016). The phasing of the diagram uses a period of 4.4930 hr, approximately the same as the 'A fragments' detected by ground-based observations. 6 Andrew Vanderburg and Saul A. Rappaport Fig. 5 Stack of 6 lightcurves for WD 1145+017 reported by Gary et al. (2016) from 2016 April. The first 5 lightcurves (blue) are from the IAC80 32(cid:48)(cid:48) telescope and the last one (green) is from a 20(cid:48)(cid:48) telescope. Fig. 6 'Waterfall' diagram for WD 1145+017 of the dips recorded over 8 months from 2015 November through 2016 June (Gary et al. 2016). The phasing here is based on the 'A-asteroid' period of 4.5004 hr, essentially the same as found with K2. Starting simultaneously with the Gansicke et al. (2016) observations, Rappaport et al. (2016) and Gary et al. (2016) began a much longer monitoring campaign of WD 1145+017 using small optical telescopes in the 30-80 cm range. A typical set of lightcurves from five sequential nights is shown in Fig. 5. Here the data are stacked vertically and phased to an assumed period of 4.4916 hours. The dips near phases 0 and 0.4 are nearly repeatable over the five nights. However, there is clearly an additional dip that is moving quickly in phase, labeled 'B-dip'. This dip feature was associated by Gary et al. (2016) with an object orbiting at the K2 'B' period. Transiting Disintegrating Planetary Debris around WD 1145+017 7 Gary et al. (2016) (see also Rappaport et al. 2016) found it useful to construct so-called 'waterfall' diagrams to display the complicated temporal behavior of the dip evolution. These diagrams are produced as follows. The dips are all formally fit with a simple analytic function (in this case an asymmetric hyperbolic secant, 'AHS'; see Rappaport et al. 2016), whose fitted parameters are the depth, center time, and ingress and egress times. One can then use these parameters to plot the kind of 'waterfall' diagram shown in Fig. 6. In that diagram, each dip is represented by a rectangular bar, with orbital phase during a single night plotted in the horizontal direction, and observation night plotted vertically. The depth of the dip is proportional to the thickness of the bar, and the duration is equal to the length of the bar. This particular plot is phased to an assumed period of 4.500 hr, the dominant mean period found by Rappaport et al. (2016). As one can see from Fig. 6 there is a collection of dips between phases 0.03 and 0.15 during 2015 November and December. After that, and for the ensuing ∼5 months, the orbiting objects started drifting in phase (the group labeled 'G6121'). The periods associated with these drifting features in the lightcurve were in the range 4.491 to 4.495 hours, in basic accord with those found by Gansicke et al. (2016). We summarize here some of what was learned after 8 months of monitoring WD 1145+017 during the 2015-2016 observing season. First, the activity level of transits was quite high with dips obscuring as much as 10% of the flux averaged around the orbit. This is at least an order of magnitude larger than during the K2 discovery period. The source dust activity level is shown as a function of time in Fig. 7. Second, numerous different periods are seen between 4.490 to 4.500 hour, possibly indicating a dozen different bodies orbiting and emitting dusty effluents. These are all within 0.2% of the K2 'A' period. Only one of the other five K2 periods, i.e., the 'B' period, became active enough to detect with the small monitoring telescopes, but remained detectable for only 2-3 weeks. Third, in all photometric monitoring to date, the transits have never been deeper than 60% of the star's total flux. Finally, we can see from Fig. 7 that the source activity at the start of the 2016-2017 observing season is now even higher than during all of the previous season, and has recently obscured as much as 17% of the orbit-averaged flux, but is currently declining to lower values (B. Gary, private communication). Fig. 7 Activity level of WD 1145+017 since its discovery with K2 (from Gary et al. 2016 and private communication). The 'activity' is defined as the dip depths averaged over an orbital cycle. Finally in regard to the 'A' period, the ground-based observations from Croll et al. (2015), Gansicke et al. (2016), Rappaport et al. (2016), and Gary et al. (2016) all find a period that is slightly, but significantly, shorter than the 4.5 hour A-period found with K2. Rappaport et al. (2016) proposed a model where a large asteroid, about 1/10 the mass of Ceres occasionally released fragments into slightly shorter period orbits. Gurri et al. (2017) simulated the subsequent interactions of multiple fragments in such a shorter-period orbit with the parent asteroid as well as amongst themselves, while Veras et al. (2017) showed explicitly how a rubble pile near the Roche limit discharges fragments into shorter (as well as longer) period orbits. However, since early 2016, there has been no additional observational evidence that the 4.5 hour K2 A-period is persistent or "special," raising some questions about this interpretation. Colour-Dependence of the Transit Depths 8 Andrew Vanderburg and Saul A. Rappaport Assuming that the dips in flux from WD 1145 are a result of dusty effluents from rocky bodies, it is important to learn what we can about the nature of the dust. Once the dust sublimates, its chemical composition, and by implication that of the planetesimal, can be investigated from its transmission spectrum as it crosses the face of the white dwarf (Xu et al. 2016; see the Discovery Observation section). While the material is still in the form of dust grains, we are limited for now, to attempting to measure (i) its size distribution and (ii) some rough information about its chemical composition from both the grain scattering properties and lifetimes against sublimation. Thus far, there have been several attempts to measure the colour-dependence of the dips in the visible (Croll et al. 2015; Alonso et al. 2016; Z. Berta-Thompson 2016, private communication) and in the NIR (Zhou et al. 2016). None of these has succeeded in finding a difference in transit depth with wavelength, and we can thereby set constraints on the dust grain sizes of ∼0.5 µm in the visible band and ∼0.8 µm in the NIR. A good example of this type of measurement is shown in Fig. 8. There are four superposed traces of the flux from WD1145 in different wavebands as a function of time over more than half an orbital cycle showing several dip features. The four traces are in bands covering 480 nm to 920 nm, and there are no systematic differences observed among the different bands. Alonso et al. (2016) showed that the depths of these transits were the same to within a statistical precision of ∼1% from which they were able to rule out small particles below ∼0.5 µm. This limit is derived from the fact that for particles whose sizes are greater than the observing wavelength, the scattering cross section quickly becomes equal to the geometric area of the dust grain. When that occurs, the scattering becomes independent of wavelength. Time-resolved spectroscopy Continued observations of the circumstellar gas absorption lines (Redfield et al. 2016; S. Xu 2016, private com- munication) show that the shape and intensity of these broad lines are highly variable on different timescales. One possible explanation for the broad and shifted absorption lines is a mildly eccentric ring of gas orbiting close to the white dwarf (see also Redfield et al. 2016). It can be shown that for a mildly eccentric orbit (i.e., with e (cid:46) 0.1), the radial velocity of the gas as a function of its transverse location, x, across the face of the host star, is given by: (cid:18) x (cid:19) km s−1 (1) vr (cid:39) v0 (ecosω − x/a) (cid:39) 2030 P1/3 min ecosω − 900 Pmin Rwd where ω, v0, and a are the argument of periastron, orbital speed, and semimajor axis for the orbiting ring of gas, re- spectively. The expression on the right is specifically for the WD 1145+017 system, Pmin is the orbital period expressed in units of minutes, and Rwd is the radius of the white dwarf. Thus, for example, if a line is observed with a range of wavelengths corresponding to radial velocities of −100 to +200 km s−1, this yields ecosω (cid:39) 0.045 and P (cid:39) 6 minutes. Fig. 8 GTC lightcurves of WD 1145+017 taken simultaneously in four wavebands and covering several dips (Alonso et al. 2016). The nearly identical dip profiles in the four bands can be used to constrain the dust grain sizes to (cid:38) 0.5µm. The divergence of the curves after phase 0.22 in the lower panel is due to atmospheric effects. Transiting Disintegrating Planetary Debris around WD 1145+017 Some Theoretical Considerations 9 The pollution of white dwarfs had been a long-standing problem in astrophysics, dating back to the 1970s (Fontaine and Michaud 1979), when it was realized that in order for heavy elements to appear in the spectra of white dwarfs, they must have been recently accreted from external sources (or they would have quickly sunk to the center of the white dwarfs, where they would be un-observable). The history of this problem is extensively discussed in "Characterizing Planetary Systems Around White Dwarfs" (B. Zuckerman and E. Young; this Handbook). In brief, the discovery of dusty debris disks near many polluted white dwarfs (Zuckerman and Becklin 1987; Becklin et al. 2005) and the fact that the abundances of heavy elements found in white dwarfs were quite similar to the elemental ratios in rocky solar system bodies (Zuckerman et al. 2007) led astronomers to believe that these heavy elements were likely the disrupted remains of rocky objects from the white dwarf progenitors' planetary systems. Origin of the Debris: Debes and Sigurdsson (2002) were among the first to suggest that heavy element pollution in white dwarfs was caused by the remnants of ancient planetary systems. Using analytic arguments and numerical studies, Debes and Sigurdsson (2002) showed that when an evolving star undergoes mass loss, previously stable planetary systems could become unstable to close encounters. The critical Hill separation for two equal-mass planets, ∆c, in terms of the semi major axes of two planets, a1 and a2, is given by: a2 − a1 (cid:114)8 (cid:17)2/3 (cid:16) m (2) ∆c = a1 (cid:39) (e2 1 + e2 2) + 9 3 M where e1 and e2 are the eccentricities of the two planets, m is the mass of each of the two planets, and M is the mass of the star. Debes and Sigurdsson (2002) noticed that when the stellar mass M decreases (for a white dwarf progenitor, often by a factor of two or more), the critical separation increases, which can push previously stable systems into the unstable regime, rearranging the planetary system dynamically. This rearrangement, Debes and Sigurdsson (2002) speculated, could cause an increased rate of comets from the outer parts of the planetary system to be scattered inwards to close encounters with the white dwarf. Fig. 9 Artist conception of what the WD 1145+017 debris disk system might look like as viewed from one of the orbiting asteroids. Image credit: NASA/JPL-Caltech Since then, other mechanisms have been proposed to bring planetary material to close encounters with the host white dwarfs. Debes et al. (2012b) showed that stellar mass loss can sweep asteroids into mean motion resonances with giant planets (akin to Jupiter), where the asteroids are perturbed into highly eccentric orbits and close encounters with the white dwarf. Similarly, Bonsor et al. (2011) showed that stellar mass loss in a planetary system can perturb objects from an exo-Kuiper belt close to the white dwarf. Additionally, numerical studies have shown that for systems that border on Hill instability, increasing the Hill radii of the planets can in many cases cause the planetary systems themselves to become unstable, and undergo strong dynamical interactions on relatively short timescales, leading to close periastron passages with the white dwarf (Veras et al. 2013, 2016b). 10 Roche Limit: Andrew Vanderburg and Saul A. Rappaport After rocky bodies have been dynamically perturbed too close to the host white dwarf, they are subject to tidal breakup. The famous result of Edouard Roche, cast in the context of bodies orbiting WD 1145+017 can be written as: scrit = ξ R∗ (ρ∗/ρp)1/3 (3) where scrit is the critical distance from a star of density ρ∗ that a planetesimal of uniform density ρp can come before it is tidally split. For Roche's original problem he found that the constant ξ had a value of 2.45. Newer treatments of the 'tidal splitting' problem have been carried out since then, and many of these are summarized by Davidsson (1999; see also Holsapple and Michel 2006). Taking into account such issues as the criteria for actually splitting the body into separated pieces; what trajectory the body is on; whether the body is corotating with its orbit; and the role that material strength plays, ξ is found to lie closer to the range of 1.2 (cid:46) ξ (cid:46) 1.7. There is also the issue of how centrally concentrated the mass of the planetesimal is, e.g., the case where its central density may be considerably higher than in its mantle. In the limiting (unrealistic) case where the planetesimal can be considered to have all its mass concentrated at the center, one can use the Roche potential to show that in this case ξ (cid:39) 2.1. All of this pertains to cases where the planetesimal under question is larger than a few km in size. Otherwise, for small bodies, they may simply be held together by solid-state forces amounting to tensile strengths that are large compared to the pressures imposed by gravity near the center of the body. One can rewrite Eqn. (3) in a more convenient way for our problem by cubing both sides and solving for the critical density (i.e., below which the body is subject to tidal breakup), to find: ρcrit = 3ξ 3M∗ 4πs3 crit = 3πξ 3 G 1 P2 orb (cid:39) 87 (ξ /2)3 1 P2 hr g cm−3 (4) where in the second term on the right we made use of Kepler's 3rd law, and where in the final expression, ξ has been normalized to a typical value of 2. For a period of 4.5 hours, as in the case of the innermost objects orbiting around WD 1145+017 we find critical densities of 7.9, 5.0, 2.6, and 1 g cm−3, depending on whether one uses Roche's value of ξ , the value of ξ from the Roche potential, or other more recent values for ξ , respectively. Debris Rings: Once large rocky bodies have been tidally disrupted due to their close approaches to the white dwarf host, the result will be an eccentric ring of large chunks of debris (Veras et al. 2014, 2015). Through a process known as a "collisional cascade" (see, e.g., Kenyon and Bromley 2002; Kenyon and Bromley 2004; Kenyon et al. 2016; Wyatt et al. 2011) the sizable rocky bodies in this orbiting ring of debris undergo a sequence of collisions which eventually break down the few large objects into a very wide array of object sizes. These will range from planetesimals, to asteroids, rocks, pebbles, and dust, with a rough power-law size distribution. In Figure 9 we show an artist's conception of what such a debris disk might look like. Note the size distribution ranging from asteroids down to dust. At the current epoch we are apparently witnessing the debris disk in WD 1145+017 after it has undergone a colli- sional cascade – which may, in fact, still be ongoing. The main goal of any comprehensive model would be to relate the properties of such a debris disk to the observational dipping activity in WD 1145+017, the high-velocity gas ab- sorption lines, and the excess NIR emission. The asymmetric transit profile and wildly variable transit depths observed from WD 1145+017 are qualitatively similar to transits of disintegrating planets around main sequence stars ("Disin- tegrating Rocky Exoplanets"; this Handbook), and it is believed that similar physical processes may be responsible for both types of transits, but with some significant differences. Perhaps the primary difference between the dusty-tailed planets around main-sequence stars and the situation in WD 1145+017 is the existence of numerous orbiting bodies in the white-dwarf system as opposed to a single more-substantive body orbiting the main-sequence stars. As we have seen, there are likely a dozen or more independently orbiting bodies in WD 1145+017 of a size that can emit sufficient dust to block a significant fraction of the white dwarf's light. Transiting Disintegrating Planetary Debris around WD 1145+017 11 Fig. 10 Sublimation mass-loss rates for different minerals on the surface of a Ceres-sized body orbiting WD 1145+017 (from Vanderburg et al. 2015). Explaining the Transits The transits around WD 1145+017 set it apart among all other polluted white dwarfs, and understanding what causes them is crucial to understanding the system as a whole. Among the various dips in flux, a number of them are uniquely identifiable, long lasting (i.e., for up to months), and of more or less constant shape and duration. Here, we describe three possible models of how the starlight is blocked. Continuous Dust Production Sublimating from Small Rocky Bodies The first model for explaining the transits of WD 1145+017 is that dust is emitted continually from an orbiting rocky body at a relatively constant rate – merely giving the illusion of a permanent feature. This scenario requires a substantive underlying body in a fairly stable orbit to produce the deep and persistent transits observed. For tidally locked objects in nearly circular 4.5 hour orbits around WD 1145+017, the temperature of the object at the substellar point is about 1675 K and the average temperature over the dayside hemisphere is a bit lower at about 1410 K. These temperatures are high enough that many minerals will sublimate. The vapor pressure, pvap, and the knowledge that any vapors produced will freely stream away from the surface of such low-mass objects, allow us to calculate a mass loss flux J from the surface of the body into a vacuum: (cid:115) M J = α pvap 2πkBTeq (cid:20) − M Lsub kBTeq (cid:21) + b where pvap = exp (5) and where M is the material's molecular mass, Lsub is the material's latent heat of sublimation, kB is Boltzmann's constant, Teq is the temperature of the body, b is an empirically measured constant, and α is the "sticking coefficient" and is roughly 0.1 to 0.3. Vanderburg et al. (2015) calculated mass loss rates for various different materials from a single Ceres sized object orbiting WD 1145+017 and found that several different materials and minerals, in particular orthoclase, albite, iron, and fayalite could sublimate and produce dust at the rates necessary to explain the transits of WD 1145+017. We show in Fig. 10 mass loss rates due to sublimation for a Ceres size object as a function of the surface equilibrium temperature for a variety of common minerals. Figure 8 of van Lieshout and Rappaport ("Disintegrating Rocky Exoplanets"; this Handbook) gives the sublimation rate as a function of planetesimal size for a fixed surface temperature of 2000 K. The realization from more recent large photometric monitoring campaigns (Gansicke et al. 2016; Gary et al. 2016) that there are likely many smaller fragments producing deep transits and that these fragments are likely considerably 100012001400160018002000Temperature (K)10410610810101012Mass Loss (g s-1)OrthoclaseAlbiteIronFayaliteEnstatiteForsteriteQuartzCorundumSiCaveragesubstellar100012001400160018002000Temperature (K)10410610810101012Mass Loss (g s-1) 12 Andrew Vanderburg and Saul A. Rappaport smaller in order to coexist in very close orbits (Rappaport et al. 2016) makes it more difficult to account for the deep transits via sublimation of refractory minerals. Vanderburg et al. (2015) calculated that the transits seen around WD 1145+017 require mass-loss rates of order 109 g s−1, and found that the mass loss fluxes calculated from Equation 5 could reproduce those mass loss rates if the surface area was that of a Ceres sized object with a radius of roughly 500 km. However, if the fragments producing most of the transits at the A-period are actually closer in size to Halley's comet (∼5-10 km radius), as suggested by Rappaport et al. (2016), then their surface areas are roughly 103-104 times smaller than Vanderburg et al. (2015) assumed. At the same mass loss fluxes, it therefore becomes difficult (but not out of the question) for sublimation of refractory elements to eject enough material to produce the observed transits. Some scenarios which might mitigate this problem and yield increased mass loss from small fragments include the presence of volatile elements (like water ice), which sublimates much faster than refractory materials like iron, or a higher surface area to mass ratio for the fragments than previously assumed. Alternatively, mass loss may not be completely driven by sublimation, and other processes such as volcanic activity or even collisions may contribute to the ejection of material causing the deep transits we see. Persistent dust from impulsive collisions The second scenario involves an impulsive ejection of dust. In this scenario, dust is ejected in discrete events. The wavelength dependence of the Mie absorption cross sections (and the corresponding inferred emissivities) for smaller dust particles causes these particles to be quickly heated to quite high temperatures (i.e., (cid:38) 2000 K) which leads to rapid sublimation. The result may be that the surviving grains are only those with sizes (cid:38) few µm. The calculated values of β , the ratio of radiation pressure forces to gravity, for these larger residual grains may be as low as β (cid:39) 0.001. Such small values of β result in minimal radiation pressure to push the grains into a different orbit than the emitter, and, likewise, the orbital decay timescale due to the Poynting-Roberston effect may be much longer than the observation timescale. Therefore, in this scenario, the dust is released and simply hangs around the emitting asteroid for considerable intervals of time. The main effect in spreading such a dust cloud may be shearing due to its finite radial extent. If the cloud is 'ribbon-like', i.e., narrow in the radial direction, then it can avoid shearing. However, if it is extended by as much in the radial direction as it is perpendicular to the mid-plane (i.e., comparable to the size of the white dwarf), then the cloud would shear completely around the orbit within a couple of days. Azimuthal Asymmetry in a Dust Ring A final scenario invokes a quasi-permanent dust ring orbiting WD 1145+017 viewed edge on, which has azimuthal density asymmetries that cause differential extinction. When an over-dense region passes in front of the host star, it would cause a transit-like event. These asymmetries could not be self-gravitating – no reasonable masses (Gurri et al. 2017; Veras et al. 2016a) could create a Hill sphere large enough to produce the dips we see – so the asymmetries would likely have to be driven by some orbiting shepherding objects, such as for some of the structures found in Saturn's rings (see e.g., Murray et al. 2005, 2008; Hyodo and Ohtsuki 2015). There are several observational arguments against this scenario. The first is that we often see transits, or diminishings of brightness, but not brightenings of the star. One would expect that the presence of over-densities that cause dimming events would also be accompanied by the presence of under-densities, which would cause brightening events when they pass in front of the host star. Second, the density waves in such a dust ring caused by shepherding bodies would be expected to show dip structures that have more manifest symmetries, and would not likely produce the isolated sharp narrow dip features that are sometimes observed in the lightcurves. Finally, it is not obvious how such an obscuring dust ring would change in overall density by more than an order of magnitude over the course of months as is seen (e.g., Fig. 7). Future Work The WD 1145+017 system is highly complex, rapidly evolving, and poses difficult observational and theoretical chal- lenges. However, there are several paths forward to learning more about this object and the general process of how planetary material is tidally disrupted. In this section, we discuss both observational and theoretical studies which should be undertaken to further our understanding. Continued Photometric Monitoring: The simplest future observations of WD 1145+017 are perhaps among the most important. In particular, it is crucial to continue observing WD 1145+017 photometrically with small to moderate- class telescopes. Continued photometric monitoring will allow us to track the transit activity level and determine the Transiting Disintegrating Planetary Debris around WD 1145+017 13 timescales on which the transit activity level changes. These observations are setting the first observational constraints on how long the tidal disruption of rocky bodies around white dwarfs last and how they progress, which are leading to new insights about the nature of the disrupted object (Veras et al. 2017). A more ambitious and far-reaching goal for photometric monitoring would be to detect and characterize objects orbiting in the sub-dominant B-F periods in the same way that objects in the A-period have been characterized to date. Now that photometric monitoring has revealed an object transiting at the sub-dominant B-period from K2, it is reasonable to believe that continued monitoring may reveal the frequency of transits of the B-F periods found in K2 relative to the numerous transits associated with the A-period, and lead to a stronger understanding of why the A-period experiences higher activity than the other periods. Finally, it would be good to carry out at least some photometry with large-aperture telescopes at shorter time-cadence to look for low-level and rapid timescale variability. Continued High-Resolution Spectroscopic Monitoring: High resolution spectroscopy of WD 1145+017 has revealed intriguing patterns and changes in the circumstellar absorption features (Xu et al. 2016; Redfield et al. 2016), but these observations are in general photon-starved. At 17th magnitude, high spectral resolution observations of WD 1145+017 are just barely possible on the short ((cid:46) 5 minute) timescales of the photometric variability with 10-meter class optical telescopes today. Future observations of WD 1145+017 with the next generation of efficient high-resolution spec- trographs on 30-meter class telescopes should yield significantly superior signal-to-noise ratios on the timescales of transits, making a detailed examination of the short-timescale variability possible and revealing the interplay between the transiting dust and circumstellar gas. UV spectroscopy: Ultraviolet spectroscopy of polluted white dwarfs often reveals the presence of additional elements in their spectra, and it is likely that WD 1145+017 will be no exception. UV spectroscopy of WD 1145+017 is crucial to obtain while the Hubble Space Telescope (HST) is still operating in order to learn more about the composition of the disintegrating material and the circumstellar gas (which has previously been found in other white dwarfs in the ultraviolet, Gansicke et al. 2012). These observations are forthcoming; an ongoing HST program led by Siyi Xu is obtaining ultraviolet spectroscopy of WD 1145+017 with the Cosmic Origins Spectograph (COS). Gaia Parallax: ESA's Gaia mission will measure parallaxes of over a billion stars, including WD 1145+017. By comparing the precise trigonometric distance to WD 1145+017 with a photometric distance estimate, it will be possible to determine whether WD 1145+017 is significantly extincted, and therefore to infer the presence of dust in the line of sight (even while out of transit). Mid-Infrared Observations: Multi-wavelength observations of debris disks around white dwarfs can yield information about the extent and orientation of the disks (Jura 2003). This information is especially interesting in the case of WD 1145+017 because of the additional information we know about the system, in particular the orbital radii and inclination of the transiting debris. Naively, one would expect that for an edge-on system with an optically thick debris disk, where the orbits and debris disk are viewed at an inclination angle close to 90◦, the luminosity of the disk would be relatively small, yet for the edge-on orbits at WD 1145+017 the disk luminosity is somewhat large. Multi-wavelength photometric observations in the mid-infrared can yield information about the orientation and size of the debris disk that are not possible with current data (which only extends out to about 5 microns). Since the end of the Spitzer cryogenic mission in 2009, there are no telescope resources capable of detecting such faint sources at wavelengths beyond 5 microns, but the James Webb Space Telescope, scheduled to launch in 2018, will once again provide that capability at unprecedented sensitivity. Measuring the WD 1145+017 debris disk's spectral energy distribution into the mid-infrared could indicate a misalignment between the debris disk and the transiting material's orbits, or it could show that the disk is likely not optically thin (like the disk surrounding SDSS J155720.77+091624.6 Farihi et al. 2016). Finally, with regard to mid-IR observations, it is worth noting that measurements of the silicate feature at 10 µm can directly reveal the presence of circumstellar silicates in dust grains (see, e.g., Jura et al. 2009). Polarimetry: It would be quite challenging to measure possible polarization of light from WD 1145+017. Because there is not likely to be any preferred grain alignment via magnetic fields, any polarization would be due to scattering. If polarization can be measured, this would be quite useful in confirming the presence of dust and learning more about its spatial distribution. Questions for Future Theoretical and Observational Investigations: Finally, here we list five questions which we believe are necessary to answer in the near future if we are to fully understand the WD 1145+017 system. While continued and further observations will help, these questions will likely require detailed calculations and study. 14 Andrew Vanderburg and Saul A. Rappaport • Is the dust production that feeds any of the dust clouds continuous, i.e., due to sublimation, or episodic, e.g., due to collisions of rocky bodies? • If the dust production is episodic, then what keeps dust-cloud dip features orbiting coherently, in some cases for months? The dust clouds must be considerably larger than the Hill's sphere of any plausible body that is emitting the dust, and therefore the dust must be unbound. • What causes the dramatic changes in dust cloud numbers and optical depth on timescales of days to years, like the change between observations by Vanderburg et al. (2015) and Croll et al. (2015), and those by Gansicke et al. (2016) and Gary et al. (2016) a year later? • What sets the apparent maximum flux decrease of ∼60% in the light of the white dwarf? Is this a result of changes in the optical depth of dust that covers essentially all of the white dwarf surface, or due to optically thick clouds that can, for some reason, never cover more than 60% of the white dwarf surface? • How do the dust clouds become large enough to cause the large transit depths we see? In order to obscure large portions of the white dwarf the debris must be orbiting in different planes that are tilted by a range of angles spanning at least ±1/2◦. In turn, this implies velocity components that are perpendicular to the mean orbital plane of ∼3 km s−1, which are much too large for thermal speeds. This seems to require collisions. However, such high speeds would lead to rapid angular spreading of any resultant dust clouds that is much larger than observed. Finite sublimation lifetimes may help in regard to this latter issue. Future observations and inquiries along these lines should improve our understanding of the many processes at work around polluted white dwarfs like WD 1145+017 and contribute to our knowledge of the ultimate fate of planetary systems (including our own solar system) after their host stars retire from the main sequence. Acknowledgements The authors are grateful to Bruce Gary, Siyi Xu, and Zach Berta-Thompson for quite helpful discussions and for sharing some of their unpublished results regarding WD 1145+017. We also acknowledge Ben Zuckerman for his insightful comments about the manuscript. References Alonso R, Rappaport S, Deeg HJ Palle E (2016) Gray transits of WD 1145+017 over the visible band. A&A589:L6 Barber SD, Patterson AJ, Kilic M et al. (2012) The Frequency of Debris Disks at White Dwarfs. ApJ760:26 Becklin EE, Farihi J, Jura M et al. (2005) A Dusty Disk around GD 362, a White Dwarf with a Uniquely High Photospheric Metal Abundance. ApJ632:L119–L122 QSO survey. ApJS78:409–421 ing WD 1145+017. ArXiv e-prints Berg C, Wegner G, Foltz CB, Chaffee FH Jr Hewett PC (1992) Spectroscopy and spectral types for 387 stellar objects from the large, bright Bonsor A, Mustill AJ Wyatt MC (2011) Dynamical effects of stellar mass-loss on a Kuiper-like belt. MNRAS414:930–939 Croll B, Dalba PA, Vanderburg A et al. (2015) Multiwavelength Transit Observations of the Candidate Disintegrating Planetesimals Orbit- Davidsson BJR (1999) Tidal Splitting and Rotational Breakup of Solid Spheres. Icarus142:525–535 Debes JH Sigurdsson S (2002) Are There Unstable Planetary Systems around White Dwarfs? ApJ572:556–565 Debes JH, Kilic M, Faedi F et al. (2012a) Detection of Weak Circumstellar Gas around the DAZ White Dwarf WD 1124-293: Evidence for the Accretion of Multiple Asteroids. ApJ754:59 Debes JH, Walsh KJ Stark C (2012b) The Link between Planetary Systems, Dusty White Dwarfs, and Metal-polluted White Dwarfs. Del Santo M, Nucita AA, Lodato G et al. (2014) The puzzling source IGR J17361-4441 in NGC 6388: a possible planetary tidal disruption Farihi J, Gansicke BT Koester D (2013) Evidence for Water in the Rocky Debris of a Disrupted Extrasolar Minor Planet. Science 342:218– Farihi J, Parsons SG Gansicke BT (2016) A Circumbinary Debris Disk in a Polluted White Dwarf System. ArXiv e-prints Fontaine G Michaud G (1979) Diffusion time scales in white dwarfs. ApJ231:826–840 Friedrich S, Koester D, Christlieb N, Reimers D Wisotzki L (2000) Cool helium-rich white dwarfs from the Hamburg/ESO survey. Gansicke BT, Marsh TR, Southworth J Rebassa-Mansergas A (2006) A Gaseous Metal Disk Around a White Dwarf. Science 314:1908 Gansicke BT, Koester D, Farihi J et al. (2012) The chemical diversity of exo-terrestrial planetary debris around white dwarfs. Gansicke BT, Aungwerojwit A, Marsh TR et al. (2016) High-speed Photometry of the Disintegrating Planetesimals at WD1145+017: Evidence for Rapid Dynamical Evolution. ApJ818:L7 Gary BL, Rappaport S, Kaye TG, Alonso R Hambsch FJ (2016) WD 1145+017 Photometric Observations During 8 Months of High ApJ747:148 event. MNRAS444:93–101 220 A&A363:1040–1050 MNRAS424:333–347 Activity. ArXiv e-prints Transiting Disintegrating Planetary Debris around WD 1145+017 15 Gurri P, Veras D Gansicke BT (2017) Mass and eccentricity constraints on the planetary debris orbiting the white dwarf WD 1145+017. Holsapple KA Michel P (2006) Tidal disruptions: A continuum theory for solid bodies. Icarus183:331–348 Hyodo R Ohtsuki K (2015) Saturn's F ring and shepherd satellites a natural outcome of satellite system formation. Nature Geoscience MNRAS464:321–328 8:686–689 Jura M (2003) A Tidally Disrupted Asteroid around the White Dwarf G29-38. ApJ584:L91–L94 Jura M, Farihi J Zuckerman B (2009) Six White Dwarfs with Circumstellar Silicates. AJ137:3191–3197 Kenyon SJ Bromley BC (2002) Collisional Cascades in Planetesimal Disks. I. Stellar Flybys. AJ123:1757–1775 Kenyon SJ Bromley BC (2004) Collisional Cascades in Planetesimal Disks. II. Embedded Planets. AJ127:513–530 Kenyon SJ, Najita JR Bromley BC (2016) Rocky Planet Formation: Quick and Neat. ApJ831:8 Koester D, Gansicke BT Farihi J (2014) The frequency of planetary debris around young white dwarfs. A&A566:A34 Manser CJ, Gansicke BT, Koester D, Marsh TR Southworth J (2016) Another one grinds the dust: variability of the planetary debris disc at the white dwarf SDSS J104341.53+085558.2. MNRAS462:1461–1469 Murray CD, Chavez C, Beurle K et al. (2005) How Prometheus creates structure in Saturn's F ring. Nature437:1326–1329 Murray CD, Beurle K, Cooper NJ et al. (2008) The determination of the structure of Saturn's F ring by nearby moonlets. Nature453:739– 744 Payne MJ, Veras D, Holman MJ Gansicke BT (2016) Liberating exomoons in white dwarf planetary systems. MNRAS457:217–231 Perez-Becker D Chiang E (2013) Catastrophic evaporation of rocky planets. MNRAS433:2294–2309 Rappaport S, Levine A, Chiang E et al. (2012) Possible Disintegrating Short-period Super-Mercury Orbiting KIC 12557548. ApJ752:1 Rappaport S, Barclay T, DeVore J et al. (2014) KOI-2700b: A Planet Candidate with Dusty Effluents on a 22 hr Orbit. ApJ784:40 Rappaport S, Gary BL, Kaye T et al. (2016) Drifting asteroid fragments around WD 1145+017. MNRAS458:3904–3917 Redfield S, Farihi J, Cauley PW et al. (2016) Spectroscopic Evolution of Disintegrating Planetesimals: Minutes to Months Variability in the Circumstellar Gas Associated with WD 1145+017. ArXiv e-prints Sanchis-Ojeda R, Rappaport S, Pall´e E et al. (2015) The K2-ESPRINT Project I: Discovery of the Disintegrating Rocky Planet with a Cometary Head and Tail EPIC 201637175b. Astrophys J, submitted Vanderburg A, Johnson JA, Rappaport S et al. (2015) A disintegrating minor planet transiting a white dwarf. Nature526:546–549 Veras D, Mustill AJ, Bonsor A Wyatt MC (2013) Simulations of two-planet systems through all phases of stellar evolution: implications for the instability boundary and white dwarf pollution. MNRAS431:1686–1708 Veras D, Leinhardt ZM, Bonsor A Gansicke BT (2014) Formation of planetary debris discs around white dwarfs - I. Tidal disruption of an Veras D, Leinhardt ZM, Eggl S Gansicke BT (2015) Formation of planetary debris discs around white dwarfs - II. Shrinking extremely extremely eccentric asteroid. MNRAS445:2244–2255 eccentric collisionless rings. MNRAS451:3453–3459 Veras D, Marsh TR Gansicke BT (2016a) Dynamical mass and multiplicity constraints on co-orbital bodies around stars. Veras D, Mustill AJ, Gansicke BT et al. (2016b) Full-lifetime simulations of multiple unequal-mass planets across all phases of stellar Veras D, Carter PJ, Leinhardt ZM Gansicke BT (2017) Explaining the variability of WD 1145+017 with simulations of asteroid tidal MNRAS461:1413–1420 evolution. MNRAS458:3942–3967 disruption. MNRAS465:1008–1022 Wyatt MC, Clarke CJ Booth M (2011) Debris disk size distributions: steady state collisional evolution with Poynting-Robertson drag and other loss processes. Celestial Mechanics and Dynamical Astronomy 111:1–28 Xu S Jura M (2014) The Drop during Less than 300 Days of a Dusty White Dwarf's Infrared Luminosity. ApJ792:L39 Xu S, Jura M, Dufour P Zuckerman B (2016) Evidence for Gas from a Disintegrating Extrasolar Asteroid. ApJ816:L22 Zhou G, Kedziora-Chudczer L, Bailey J et al. (2016) Simultaneous infrared and optical observations of the transiting debris cloud around WD 1145+017. MNRAS463:4422–4432 Zuckerman B Becklin EE (1987) Excess infrared radiation from a white dwarf - an orbiting brown dwarf? Nature330:138–140 Zuckerman B, Koester D, Melis C, Hansen BM Jura M (2007) The Chemical Composition of an Extrasolar Minor Planet. ApJ671:872–877 Zuckerman B, Melis C, Klein B, Koester D Jura M (2010) Ancient Planetary Systems are Orbiting a Large Fraction of White Dwarf Stars. ApJ722:725–736
1706.02566
1
1706
2017-06-08T13:14:53
Analog Experiments on Tensile Strength of Dusty and Cometary Matter
[ "astro-ph.EP" ]
The tensile strength of small dusty bodies in the solar system is determined by the interaction between the composing grains. In the transition regime between small and sticky dust ($\rm \mu m$) and non cohesive large grains (mm), particles still stick to each other but are easily separated. In laboratory experiments we find that thermal creep gas flow at low ambient pressure generates an overpressure sufficient to overcome the tensile strength. For the first time it allows a direct measurement of the tensile strength of individual, very small (sub)-mm aggregates which consist of only tens of grains in the (sub)-mm size range. We traced the disintegration of aggregates by optical imaging in ground based as well as microgravity experiments and present first results for basalt, palagonite and vitreous carbon samples with up to a few hundred Pa. These measurements show that low tensile strength can be the result of building loose aggregates with compact (sub)-mm units. This is in favour of a combined cometary formation scenario by aggregation to compact aggreates and gravitational instability of these units.
astro-ph.EP
astro-ph
Analog Experiments on Tensile Strength of Dusty and Cometary Matter Grzegorz Musiolik, Caroline de Beule, Gerhard Wurm aFakultät für Physik, Universität Duisburg-Essen, Lotharstr. 1, 47048 Duisburg, Germany 7 1 0 2 n u J 8 . ] P E h p - o r t s a [ 1 v 6 6 5 2 0 . 6 0 7 1 : v i X r a Abstract The tensile strength of small dusty bodies in the solar system is determined by the interaction between the composing grains. In the transition regime between small and sticky dust (µm) and non cohesive large grains (mm), particles still stick to each other but are easily separated. In laboratory experiments we find that thermal creep gas flow at low ambient pressure generates an overpressure sufficient to overcome the tensile strength. For the first time it allows a direct measurement of the tensile strength of individual, very small (sub)-mm aggregates which consist of only tens of grains in the (sub)-mm size range. We traced the disintegration of aggregates by optical imaging in ground based as well as microgravity experiments and present first results for basalt, palagonite and vitreous carbon samples with up to a few hundred Pa. These measurements show that low tensile strength can be the result of building loose aggregates with compact (sub)-mm units. This is in favour of a combined cometary formation scenario by aggregation to compact aggreates and gravitational instability of these units. 1. Introduction The solar system swarms with small dusty bodies. Cometary activity close to the sun bears witness to their dusty nature. Small bodies are partly modified by or due to collisions over the last billion of years (Krivov et al., 2006). However, especially for comets most dust fea- tures are likely inherited from the origin within the so- lar nebula. Cometary dust was studied in detail by the Stardust mission to comet 81P/Wild 2 (Brownlee et al., 2006; Hörz et al., 2006; Trigo-Rodríguez et al., 2008). More generally, ideas of dusty growth come from exper- iments or modeling of early phases of planet formation (Blum & Wurm , 2008; Johansen et al., 2014). Comet flybys of space probes, e.g at Borrelly, Wild 2, and Churyumov-Gerasimenko show that comets are highly porous (Brownlee et al., 2006; Davidsson & Gutiérrez, 2004; Davidsson & Gutierrez, 2004; Pätzold et al., 2016) and mainly built up of par- ticles with sizes from tens of nanometers to a millimeter (Hörz et al., 2006; Fulle et al., 2015). With densities of 0.18-0.3 g/cm3, 0.38-0.6 g/cm3, and 0.533 g/cm3, re- spectively, they are much less dense than pure water ice. High porosities can be the result of quite different internal structures. One extreme might be macropores between otherwise solid, monolithic, large fragments. Email address: [email protected] (Grzegorz Musiolik) Preprint submitted to Elsevier This resembles a rubble-pile structure. In this case, the building blocks of the pile might be rather solid with an increased strength due to collisional processing and formation of aqueous alteration minerals which fill the pores as e.g. suggested by Trigo-Rodríguez & Blum (2009). The drill on the lander Philae on comet Churyumov-Gerasimenko as part of the Rosetta mis- sion, e.g. seemed to have hit solid ground 3 cm below the surface (Spohn et al., 2015). This would be in favour of such ideas. Also the dust can provide porosity on differenct scales. On one side there might be very compact dust aggregates, which are then building the larger body again with macro-pore rubble-pile structure. On the other side a homogeneous micro-porosity between mi- crometer dust grains with low volume filling factor can also provide a global high porosity. An analysis by Gustafson & Adolfsson (1996) shows that packing fac- tors of meteors can even be as low as 0.12, which means that the major part of volume might be cavities. Some of this structure can be inferred from obser- vations e.g. of the dust production during the active phase or particle entry and ablation into Earth's at- mosphere or IDP analysis (Borovicka , 1993; Flynn, 1989; Trigo-Rodríguez et al., 2003; Fulle et al., 2015; Brownlee et al., 1985). Most of the cometary inter- planetary dust particles (IDPs) collected in the strato- sphere are dense aggregates; typically from 10 µm to July 24, 2018 100 µm in size and consisting of up to 1 µm grains (Brownlee et al., 1985; Mannel et al., 2016). Fulle et al. (2015) also conclude from measurements of dust ejected from comet Churyumov-Gerasimenko that only a small fraction of the dust is really highly porous. Related to the morphology, it is the tensile strength which therefore might be a quantity to constrain early planet formation scenarios. For solid monolithic "rocky" material, values for tensile strength are way too large to allow shedding of a dust population. If larger bodies are composed of particles on the order of 1 µm in a homogeneous way, their tensile strength is still on the order of 1 kPa (Blum et al., 2006). This assumes that particles only stick together by surface forces and not by chemical bonding, i.e. are not sintered together but that they are packed densely. Such values might be expected from a pure collisional formation of cometesimals with a homogeneous porosity also on large size scales. From meteor observations Trigo-Rodríguez & Llorca (2006, 2007) conclude that cometary particles of sizes from 102 to 104 microns entering the atmosphere themselves have tensile strengths of 0.4 to 10 kPa. Tsuchiyama et al. (2009) measured the tensile strength for ∼250µm car- bonaceous chondrites to 0.3-30 Mpa. All this gives clear evidence that on the small scale dust is rather compact and firmly sticks to each other. Recently discussed are gravitational collapse scenar- ios for planet (comet) formation where aggregates of mm to cm size form first by collisions in agreement to a large tensile strength. However, these are then con- centrated gently and mostly bound together by gravity later (Johansen et al., 2014). In this case, the overall tensile strength of the nucleus' surface is determined by the contacts between these larger granules. The ten- sile strength measured for such dust granule bodies in analog laboratroy experiments is only at the 1 Pa level (Skorov & Blum, 2012; Blum et al., 2014; Brisset et al., 2016). These values might be set in the context of pro- cesses occuring during active cometary phases. When comets approach the inner solar system the sublima- tion of ices (H2O, CO2, CO) leads to a near surface pressure. If this pressure is larger than the tensile strength of the overlaying material, dust is ejected. Sub- limation of ice might provide a pressure on the order of a few Pa (Skorov & Blum, 2012). Groussin et al. (2015) estimate the tensile strength of different parts of comet Churyumov-Gerasimenko to values between a few Pa and 1kPa at maximum. Combined with the measured tensile strength of loosly bound compact ag- gregates, Skorov & Blum (2012); Blum et al. (2014); Groussin et al. (2015) conclude that these are indica- 2 tions of an early gravitational instability scenario for comets. In any case all these findings indicate the ex- istence of different size scales. Constituents with higher tensile strength are assembled to larger bodies with a lower tensile strength. Obviously, the relation between the grain size and the tensile strength is important. Nonetheless, besides the work by Skorov & Blum (2012); Blum et al. (2014) on large assemblies of dust aggregates there are no dedi- cated laboratory experiments. Tensile strength has never been explicitly measured before for an individual small aggregate with only few grains within. A question im- portant to judge on formation scenarios would e.g. be: What size distribution of sticky aggregates can build larger assemblies of what tensile strength? This gen- eral question is far beyond this paper but we approach this problem here by new laboratory experiments. We developed a technique and used it for the first time to determine the tensile strength of small individual dust aggregates which are only (sub)-mm in size with con- stituents on the same size scale. Our experiments are based on thermal creep gas flow at low pressure. While we use it as technique here to measure tensile strength for small aggregates the inter- pretation of this work might go beyond. Thermal creep is not restricted to a laboratory environment. Particle disintegration by thermal creep might occur naturally during planet formation and evolution on the surface of illuminated bodies (Kelling et al., 2011; Kocifaj et al., 2011; de Beule et al. , 2014). Therefore, this work not only shows laboratory measurements but also allows speculations on the existence of a potentially disastrous process, the knudsen barrier, for weak bodies under cer- tain conditions (Wurm, 2007). 2. Experimental setup: Ground based For the experiments, free moving small aggregates are needed. As we intend to measure low tensile strength, these aggregates are rather fragile by defini- tion. We used two different approaches for our measure- ments. One setup works in the ground based laboratory. The other setup was used in drop tower experiments. Both are complementary. In this section, we describe the ground based experiments. On the ground, free moving aggregates of different sizes can be generated in the following way: At low ambient pressure dust samples are placed on a heater. This leads to thermal creep through the aggregates and an overpressure between the dust and the heater, which levitates dust aggregates (Kelling & Wurm, 2009). As this support is adjusting itself to only compensate the weight and is distributing stress over a larger area and partially volume, even low-tensile-strength aggregates can be levitated by this method. Thermal creep is strong enough if the ratio between the mean free path of the gas molecules and the grain or pore size within the aggregate is comparable to or larger than 1. Therefore, low ambient pressure is needed to levitate the aggregates. The main part of the setup thus consists of a heater placed within a vacuum chamber (Fig.1). Before the experimental run, the heater is kept at room temperature and for the test experiments basaltic dust with grain sizes smaller than 125 µm is placed in its center (Fig.1 (1)). The vacuum chamber is evacuated to 200 Pa and the heater is heated to 620 K (Fig.1 (2)). Eventually, dust aggregates start to levitate over the hot surface (Fig.1 (3)). At this stage, the heater is tilted and the levitating aggregates slip down from the hot surface and are in free fall (Fig.1 (4)). No longer exposed to the hot surface, the aggregates rapidly cool from the outside in due to thermal radiation (Fig.1 (5)). Hence, the maximum temperature is inside the aggregate. In the same way as thermal creep leads to levitation for the aggregates with hot bottom and cool top, gas is now pumped into the aggregate by thermal creep. On small millisecond timescales, an overpressure builds up inside the aggregate. This effect is also known as Knudsen compressor (Knudsen, 1909). If the pres- sure is larger than the tensile strength, the aggregate is explosively disintegrating (Fig.1 (6)). The aggre- gate disintegration is observed by a camera with 25000 frames per second at a resolution of 6 µm. This tem- poral and spatial resolution allows to measure the initial acceleration at the start of disintegration. With observed fragment size and known density this directly translates into a tensile strength at the moment the aggregate frag- mented. Figure 2 shows three snapshots of a time se- quence of one of such disintegration observed. 3. Experimental setup: Microgravity The ground based experiments described above allow an observation at high frame rates and high spatial resolution. Therefore, the accelerated motion upon disintegration can be measured (in 2d) and the values for the tensile strength can directly be determined. Due to the nature of high resolution imaging, only few aggregates which fragment just at the right moment can be observed with this method. No information is gained about a broader aggregate sample, which (2) (3) thermal creep Heater (5) (6) (1) (4) FG Figure 1: Principle of the laboratory experiment. A dust sample is placed on a heater within a vacuum chamber (200 Pa) forming ag- gregates (1). The heater is set to 620 K (2) and based on thermal creep described by (Kelling & Wurm, 2009), the aggregates start to levitate over the hot surface (3). The heater is then tilted and the free aggregates slip into free fall (4). Now their outside can cool and a temperature maximum is established inside the aggregate (5). Ther- mal creep then leads to a gas flow from the cool surroundings to the warm inside. On short timescales, this generates an overpressure and disintegration of the aggregates occurs (6). 1 mm time [ms]: 1 1.8 2.6 Figure 2: Unprocessed data of a basaltic dust aggregate disintegrating in free fall due to an overpressure induced by thermal creep in the laboratory experiment. might disintegrate at locations not observed or which might not fragment at all. To sample a larger number of aggregates it would be beneficial if tensile strengths could be deduced from low resolution imaging of the fragments after the explosion. Then a larger field of view could be observed and statistical sampling would be possible. We therefore also carried out first drop tower experiments where aggregate fragments can be traced over long timescales of seconds. A sketch of the experiment is shown in Fig.3. In this case, suspended aggregates of particle samples were generated by light induced ejections from a particle bed. This ejection mechanism is described, e.g. by Wurm & Krauss (2006); Kelling et al. (2011); de Beule et al. (2013); de Beule et al. (2014), and we refer to these papers for details. In the context of this work, it is only important that the released aggregates were part of an illuminated dust bed. An infrared laser beam (955 nm) illuminated a 3.4 cm spot on the dust 3 vacuum chamber camera l d e i f t h g i r b dust bed Figure 3: Setup of the microgravity experiment. Suspended dust ag- gregates are provided by the effect of light induced erosion in mi- crogravity. Once the aggregates left the dust bed and the radiation, they cool from outside-in and an overpressure evolves inside, leading eventually to disintegration. bed with 7 cm diameter. The light flux was 5.4 and 12.7 kW/m2. We used basalt and palagonite samples in this work. The palagonite was tempered for 1 hour at 900 K to avoid effects of adhering water during the experiments. Furthermore, in a parallel setup a red laser beam (655 nm) with a 5.5 mm spot and a flux of 12.6 kW/m2 was used for a sample of vitreous carbon spheres. We estimate the temperature of the dust bed to 420 K - 650 K (Kocifaj et al., 2011). The sample was in a vacuum chamber at 400 Pa ambient pressure. Once agglomerates leave the surface, they are free to radiate into space and cool from the outside-in, especially if they leave the laser beam. As in the ground based experiments, a radial temperature gradient within the aggregates evolves. The experiments reported here were carried out under microgravity at the drop tower in Bremen. This allows sufficient time for aggregate observations and trajectory reconstruction. 4. Data analysis 4.1. Disintegration at 25000 fps With the given frame rate of 25000 fps in the ground based experiments, the initial acceleration can be re- solved which is shown in Fig.4 for one fragment. We subtract the gravitational acceleration and rescaled the Figure 4: Example of a fragment position with time in both dimen- sions observed after a particle disintegration. Gravitational accelera- tion was subtracted. Overplotted are linear (red, dashed) and parabolic (black, solid) fits. positions by an arbitrary linear motion to center the tra- jectory somewhat. The latter is only for visual reasons and is neither showing the trajectory in the center of mass system nor it has an effect on the determination of the acceleration. After removal of the gravity de- pendence, the trajectory of a fragment can be divided into three distinct parts: a linear motion before disin- tegration as part of the aggregate, an accelerated part during disintegration and a second linear part after the expansion. Therefore, we approximately fit the result- ing curves with two linear parts before and after the explosion with a parabola in between. Due to the ex- pansion of the gas within the aggregate the acceleration decreases with time. The approximation of a parabola with constant acceleration therefore underestimates the initial acceleration by a small factor. However, within the accuracy of the data the deviation from a parabola cannot be quantified. We consider the deviation negli- gible compared to the current uncertainties. We further assume that the velocity is continuous at the connec- tions and set the end points, so that the linear tracks be- fore and after the acceleration have the lowest squared 4 difference to the data. The mean force acting on a fragment can be derived from the acceleration data (parabolic part of the track) if the mass of the fragment is known. Taking the cross section of the fragment this also gives a measure of the pressure acting during the explosion which equals the tensile strength. Mass and cross section are estimated from the observed size of the fragment. We approximate the grains by spherical particles of equivalent observed cross section. From this size we estimate the mass by assuming a bulk density of 2.89 g cm−3. For 10 frag- ments analyzed for one fragmentation we get the mean value for the tensile strength ∆Pacc= 31.4± 17.4 Pa. The observation of such explosions is not trivial and we only consider this as first measurement here to show the capability of the technique. Also, the data lack a 3d information. Therefore, accelerations are systematically underestimated. (1) Figure 5: Example of a fragment position with time in both dimen- sions observed after a particle disintegration. x and y refers to camera coordinates. Overplotted are linear fits. g~cm3 for basalt, 1.45 g~cm3 for vitreous carbon and 2.5 g~cm3 for the palagonite sample. Adding the mass of all individual grains or smaller aggregates after disintegra- tion often differs from the deduced mass, which can be deduced independently in the same way for the original aggregate. The difference is typically a factor of about 0.5. We attribute the "missing" half of the mass to the porosity of the aggregate with 50% being a typical value for dust samples consistent with our observation. Fig.7 shows the velocity of fragments after disintegra- tion of the aggregate over their mass. In total, 7 dis- integrations are shown. There is a clear trend that par- ticles with larger mass move slower. The spread has different origins. Particles can move towards or away from the observer and would be misleadingly attributed as slower in a 2d projection. Also, some fragments might pick up rotational energy. Related to this, there is a variation among the fragments in shape which in- fluences the energy that a fragment can get. The dot- ple explosion model as detailed below. A power law ted lines have a slope of−1~3 resulting from the sim- of −1~3 was fitted to fragments of individual explo- sions and particle velocities were scaled to a relative velocity of 1 at a mass of 1 ng to compare the differ- ent events. To complete the data, the size distributions of the dust samples were measured with a Mastersizer 3000 (Malvern Instruments) and are shown in Fig.8. We note that the basalt sample of the ground based experi- ment was sieved to below 125µm. 5. Disintegration model In the following, we assume that an overpressure within the aggregate is responsible for the disintegra- 5 4.2. Disintegration at 1000 fps During the further analysis we take a look at disinte- grating aggregates at 1000 fps from the drop tower ex- periments. Here, it is no longer possible to resolve the acceleration as shown above. With the help of a simple disintegration model, as detailed below, we nevertheless can estimate the value for the tensile strength. The idea is to determine the kinetic energy of all fragments and compare this to the energy released by the expansion of the gas due to the overpressure. Fig.5 shows an example for a recorded fragment trajec- tory. From the first image, where the particle is observed to 50 images later, the motion can be described by a constant velocity. Note, that there are different scales compared to Fig.4. All fragments can be described by a linear motion, which allows a straight forward determi- nation of the kinetic energies. On very long timescales the fragments change their trajectories, as particles cou- ple to the gas and residual flows within the experiment chamber. This is not important here and is not consid- ered further. The motion of the visible center of mass of the aggre- gate before fragmentation is also determined. The mo- tion of the fragment after the disintegration can then be described in the center of mass system. Fig.6 shows the final data of an explosion reduced to the size and veloc- ity of the fragments in the center of mass system. As mentioned above, we define the size of the grains by the radius of a sphere with equivalent observed cross sec- tion. To estimate the mass we use a bulk density of 2.89 Figure 6: Velocities (2d projections) and sizes of the fragments in the center of mass system for a disintegration of a basalt aggregate. The sizes of the circles are proportional to the actual sizes of the fragments. tion. It is worth to mention that this occurs on very small spatial scales but 100 µm scales for pressure induced tension were also seen in recent work by de Beule et al. (2015). As a model, we approximate an aggregate as a core- mantle structure. The individual grains form a shell sur- rounding a central pore as seen in Fig.9. If the pres- sure inside the pore-space increases beyond the ten- sile strength due to thermal creep, the aggregate (shell) is disrupted and smaller sub-units down to individual grains are accelerated by the pressure as long as the ex- pansion takes place. The maximum pressure difference that can be achieved is given by (Knudsen, 1909) ∆Pmax= Po⎛⎝  Ti To − 1⎞⎠ (2) where Po is the ambient pressure, Ti is the temperature ature. Ti depends on the intensity of the corresponding laser and the duration of exposure, which is not exactly known. Assuming values between 1-5 seconds we get within the pore, and To= 300 K is the ambient temper- Ti ∈ [420K, 480K] for the 5.4 kW/m2 laser (used for basalt) and Ti ∈ [560K, 650K] for the 12.6 kW/m2 and sults lead to overpressures ∆Pmax ∈ [73Pa, 105Pa] for basalt and ∆Pmax∈ [146Pa, 189Pa] for JSC-600 and the 12.7 kW/m2 lasers (used for vitreous carbon spheres and JSC-600) referring to Kocifaj et al. (2011). These re- Figure 7: 2d velocity over mass. The velocities for different disinte- grations are scaled to 1 at a mass of 1 ng in order to compare aggre- gates with different absolute velocities. In addition, lines proportional to m−1~3 according to a simple explosion model are shown. The up- per and lower lines are provided to guide the eye. The data cover 7 events (4 for basalt, 2 for palagonite (JSC600) and 1 for carbon glass spheres). glass-spheres. The real pressure at disintegration can be lower. The actual value for the pressure is deduced from the kinetic energy of the fragments which can be measured. Keeping in mind that velocities vi for fragments with masses mi are 2d projections only, we get the lower limit of the total kinetic energy in the center of mass system mi 2 v2 i ∆Ekin= Qi which is on the order of 0.1 nJ. The pressure release energy can be calculated by (3) (4) Epres= S (Pi− Po)dV. As it is not possible to resolve the pressure change dur- ing the expansion in the images taken with 1000 fps, we assume that the shell is expanding by a distance of spa- where κ is the isentropic exponent. This gives  tial resolution adiabatically or with a constant b= PiV κ Epres= S VE VI  b +Po(VI−VE) by b= (Po+ ∆P)V κ (5) with the initial volume of the pore VI and the volume of the pore after the expansion VE. The constant b is given bV1−κ E − V1−κ 1− κ V κ − Po dV= I . Thus we get I ∆P= 6 Epres− Po(VI− VE)(cid:6)(1− κ) I V1−κ E − V1−κ  V κ I − Po (6) Figure 8: Volume size distributions in arb. units (probability density but not normalized) of the dust samples studied. Figure 10: Tensile strength for different aggregates. The upper limits are the maximum pressure differences from Eq.(2). The lower limits result from the estimation of the smallest expanding volume VE. VI VE THERMAL CREEP DISINTEGRATION Figure 9: Aggregate model with a central pore surrounded by the in- dividual grains and thin capillaries allowing thermal creep to enter the pore. in total. As the fragmentation occurs in a dry air envi- ronment, we use κ= 1.4 (Kouremenos et al., 1986). As- suming an initial pore size between [0.3RA, 0.8RA] with the radius of the aggregate RA this allows to estimate VI. In order to calculate VE we consider radii between [RA+ 17µm, RA+ 188µm], which is an estimation based on the recorded dataset for 25000 fps ground based ex- periments. It represents the range of determined ex- pansion volumes, where fragments are accelerated. Fi- nally, Eq.(6) allows us to calculate the tensile strength as Epres= ∆Ekin. Here we assume that the volume work Epres is fully transferred into kinetic energy of the frag- ments. The results are shown in Fig.10. If the process was not fully adiabatic, only a fraction of Epres could be converted into volume work and thus also into ki- netic energy. With this, the tensile strength ∆P could be rather overestimated by the model due to ∆P∼ Epres. As every fragment is accelerated by the same pressure, the acceleration is proportional to 1~r with a fragment size r. As all fragments are accelerated over the same time, also the final (measured) velocities are depending on 1~r or 1~m1~3. The mass dependence of the ejecta velocity seen in Fig.7 is in agreement with this. For the ground based experiments we can apply both methods. As seen above we can directly measure the acceleration and deduce a tensile strength of ∆Pacc = 31.4± 17.4 Pa.. Using the energy method we get a tensile strength of ∆PE= 44.7+60.3 −30.1Pa for the same aggregate. This is in good agreement and the average values only differ by a factor 1.4. 6. Conclusion We measured the tensile strength of small aggregates consisting of basalt, JSC and vitreous carbon on the or- der of∼ 100Pa for 10 individual grains in the size range of 50 µm to 500 µm for the first time. We used high as well as low temporal and spatial resolution. We showed that both methods produce similar values. Therefore, low resolution imaging with a wider field of view can provide statistical information on tensile strength of a larger aggregate sample in future experiments. This is the first proof of concept that low tensile strength of par- ticle aggregates can be measured. This technique allows to quantify values in the range expected for cometary matter. The measured tensile strength is on the order of 10 Pa to 100 Pa for three different materials. While these are first measurements they show that a size scale of 100 µm is possible as constituent to build the weak surface of comets. These constituents themselves can be more stable sub-units. In general, 7 the results are consistent with former work on me- teroid fragmentation and strength measurements by the Stardust mission of 81P/Wild 2 (Brownlee et al., 2006; Trigo-Rodríguez & Llorca , 2006, 2007; Hörz et al., 2006). The results fit also the idea of a gravitational instability formation scenario for comets where com- pact aggregates with high tensile strength grow first and are then assembled to weak larger bodies by grav- ity as suggested by Blum et al. (2014); Groussin et al. (2015); Skorov & Blum (2012). It has to be kept in mind that a minimum aggregate size is needed to con- centrate them. E.g. Bai & Stone (2010); Carrera et. al (2015); Drazkowska & Dullemond (2014) show that Stokes numbers (ratio between gas-grain friction time and orbital period) larger than 10−2 are needed for streaming instabilities though this might even be re- duced further (Yang et al., 2016). This number depends on the disk model but especially in the comet forming region in the outer solar system this might rather cor- respond to cm or larger aggregates. Our results show that the tensile strength is already low for much smaller 100µm constituents. So while different tensile strength for the constituent and the aggregate is in support of in- stabilities, there is a mismatch in size scales. In a proba- bly simplified view, the interpretation is that either grav- itational instabilities did not form comets or instabilities were able to concentrate much smaller grains than pre- viously expected. 7. Acknowledgements This project was supported by DLR Space Manage- ment with funds provided by the Federal Ministry of Economics Affairs and Energy (BMWi) under grant number DLR 50 WM 1242. G. Musiolik is funded by the DFG. We thank B. Steffentorweihen for help with the lab experiments. We also appreciate the reviews of the two referees. References Blum, J., Schräpler, R., Davidsson, B. J. R., & Trigo-Rodríguez, J. M. 2006, ApJ, 652, 1768 Blum, J., & Wurm, G., 2008, ARAA, 46, 21-56 Blum, J., Gundlach, B., Mühle, S., & Trigo-Rodriguez, J. M., 2014, Icarus, 235, 156 Bai, X.-N. & Stone J.M., 2010, ApJL, 722, L220. Borovicka, J., 1993, Astronomy & Astrophysics, 279, 627-645. Brisset, J. & Heisselmann, D. & Kothe, S. & Weidling, R. & Blum, J., 2016, Astronomy & Astrophysics 593, A3 Brownlee, D., 1985, Ann. Rev. Earth Planet. Sci. 13, 147-173 Brownlee, D.& Tsou, P.& AlÃl'on, J.& Alexander, C. M. D.& Araki, T.& Bajt, S., ... & Borg, J., 2006, Science, 314(5806), 1711-1716. 8 Carrera, D. & Johansen, A. & Davies, M. B., 2015, Astronomy & Astrophysics, 579, A43 Davidsson, B. J. R., & Gutiérrez, P. J. 2004, Icarus, 168, 392 Davidsson, B. J. R., & Gutierrez, P. J. 2004, Bulletin of the American Astronomical Society, 36, 1118 de Beule, C., Kelling, T., Wurm, G., Teiser, J., & Jankowski, T., 2013, ApJ, 763, 11 de Beule, C., Wurm, G. and Kelling, T., Küppers, M., Jankowski, T. & Teiser, J., 2014, Nature, 10 de Beule, C., Wurm, G., Kelling, T., Koester, M., Kocifaj, M., 2015, Icarus, 260, 23 Dominik, C., Tielens, A.G.G.M., 1997, ApJ, 480:647-673 Drazkowska, J. & Dullemond, C. P., 2014, Astronomy & Astro- physics, 572, A78. Flynn, G. J., 1989, Icarus, 77(2), 287-310. Fulle, M., Della Corte, V., Rotundi, A., et al., 2015, ApJ, 802, L12 Groussin, O., Jorda, L., Auger, A. T., KÃijhrt, E., Gaskell, R., Ca- panna, C., ... & Knollenberg, J., 2015, Astronomy & Astrophysics, 583, A32. Gustafson, B.Å. S., and L. G. Adolfsson 1996, The Cosmic Dust Connection. Springer Netherlands, 349-355. Hörz, F., Bastien, R., Borg, J., Bradley, J. P., Bridges, J. C., Brownlee, D. E., ... & Djouadi, Z., 2006, Science, 314(5806), 1716-1719. Johansen, A., Blum, J., Tanaka, H., et al. 2014, Protostars and Planets VI, 547 Kelling, T. & Wurm, G. 2009, Phyis. Rev. Lett., 103, 21 Kelling, T., Wurm, G., Kocifaj, M., Klacka, J., & Reiss, D., 2011, Icarus, 212, 935 Knudsen, M. 1909, Annalen der Physik, 336, 205 Kocifaj, M., Klacka, J., Kelling, T., & Wurm, G. 2011, Icarus, 212, 935 Kouremenos, D.A., Acta Mechanica, 1987, 1-4, 81-99 Krivov, A. V., Löhne, T., & Sremcevi´c, M. 2006, Astronomy & As- trophysics, 455, 509 Mannel, T. & Bentley, M. S. & Schmied, R. & Jeszenszky, H. & Levasseur-Regourd, A. C. & Romstedt, J. & Torkar, K. 2016, Monthly Notices of the Royal Astronomical Society, 462, S304 Pätzold, M. & Andert, T. &, Hahn, M. &, Asmar, S.W. & Barriot, J.- P.& Bird, M. K. & Häusler, B.& Peter, K. & Tellmann, S. & Grün, E. & Weissman, P. R.& Sierks, H. & Jorda, L. & Gaskell, R. & Preusker F. & Scholten, F., Nature, 530, 63-65 Skorov, Y. V. & Blum, J. 2012, Icarus, 221, 1 Spohn, T., Knollenberg, J., Ball, A. J., et al. 2015, Science, 349, 020464 Trigo-Rodríguez, J. M., Llorca, J., Borovicka, J. & Fabregat, J. 2003, Meteoritics & Planetary Science, 38(8), 1283-1294. Trigo-Rodríguez, J.M., & Llorca, J. 2006, Monthly Notices of the Royal Astronomical Society, 372, 655 Trigo-Rodríguez, J.M., & Llorca, J. 2007, Monthly Notices of the Royal Astronomical Society, 375, 415 Trigo-Rodríguez, J. M.& Domínguez, G.& Burchell, M. J.& Hörz, F. & Llorca, J. 2008, Meteoritics & Planetary Science, 43: 75â A¸S86 Trigo-Rodríguez, J. M. & Blum, J. 2009, Planetary and Space Science 57.2: 243-249. Tsuchiyama, A. & Mashio, E. & Imai, Y. & Noguchi, T. & Miura, Y. & Yano, H. & Nakamura, T., 2009, Meteoritics and Planetary Science Supplement, 72, 5189 Wurm, G. ,& Krauss, O. 2006, Phys. Rev. Lett., 96, 134301 Wurm, G. , 2007, Monthly Notices of the Royal Astronomical Society, 380, 683-690 Yang, C.-C. ,& Johansen, A. ,& Carrera, D. 2016, arXiv preprint arXiv:1611.07014 (2016).
1412.6230
3
1412
2015-01-31T12:04:11
Using the Inclinations of Kepler Systems to Prioritize New Titius-Bode-Based Exoplanet Predictions
[ "astro-ph.EP" ]
We analyze a sample of multiple-exoplanet systems which contain at least 3 transiting planets detected by the Kepler mission ("Kepler multiples"). We use a generalized Titius-Bode relation to predict the periods of 228 additional planets in 151 of these Kepler multiples. These Titius-Bode-based predictions suggest that there are, on average, ~2 planets in the habitable zone of each star. We estimate the inclination of the invariable plane for each system and prioritize our planet predictions by their geometric probability to transit. We highlight a short list of 77 predicted planets in 40 systems with a high geometric probability to transit, resulting in an expected detection rate of ~15%, ~3 times higher than the detection rate of our previous Titius-Bode-based predictions.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 19 (2015) Printed 24 January 2020 (MN LATEX style file v2.2) Using the Inclinations of Kepler Systems to Prioritize New Titius-Bode-Based Exoplanet Predictions T. Bovaird1,2(cid:63), C. H. Lineweaver1,2,3 and S. K. Jacobsen4 1Research School of Astronomy and Astrophysics, Australian National University, Canberra, ACT 2611, Australia 2Planetary Science Institute, Australian National University 3Research School of Earth Sciences,Australian National University 4Niels Bohr Institute, University of Copenhagen, Copenhagen, Denmark 24 January 2020 ABSTRACT We analyze a sample of multiple-exoplanet systems which contain at least 3 transiting plan- ets detected by the Kepler mission ('Kepler multiples'). We use a generalized Titius-Bode relation to predict the periods of 228 additional planets in 151 of these Kepler multiples. These Titius-Bode-based predictions suggest that there are, on average, 2± 1 planets in the habitable zone of each star. We estimate the inclination of the invariable plane for each system and prioritize our planet predictions by their geometric probability to transit. We highlight a short list of 77 predicted planets in 40 systems with a high geometric probability to transit, resulting in an expected detection rate of ∼ 15 per cent, ∼ 3 times higher than the detection rate of our previous Titius-Bode-based predictions. Key words: exoplanets, Kepler, inclinations, Titius-Bode relation, multiple-planet systems, invariable plane 1 INTRODUCTION The Titius-Bode (TB) relation's successful prediction of the pe- riod of Uranus was the main motivation that led to the search for another planet between Mars and Jupiter, e.g. Jaki (1972). This search led to the discovery of the asteroid Ceres and the rest of the asteroid belt. The TB relation may also provide useful hints about the periods of as-yet-undetected planets around other stars. In Bovaird & Lineweaver (2013) (hereafter, BL13) we used a gen- eralized TB relation to analyze 68 multi-planet systems with four or more detected exoplanets. We predicted the existence of 141 new exoplanets in these 68 systems. Huang & Bakos (2014) (here- after, HB14) performed an extensive search in the Kepler data for 97 of our predicted planets in 56 systems. This resulted in the con- firmation of 5 of our predictions. (Fig. 4 and Table 1). In this paper we perform an improved TB analysis on a larger sample of Kepler multiple-planet systems1 to make new exoplanet orbital period predictions. We use the expected coplanarity of multiple-planet systems to estimate the most likely inclination of the invariable plane of each system. We then prioritize our original and new TB-based predictions according to their geometric prob- ability of transiting. Comparison of our original predictions with the HB14 confirmations shows that restricting our predictions to those with a high geometric probability to transit should increase the detection rate by a factor of ∼ 3 (Fig. 8). (cid:63) E-mail: [email protected] 1 Accessed November 4th, 2014: http://exoplanetarchive.ipac.caltech.edu/cgi- bin/TblView/nph-tblView?app=ExoTbls&config=cumulative c(cid:13) 2015 RAS As in BL13, our sample includes all Kepler multi-planet sys- tems with four or more exoplanets, but to these we add three-planet systems if the orbital periods of the system's planets adhere better to the TB relation than the Solar System (Eq. 4 of BL13). Using these criteria we add 77 three-planet systems to the 74 systems with four or more planets. We have excluded 3 systems: KOI-284, KOI-2248 and KOI-3444 because of concerns about false positives due to adjacent-planet period ratios close to 1 and close binary hosts (Lissauer et al. 2011; Fabrycky et al. 2014; Lillo-Box et al. 2014). We have also excluded the three-planet system KOI-593, since the period of KOI-593.03 was recently revised, excluding the system from our three-planet sample. Thus, we analyze 151 Kepler multiples, with each system containing 3, 4, 5 or 6 planets. 1.1 Coplanarity of exoplanet systems Planets in the Solar System and in exoplanetary systems are be- lieved to form from protoplanetary disks (e.g. Winn & Fabrycky (2014)). The inclinations of the 8 planets of our Solar System to the invariable plane are (in order from Mercury to Neptune) 6.3◦, 2.2◦, 1.6◦, 1.7◦, 0.3◦, 0.9◦, 1.0◦, 0.7◦ (Souami & Souchay 2012). Jupiter and Saturn contribute ∼ 86 per cent of the total plane- tary angular momentum and thus the angles between their orbital planes and the invariable plane are small: 0.3◦ and 0.9◦ respec- tively. In a given multiple-planet system, the distribution of mutual inclinations between the orbital planes of planets is well described by a Rayleigh distribution (Lissauer et al. 2011; Fang & Margot 2012; Figueira et al. 2012; Fabrycky et al. 2014; Ballard & John- 5 1 0 2 n a J 1 3 . ] P E h p - o r t s a [ 3 v 0 3 2 6 . 2 1 4 1 : v i X r a 2 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 1. Panel a): Our coordinate system for transiting exoplanets. The x-axis points towards the observer. (cid:126)Lj is the 3-D angular momentum of the jth planet, and is perpendicular to the orbital plane of the jth planet. (cid:126)(cid:104)L(cid:105) is the sum of the angular momenta of all detected planets (Eq. A2) and is perpendicular to the invariable plane of the system. We have chosen the coordinate system without loss of generality such that (cid:126)(cid:104)L(cid:105) has no component in the y direction. φ j is the angle between (cid:126)Lj and (cid:126)(cid:104)L(cid:105). Let (cid:126)L j be the projection of (cid:126)Lj onto the x− z plane. ∆θ j is the angle between (cid:126)(cid:104)L(cid:105) and (cid:126)L j. (cid:104)θ(cid:105) is the angle between the z axis and (cid:126)(cid:104)L(cid:105). i j is the inclination of the planet (Eq. 3). θ j = 90− i j and is the angle between the z axis and (cid:126)L j such that θ j = (cid:104)θ(cid:105) + ∆θ j (Eq. A4). Panel b) shows the x− z plane of Panel a) with the y axis pointing into the paper. The observer is to the right. The grey shaded region represents the 'transit region' where the centre of a planet will transit its host star as seen by the observer (impact parameter b (cid:54) 1, see Eq. 1). Four planets, b, c, d and j, are represented by blue dots. The intersection of the orbital plane of the jth planet with the x− z plane is shown (thin black line). All angles shown in Panel a) (with the exception of φ j) are also shown in Panel b. The thick line is the intersection of the invariable plane of the system with the x− z plane. Because we are only dealing with systems with multiple transiting planets, all these angles are typically less than a few degrees but are exaggerated here for clarity. 1.5◦. We use this angle to determine the probability of detecting additional transiting planets in each system. Estimates of the inclination of a transiting planet come from the impact parameter b which is the projected distance between the center of the planet at mid-transit and the center of the star, in units of the star's radius. b = cosi (1) a R∗ Figure 2. The coplanarity of planets in the Solar System relative to the invariable plane. With the exception of Mercury, the angles between the orbital planes of the planets and the invariable plane are well represented by a Rayleigh distribution with a mode of ∼ 1◦. son 2014). For the ensemble of Kepler multi-planet systems, the mode of the Rayleigh distribution of mutual inclinations (φ j − φi) is typically ∼ 1◦ − 3◦ (Appendix A1 & Table A1). Thus, Kepler multiple-planet systems are highly coplanar. The Solar System is similarly coplanar. For example, the mode of the best fit Rayleigh distribution of the planet inclinations (φ j) relative to the invariable plane in the Solar System is ∼ 1◦ (see Fig. 1 and 2)2. The angle ∆θ j is a Gaussian distributed variable with a mean of 0 (centered around (cid:104)θ(cid:105)) and standard deviation σ∆θ . Based on previous analyses (Table A1), we assume the typical value σ∆θ = √ 2 See Appendix A for an explanation of why the distribution of mutual 2 wider than the distribution of the inclinations is on average a factor of angles φ j in Fig. 1, between the invariable plane of the system and the orbital planes of the planets. where R∗ is the radius of the star and a is the semi-major axis of the planet. For edge-on systems, typically 85◦ < i < 95◦. However, since we are unable to determine whether b is in the positive z direction or the negative z direction (Fig.1b and Fig. B1), we are unable to determine whether i is greater than or less than 90◦. By convention, for transiting planets the sign of b is taken as positive and thus the corresponding i values from Eq. 1 are taken as i (cid:54) 90◦. The impact parameter is also a function of four transit light curve observables (Seager & Mallén-Ornelas 2003); the period P, the transit depth ∆F, the total transit duration tT , and the total tran- sit duration minus the ingress and egress times tF (the duration where the light curve is flat for a source uniform across its disk). Thus the impact parameter can be written, b = f (P,∆F,tT ,tF ). (2) Eliminating b from Eqs. (1) and (2) yields the inclination i as a function of observables, i = cos−1 f (P,∆F,tT ,tF ) (3) (cid:20) R∗ a (cid:21) From Eq. 1 we can see that for an impact parameter b = 0 (a transit through the center of the star), we obtain i = 90◦; an 'edge- on' transit. The convention i (cid:54) 90◦ is unproblematic when only a single planet is found to transit a star but raises an issue when multiple exoplanets transit the same star, since the degree of coplanarity depends on whether the actual values of i j ( j = 1,2, ..N where N c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 θjR*icbdj024681012 2101 Distance from host star [RO •] R*/RO •〈θ〉b)a)To observerjTo observerzyxzθ〈θ〉LLxiφL〈L〉〈L〉jjjjj∆θjj∆θj Kernel density estimateRayleigh 1°Rayleigh 1.5°01234567 012 Inclination to invariable plane φ Number of planets Inclinations and Titius-Bode Predictions of Kepler Systems 3 is the number of planets in the system) are greater than or less than 90◦. For example, the actual values of i j in a given system could be all > 90◦, all < 90◦ or some in-between combination. Although we do not know the signs of θ j = 90◦ − i j for individual planets, we can estimate the inclination of the invariable plane for each system, by calculating all possible permutations of the θ j values for each system (see Appendix A2). In this estimation, we use the plausible assumption that the coplanarity of a system should not depend on the inclination of the invariable plane relative to the observer. 1.2 The probability of additional transiting exoplanets We wish to develop a measure of the likelihood of additional transiting planets in our sample of Kepler multi-planet systems. The more edge-on a planetary system is to an observer on Earth, the greater the probability of a planet transiting at larger periods. Similarly, a larger stellar radius leads to a higher probability of additional transiting planets (although with a reduced detection efficiency). We quantify these tendencies under the assumption that Kepler multiples have a Gaussian opening angle σ∆θ = 1.5◦ around the invariable plane, and we introduce the variable, acrit. Planets with a semi-major axis greater than acrit have less than a 50% geometric probability of transiting. More specifically, acrit is defined as the semi-major axis where Ptrans(acrit) = 0.5 (Eq. B1) for a given system. In a given system, a useful ratio for estimating the amount of semi-major axis space where additional transiting planets are more likely, is acrit/aout, where aout is the semi-major axis of the detected planet in the system which is the furthest from the host star. The larger acrit/aout, the larger the semi-major axis range for additional transiting planets beyond the outermost detected planet. Values for this ratio less than 1 mean that the outermost detected planet is beyond the calculated acrit value, and imply that addi- tional transiting planets beyond the outermost detected planet are less likely. Figure 3 shows the acrit/aout distribution for all systems in our sample. The fact that this distribution is roughly symmetric around acrit/aout ∼ 1 strongly suggests that the outermost tran- siting planets in Kepler systems are due to the inclination of the system to the observer, and are not really the outermost planets. In Section 2 we discuss the follow-up that has been done on our BL13 planet detections. In Section 3 we show that the ∼ 5% follow-up detection rate of HB14 is consistent with selection ef- fects and the existence of the predicted planets. In Section 4 we extend and upgrade the TB relation developed in BL13 and predict the periods of undetected planets in our updated sample. We then prioritize these predictions based on their geometric probability to transit and emphasize for further follow-up a subset of predictions with high transit probabilities. We also use TB predictions to esti- mate the average number of planets in the circumstellar habitable zone. In Section 5 we discuss how our predicted planet insertions affect the period ratios of adjacent planets and explore how period ratios are tightly dispersed around the mean period ratio within each system. In Section 6 we summarize our results. 2 FOLLOW-UP OF BL13 PREDICTIONS BL13 used the approximately even logarithmic spacing of exo- planet systems to make predictions for the periods of 117 addi- tional candidate planets in 60 Kepler detected systems, and 24 additional predictions in 8 systems detected via radial velocity(7) c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 Figure 3. Histogram of the acrit/aout values for our sample of Kepler mul- tiples. The distribution peaks for a ratio just below 1. Approximately half of the systems lie to the right of 1. For these systems, if there are plan- ets with semi-major axes a such that aout < a < acrit then the geometric probability of them transiting is greater than 50%. Note that this does not account for the detectability of these planets (e.g. they could be too small to detect). The majority of the predicted planets that are insertions (a < aout) have geometric transit probabilities greater than 50 per cent, when acrit/aout < 1. Systems on the right have more room for detections, and in general, predicted planets in these systems have higher values of Ptrans. The blue curve is the expected distribution of our sample of Kepler multiples if they all have planets at TB-predicted semi-major axes extrap- olated out to ∼ 4 × acrit. The blue curve is consistent with the observed distribution, indicating that our acrit/aout distribution is consistent with the system in our sample containing more planets than have been detected. Table 1. Systems with candidate detections by HB14 (in bold) plus KOI-1151 (Petigura et al. 2013a) and KOI-1860 b, after planet pre- dictions were made by BL13. System Predicted Detected Predicted Detected Period (days) Period (days) Radius (R⊕) Radius (R⊕) KOI-719 KOI-1336 KOI-1952 KOI-2722a KOI-2859 KOI-733 KOI-1151a KOI-1860b 14± 2 26± 3 13± 2 16.8± 1.0 5.2± 0.3 N/A 9.6± 0.7 25± 3 15.77 27.51 13.27 16.53 5.43 15.11 10.43 24.84 (cid:54) 0.7 (cid:54) 2.4 (cid:54) 1.5 (cid:54) 1.6 (cid:54) 0.8 N/A (cid:54) 0.8 (cid:54) 2.7 0.42 1.04 0.85 1.16 0.76 3.0 0.7 1.46 a Predicted by preprint of BL13 (draft uploaded 11 Apr 2013: http://arxiv.org/pdf/1304.3341v1.pdf), detected planet reported by Ke- pler archive and included in analysis of BL13. b October 2014 Kepler Archive update, during the drafting of this paper. and direct imaging(1), which we do not consider here. NASA Ex- oplanet Archive data updates, confirmed our prediction of KOI- 2722.05 (Table 1). HB14 used the planet predictions made in BL13 to search for 97 planets in the light curves of 56 Kepler systems. Within these 56 systems, BL13 predicted the period and maximum radius: the largest radius which would have evaded detection, based on the lowest signal-to-noise of the detected planets in the same system. Predicted planets were searched for using the Kepler Quarter 1 to Quarter 15 long cadence light curves, giving a baseline exceeding 1000 days. Once the transits of the already known planets were de- tected and removed, transit signals were visually inspected around the predicted periods. Of the 97 predicted planets searched for by HB14, 5 candi- dates were detected within ∼ one-sigma of the predicted periods LessroomMore room for detections01234 010203040 acrit/aout Number of systems 4 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 4. Exoplanet systems where an additional candidate was detected after a TB relation prediction was made (see Table 1). The systems are shown in descending order of acrit/aout. Previously known planets are shown as blue circles. The predictions of BL13 and their uncertainties are shown by the red filled rectangles if the Ptrans value of the predicted planet is (cid:62) 0.55 (Eq. B1 and Fig. 8), or by red hatched rectangles otherwise. The new candidate planets are shown as green squares. The critical semi-major axis acrit (Section 1.2), beyond which Ptrans(acrit) < 0.5, is shown by a solid black arc. The uncertainties (width of red rectangles) in this figure and Table 1, are slightly wider than Figure 5 due to excessive rounding of predicted period uncertainties for some systems in BL13. Figure 5. Same as Fig. 4 except here our TB predictions are based on γ with n2 ins (Eq. 9) rather than on the γ with nins of Eq. 5 of BL13. Comparing Fig. 4 with this figure, in the KOI-719 system the number of predicted planets goes from 4 to 2, while in KOI-1151 the number of predicted planets goes from 3 to 2. In both cases the detected planet is more centrally located in the predicted region. c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 KOI-719KOI-1336KOI-1952KOI-2722KOI-2859KOI-733KOI-1151KOI-18600.00.10.20.30.40.50.00.10.20.30.40.5 Semi-Major Axis [AU] Mercury Earth Uranus Saturn KOI-719KOI-1336KOI-1952KOI-2722KOI-2859KOI-733KOI-1151KOI-18600.00.10.20.30.40.50.00.10.20.30.40.5 Semi-Major Axis [AU] Mercury Earth Uranus Saturn Inclinations and Titius-Bode Predictions of Kepler Systems 5 Figure 6. Reported dip significance parameter (DSP) from Table 1 of Huang & Bakos (2014) for previously known exoplanets in the five sys- tems with a new detection. A linear trend can be seen for the DSP and the signal-to-noise ratio as reported by the Kepler team (Christiansen et al. 2012). HB14 required planet candidates to have a DSP > 8 to survive their vetting process. The sizes of the blue dots correspond to the same planetary radii representation used to make Figs. 11, 12, 13 & 14. (5 planets of the 6 planets in bold in Table 1, see also Fig. 4). No- tably all new planet candidates have Earth-like or lower planetary radii. One additional candidate was detected in KOI-733 which is incompatible with the predictions of BL13. This candidate is unique in that it should have been detected previously, based on the signal-to-noise of the other detected planets in KOI-733. In Table 1, the detected radii are less than the maximum predicted radii in each case. The new candidate in KOI-733 has a period of 15.11 days and a radius of 3 R⊕. At this period, the maximum ra- dius to evade detection should have been 2.2 R⊕. With the possible exception of KOI-1336 where a dip significance parameter (DSP, Kovács & Bakos (2005)) was not reported, all detected candidates have a DSP of (cid:62) 8, which roughly corresponds to a Kepler SNR (Christiansen et al. 2012) of (cid:38) 12 (see Figure 6). HB14 required DSP > 8 for candidate transit signals to survive their vetting pro- cess. 3 IS A 5% DETECTION RATE CONSISTENT WITH SELECTION EFFECTS? From a sample of 97 BL13 predictions, HB14 confirmed 5. How- ever, based on this ∼ 5% detection rate, HB14 concluded that the predictive power of the TB relation used in BL13 was question- able. Given the selection effects, how high a detection rate should one expect? We do not expect all planet predictions to be detected. The predicted planets may have too large an inclination to tran- sit relative to the observer. Additionally, there is a completeness factor due to the intrinsic noise of the stars, the size of the plan- ets, and the techniques for detection. This completeness for Ke- pler data has been estimated for the automated lightcurve analy- sis pipeline TERRA (Petigura et al. 2013a). Fig. 7 displays the TERRA pipeline injection/recovery completeness. After correct- ing for the radius and noise of each star, relative to the TERRA sample in Fig. 7, the planet detections in Table 1 have an average detection completeness in the TERRA pipeline of ∼ 24%. That is, if all of our predictions were correct and if all the planets were in approximately the same region of period and radius space as the green squares in Fig. 7, and if all of the planets transited, we would expect a detection rate of ∼ 24% using the TERRA pipeline. It is unclear how this translates into a detection rate for a manual inves- tigation of the lightcurves motivated by TB predictions. We wish to determine, from coplanarity and detectability ar- guments, how many of our BL13 predictions we would have ex- pected to be detected. An absolute number of expected detections is most limited by the poorly known planetary radius distribution c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 Figure 7. The simulated detection completeness of the new candidate plan- ets in the TERRA pipeline (modified from Figure 1 of Petigura et al. (2013a)). Here we overplot as green squares, the 8 planets listed in Table 1. The completeness curves are averaged over all stellar noise and stellar radii in the Petigura et al. (2013a) sample (the 42,000 least noisy Kepler stars). Green circles indicate the 'effective radius' of the new candidates, based on the noise and radius of their host star in comparison to the median of the quietest 42,000 sample. From the signal-to-noise, the effective radius can be calculated by Rp,eff/Rp = (R∗/R∗,median) × (CDPP/CDPPmedian)1/2, where CDPP is the combined differential photometric precision defined in Christiansen et al. (2012). Taking a subset of 42,000 stars from the Ke- pler input catalog with the lowest 3-hour CDPP (approximately represen- tative of the sample in the figure), we obtain CDPPmedian ≈ 60 ppm and R∗,median ≈ 1.15 R(cid:12). Using the effective radius and excluding the outlier KOI-733, the mean detection completeness for the 7 candidate planets in Table 1 is ∼ 24%. below 1 Earth radius (Howard et al. 2012; Dressing & Charbon- neau 2013; Dong & Zhu 2013; Petigura et al. 2013b; Fressin et al. 2013; Silburt et al. 2014; Morton & Swift 2014; Foreman-Mackey et al. 2014). Large uncertainties about the shape and amplitude of the planetary radius distribution of rocky planets with radii less than 1 Earth radius make the evaluation of TB-based exoplanet predictions difficult. Since the TB relation predicted the asteroid belt (Masteroid < 10−3MEarth) there seems to be no lower mass limit to the objects that the TB relation can predict. This makes estimation of the detection efficiencies strongly dependent on as- sumptions about the frequency of planets at small radii. Let the probability of detecting a planet, Pdetect, be the product of the geometric probability to transit Ptrans as seen by the observer (Appendix B) and the probability PSNR that the planetary radius is large enough to produce a signal-to-noise ratio above the detection threshold, Pdetect = Ptrans PSNR. (4) The geometric probability to transit, Ptrans, is defined in Eq. B1 and illustrated in Fig. 1 and Fig. B1. The 5 confirmations from our previous TB predictions are found in systems with a much higher than random probability of transit (Fig. 8). This is expected if our estimates of the invariable plane are reasonable. To estimate PSNR we first estimate the probability that the radius of the planet will be large enough to detect. In BL13 we 010203040506070 020406080100 DSP Kepler SNR51020304050100200300400Orbital period (days)0.4123451020Planet size (Earth-radii)0102030405060708090100Survey Completeness (C) % 6 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 8. Histogram of Ptrans (Eq. B1), the geometric probability of transit, for the 97 predicted planets from BL13, that were followed up by HB14. The blue histogram represents the 5 new planets detected by HB14 (Ta- ble 1). As expected, the detected planets have high Ptrans values compared to the entire sample. The red and blue solid lines represent the empirical cumulative distribution function for the two distributions. A K-S test of the two distributions yields a p-value of 1.8× 10−2. Thus, Ptrans can be used to prioritize our predictions and increase their probability of detection. Figure 9. The assumed distribution of planetary radii described in Sec- tion 3. The distribution is poorly constrained below 1 Earth radii, indicated by the gray dashed line. For low mass stars, the planetary radius distribu- tion may decline below 0.7 R⊕ (Dressing & Charbonneau 2013; Morton & Swift 2014). Alternative estimations show the planetary radius distribution continuing to increase with smaller radii (continuing the flat logarithmic distribution), down to 0.5 R⊕ (Foreman-Mackey et al. 2014). For our anal- ysis we have extrapolated the flat distribution (in log R) down to Rlow. We indicate three regions for a hypothetical system at a specific predicted pe- riod. The 'already detected' region refers to the range of planetary radii which should already have been detected, based on the lowest signal-to- noise ratio of the detected planets in that system. Rmin is the smallest radius which could produce a transit signal that exceeds the detection threshold, and is the boundary between the undetectable and detectable regions. estimated the maximum planetary radius, Rmax, for a hypothetical undetected planet at a given period, based on the lowest signal- to-noise of the detected planets in the same system. We now wish to estimate a minimum radius that would be detectable, given the individual noise of each star. We refer to this parameter as Rmin, which is the minimum planetary radius that Kepler could detect around a given star (using a specific SNR threshold). For each star we used the mean CDPP (combined differential photometric pre- cision) noise from Q1-Q16. When the number of transits is not reported, we use the approximation Ntrans ≈ Tobs f0/P, where Tobs is the total observing time and f0 is the fractional observing up- time, estimated at ∼0.92 for the Kepler mission (Christiansen et al. 2012). The probability PSNR depends on the underlying planetary ra- dius probability density function. We assume a density function of Figure 10. The mean detection rate Pdetect (Eq. 4) of the BL13 predictions is dependent on the transit signal-to-noise threshold (SNRth in Eq. 8) used in lightcurve vetting (x-axis), and on the probability of low-radii planets, i.e. on how far the flat logarithmic planetary radius distribution in Fig. 9 should be extrapolated (Rlow on the y-axis). For example, in the denomi- nator of Eq. 6, integrating down to a radius Rlow of 0.6 R⊕ and setting a SNR threshold of 7 (which sets Rmin in the numerator) gives an expected detection fraction of ∼ 25% (blue values in the upper left of the plot). Inte- grating down to a radius of 0.2 R⊕ and having a DSP threshold of 8 (con- verted from SNR according to Fig. 6) gives an expected detection fraction of ∼ 5% (red values in the lower right). the form: f (R) = d f dlogR = (cid:40) R (cid:62) 2.8 R⊕ k(logR)α , k(log2.8)α , R < 2.8 R⊕ (5) where k = 2.9 and α = −1.92 (Howard et al. 2012). The dis- continuous distribution accounts for the approximately flat num- ber of planets per star in logarithmic planetary radius bins for R (cid:46) 2.8 R⊕ (Dong & Zhu 2013; Fressin et al. 2013; Petigura et al. 2013b; Silburt et al. 2014). For R (cid:46) 1.0 R⊕ the distribution is poorly constrained. For this paper, we extend the flat distribu- tion in log R down to a minimum radius Rlow = 0.3 R⊕. It is im- portant to note that for the Solar System, the poorly constrained part of the planetary radius distribution contains 50 per cent of the planet population. For reference the radius of Ceres, a "planet" predicted by the TB relation applied to our Solar System has a radius RCeres = 476 km = 0.07R⊕. The probability that the hypothetical planet has a radius that exceeds the SNR detection threshold is then given by (cid:82) Rmax (cid:82) Rmax Rmin Rlow PSNR = f (R)dR f (R)dR (6) We do not integrate beyond Rmax since we expect a planet with a radius greater than Rmax would have already been detected. We define Rmax by, Rmax = Rmin SNR , (7) where Rmin SNR and Pmin SNR are the radius and period respectively of the detected planet with the lowest signal-to-noise in the system. Ppredict is the period of the predicted planet. Rmin depends on the SNR in the following way: (cid:19)1/4 (cid:18) Ppredict Pmin SNR Rmin = R∗(cid:112)SNRth CDPP (cid:18) 3hrs (cid:19)1/4 ntr tT , (8) where SNRth is the SNR threshold for a planet detection, ntr is the number of expected transits at the given period and tT is the c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 0.0 0.2 0.4 0.6 0.8 1.0 051015200%20%40%60%80%100% Ptrans Number of predicted planetsMarsMercury RlowRmax-0.50.00.51.01.5 110Log planet radius [R⊕] 0.00.20.40.60.8 f (R)undetectable detectable already detected already detected Relative frequency?0.1-1.0RminCeres 7891011121314150.10.20.30.40.50.6 5%10%15%20%25%56789 Transit SNR threshold Rlow [R⊕]Transit DSP thresholdP detect5%10%15%20%3%HB14 Inclinations and Titius-Bode Predictions of Kepler Systems 7 transit duration in hours. See Figure 9 for an illustration of how the integrals in PSNR (Eq. 6) depend on the planet radii limits, Rmin and Rlow. While Ptrans is well defined, PSNR is dependent on the SNR threshold chosen (SNRth), the choice of Rlow and the poorly con- strained shape of the planetary radius distribution below 1 Earth radius. This is demonstrated in Figure 10, where the mean Pdetect from the predictions of BL13 (for DSP > 8) can vary from ∼ 2 per cent to ∼ 11 per cent. Performing a K-S test on PSNR values (analogous to that in Figure 8) indicates that the PSNR values for the subset of our BL13 predictions that were detected, are drawn from the same PSNR distribution as all of the predicted planets. For this reason we use only Ptrans, the geometric probability to tran- sit, to prioritize our new TB relation predictions. We emphasize a subset of our predictions which have a Ptrans value (cid:62) 0.55, since all of the confirmed predictions of BL13 had a Ptrans value above this threshold. Only ∼ 1/3 of the entire sample have Ptrans values this high. Thus, the ∼ 5 per cent detection rate should increase by a factor of ∼ 3 to ∼ 15 per cent for our new high-Ptrans subset of planet period predictions. 4 UPDATED PLANET PREDICTIONS 4.1 Method and Inclination Prioritization We now make updated and new TB relation predictions in all 151 systems in our sample. If the detected planets in a system adhere to the TB relation better than the Solar System planets (χ2/d.o.f. < 1.0, Equation 4 of BL13), we only predict an extrapo- lated planet, beyond the outermost detected planet. If the detected planets adhere worse than the Solar System, we simulate the in- sertion of up to 9 hypothetical planets into the system, covering all possible locations and combinations, and calculate a new χ2/d.o.f. value for each possibility. We determine how many planets to in- sert, and where to insert them, based on the solution which im- proves the system's adherence to the TB relation, scaled by the number of inserted planets squared. This protects against overfit- ting (inserting too many planets, resulting in too good a fit). In Eq. 5 of BL13 we introduced a parameter γ, which is a measure of the fractional amount by which the χ2/d.o.f. improves, divided by the number of planets inserted. Here, we improve the definition of γ by dividing by the square of the number of planets inserted, (cid:18) χ2 (cid:19) f i −χ2 χ2 f n2 ins γ = (9) i and χ2 where χ2 f are the χ2 of the TB relation fit before and af- ter planets are inserted respectively, while nins is the number of inserted planets. Importantly, when we calculate our γ value by dividing by the number of inserted planets squared, rather than the number of planets, we still predict the BL13 predictions that have been detected. In two of these systems fewer planets are predicted and as a result the new predictions agree better with the location of the detected candidates. This can be seen by comparing Figures 4 and 5. We compute Ptrans for each planet prediction in our sample of 151 Kepler systems. We emphasize the 40 systems where at least one inserted planet in that system has Ptrans (cid:62) 0.55. Period predictions for this subset of 40 systems are displayed in Table C1 and Figures 11, 12 and 13. As discussed in the previous section, we expect a detection rate of ∼ 15 per cent for this high-Ptrans c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 sample. Predictions for all 228 planets (regardless of their Ptrans value) are shown in Table C2 (where the systems are ordered by the maximum Ptrans value in each system). 4.2 Average Number of Planets in Circumstellar Habitable Zones Since the search for earth-sized rocky planets in circumstellar hab- itable zones (HZ) is of particular importance, in Fig. 14, for a subset of Kepler multiples whose predicted (extrapolated) planets extend to the HZ, we have converted the semi-major axes of de- tected and predicted planets into effective temperatures (as in Fig. 6 of BL13). One can see in Fig. 14 that the habitable zone (shaded green) contains between 0 and 4 planets. Thus, if the TB relation is approximately correct, and if Kepler multi-panet systems are rep- resentative of planetary systems in general, there are on average ∼ 2 habitable zone planets per star. More specifically, in Table 2, we estimate the number of plan- ets per star in various 'habitable zones', namely (1) the range of Teff between Mars and Venus (assuming an albedo of 0.3), dis- played in Fig.14 as the green shaded region, (2) the Kopparapu et al. (2013) "optimistic" and (3) "conservative" habitable zones ("recent Venus" to "early Mars", and "runaway greenhouse" to "maximum greenhouse" respectively). We find, on average, 2± 1 planets per star in the "habitable zone", almost independently of which of the 3 habitable zones one is referring to. Using our es- timates of the maximum radii for these predominantly undetected (but predicted) planets, as well as the planetary radius distribu- tion of Fig. 9, we estimate that on average, ∼ 1/6 of these ∼ 2 planets, or ∼ 0.3, are 'rocky'. We have assumed that planets with R (cid:54) 1.5R⊕ are rocky (Rogers 2014; Wolfgang & Lopez 2014). 5 ADJACENT PLANET PERIOD RATIOS HB14 concluded that the percentage of detected planets (∼ 5%) was on the lower side of their expected range (∼ 5%− 20%) and that the TB relation may over-predict planet pairs near the 3:2 mean-motion resonance (compared to systems which adhered to the TB relation better than the Solar System, without any planet insertions. i.e. χ2/d.o.f (cid:54) 1). There is some evidence that a peak in the distribution of period ratios around the 3:2 resonance is to be expected from Kepler data, after correcting for incompleteness Steffen & Hwang (2014). In this section we investigate the period ratios of adjacent planets in our Kepler multiples before and after our new TB relation predictions are made. We divide our sample of Kepler multiples into a number of subsets. Our first subset includes systems which adhere to the TB relation better than the Solar System (where we only predict an ex- trapolated planet beyond the outermost detected planet). Systems which adhere to the TB relation worse than the Solar System we divide into two subsets, before and after the planets predicted by the TB relation were inserted. Adjacent planet period ratios can be misleading if there is an undiscovered planet between two detected planets, which would reduce the period ratios if it was included in the data. To minimize this incompleteness, we also construct a subset of systems which are the most likely to be completely sam- pled (unlikely to contain any additional transiting planets within the range of the detected planet periods). Systems which adhere to the TB relation better than the Solar System (χ2/d.o.f (cid:54) 1) were considered by HB14 as being the sam- ple of planetary systems that were most complete and therefore 8 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 11. The architectures and invariable plane inclinations for Kepler systems in our sample which contain at least one planet with a geometric probability to transit Ptrans (cid:62) 0.55. There are 40 such systems out of the 151 in our sample. The 14 with the highest Ptrans values are plotted here. The remaining 26 are plotted in the next two figures. The order of the systems (from top to bottom) is determined by the highest Ptrans value in each system. The thin horizontal dotted line represents the line-of-sight to Earth, i.e. where the i value of a planet would be 90◦. The thick grey line in each system is our estimate of the invariant plane angle, (cid:104)θ(cid:105) (Appendix A2). The value of (cid:104)θ(cid:105) is given in degrees to the right of each panel (see also Fig. A1). The green wedge has an opening angle σ∆θ = 1.5◦ and is symmetric around the invariable plane, but is also limited to the grey region where a planet can be seen to transit from Earth (b (cid:54) 1, Eq. 1). The thick black arc indicates the acrit value beyond which less than 50 per cent of planets will transit (Eq. B1). Predicted planets and their uncertainties are shown by solid red rectangles if the Ptrans value of the predicted planet is (cid:62) 0.55, or by red hatched rectangles otherwise. Thus, the 77 solid red rectangles in the 40 systems shown in Figs. 11,12 & 13 make up our short list of highest priority predictions (Table C1). Our estimate of the most probable inclination ambiguities in a system are represented by vertically separated pairs of blue dots, connected by a thin black line (see Appendix A2). c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 〈θ〉0.4KOI-11980.8KOI-19550.6KOI-10820.5KOI-9521.6KOI-5002.7KOI-40321.2KOI-7071.0KOI-13364.1KOI-28590.3KOI-2500.3KOI-1680.5KOI-25850.3KOI-10521.3KOI-5050.00.10.20.30.40.50.00.10.20.30.40.5 Semi-Major Axis [AU] Mercury Earth Uranus Saturn Inclinations and Titius-Bode Predictions of Kepler Systems 9 Figure 12. The same as Figure 11 but for the next 16 systems in our sample which have at least one planet with Ptrans (cid:62) 0.55. Note that some of the new detected planets from the predictions of BL13 have been included in the Kepler data archive (see Table 1), and that these planets are included in our analysis. c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 〈θ〉0.3KOI-18311.5KOI-2480.3KOI-8800.3KOI-15670.8KOI-19520.0KOI-3510.2KOI-7010.3KOI-13060.5KOI-27220.4KOI-13580.5KOI-16270.8KOI-18330.7KOI-31581.2KOI-20550.5KOI-2451.5KOI-7490.00.10.20.30.40.50.00.10.20.30.40.5 Semi-Major Axis [AU] Mercury Earth Uranus Saturn 10 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 13. The same as Figures 11 & 12 but for the remaining 10 systems in our sample which have at least one planet with Ptrans (cid:62) 0.55. had a distribution of adjacent planet period ratios most representa- tive of actual planetary systems. However, the choice of BL13 to normalize the TB relation to the Solar System's χ2/d.o.f is some- what arbitrary. The Solar System's χ2/d.o.f is possibly too high to consider all those with smaller values of χ2/d.o.f to be completely sampled. We want to find a set of systems which are unlikely to host any additional planets between adjacent pairs, due to the system being dynamically full (Hayes & Tremaine 1998). We do this by identifying the systems where two or more sequential planet pairs are likely to be unstable when a massless test particle is inserted between each planet pair (dynamical spacing ∆ < 10, Gladman (1993), BL13). The dynamical spacing ∆ is an estimate of the stability of ad- jacent planets. If inserting a test particle between a detected planet pair results in either of the two new ∆ values being less than 10, we consider the planet pair without the insertion to be complete. That is, there is unlikely to be room, between the detected planet pair, where an undetected planet could exist without making the planet pair dynamically unstable. Therefore, since the existence of an undetected planet between the planet pair is unlikely, we refer to the planet pair as 'completely sampled'. Estimating completeness based on whether a system is dynamically full is a reasonable ap- proach, since there is some evidence that the majority of systems are dynamically full (e.g. Barnes & Raymond (2004)). For Ke- pler systems in particular, Fang & Margot (2013) concluded that at least 45 per cent of 4-planet Kepler systems are dynamically packed. If at least two sequential adjacent-planet pairs (at least three sequential planets) satisfy this criteria, we add the subset of the system which satisfies this criteria to our 'most complete' sam- ple. We use this sample to analyze the period ratios of Kepler sys- tems. The period ratios of the different samples described above are shown in Figure 15. One criticism from HB14 was that the TB relation from BL13 inserted too many planets. To address this criticism we have rede- fined γ to be divided by the number of inserted planets squared (denominator of Eq. 9). This introduces a heavier penalty for in- serting planets. Figure 15 displays the distributions of period ratios when using the γ from BL13 and the new γ of Equation 9 (panels b and c respectively). When using our newly defined γ, the mean c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 〈θ〉0.3KOI-7300.6KOI-7190.4KOI-10601.6KOI-30830.8KOI-1560.7KOI-1370.5KOI-11510.3KOI-10150.8KOI-20291.1KOI-6640.00.10.20.30.40.50.00.10.20.30.40.5 Semi-Major Axis [AU] Mercury Earth Uranus Saturn Inclinations and Titius-Bode Predictions of Kepler Systems 11 Figure 14. The effective temperatures of planets within the 31 systems from our sample which extend out to the green habitable zone (HZ) after our planet predictions are made. For the purpose of estimating the number of HZ planets per star (see Table 2), we extrapolate additional planets (gray squares) beyond the HZ. The sizes of the red hashed squares represent the Rmax of the predicted planet. Table 2. The estimated number of planets per star within various 'habitable zones'. All planets Sample All 151 systems Least extrapolationb Mars-Venus K13 "optimistic" K13 "conservative" 2.0± 1.0 1.6± 0.9 1.5± 0.8 1.3± 0.7 2.3± 1.2 1.7± 0.8 Rocky planets (R (cid:54) 1.5 R⊕) Mars-Venus K13 "optimistic" K13 "conservative" 0.15 0.40 0.15 0.35 0.10 0.30 a K13 "optimistic" and "conservative" habitable zones refer to the "recent Venus" to "early Mars" and "runaway greenhouse" to "maximum greenhouse' regions from Kopparapu et al. (2013) respectively'. b The 31 systems in the sample shown in Fig. 14 are those which need the least extrapolation (red hashed squares) to extend out to (or beyond) the green Mars-Venus HZ. c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 MerVenEarMarJupSunKOI-518KOI-435KOI-571KOI-701KOI-2183KOI-841KOI-1422KOI-1922KOI-1430KOI-1258KOI-564KOI-812KOI-612KOI-806KOI-351KOI-904KOI-775KOI-2926KOI-2714KOI-2073KOI-481KOI-719KOI-1895KOI-250KOI-899KOI-1426KOI-757KOI-620KOI-886KOI-952KOI-28210001001000100 Effective temperature [K]H Z Mercury Earth Uranus Saturn 12 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Figure 16. Each system has a mean value for the adjacent-planet period ratios within that system. This figure shows the distribution of the period ratios, offset from the mean period ratio of the system. The colors of the distributions correspond to subsets in Figure 15. The green distribution is from the 'most complete sample' in panel e) and is our best estimate of what a distribution should look like if an appropriate number of planets have been inserted. The gray distribution indicates that the sampling of these systems is highly incomplete. The thick and thin blue lines repre- sent panel c) in two different ways. The thick blue line uses planet peri- ods drawn randomly from a normal distribution centered on the periods predicted by the TB relation, with the width set to the uncertainty in the period predictions (Tables C1 and C2). The thin blue line uses periods at their exact predicted value. the incompleteness in Kepler multiple-planet systems and make predictions about the probable locations of the undetected planets. 6 CONCLUSION Huang & Bakos (2014) investigated the TB relation planet predic- tions of Bovaird & Lineweaver (2013) and found a detection rate of ∼ 5% (5 detections from 97 predictions). Apart from the detec- tions by HB14, only one additional planet (in KOI-1151) has been discovered in any of the 60 Kepler systems analyzed by BL13 -- indicating the advantages of such predictions while searching for new planets. Completeness is an important issue (e.g. Figures 7 & 10). Some large fraction of our predictions will not be detected be- cause the planets in this fraction are likely to be too small to pro- duce signal-to-noise ratios above some chosen detection thresh- old. Additionally, the predicted planets may have inclinations and semi-major axes too large to transit their star as seen from Earth. All new candidate detections based on the predictions of BL13 are approximately Earth-sized or smaller (Table 1). For a new sample of Kepler multiple-exoplanet systems con- taining at least three planets, we computed invariable plane incli- nations and assumed a Gaussian opening angle of coplanarity of σ∆θ = 1.5◦. For each of these systems we applied an updated gen- eralized TB relation, developed in BL13, resulting in 228 predic- tions in 151 systems. We emphasize the planet predictions which have a high geo- metric probability to transit, Ptrans (cid:62) 0.55 (Figure 8). This subset of predictions has 77 predicted planets in 40 systems. We expect the detection rate in this subset to be a factor of ∼ 3 higher than the detection rate of the BL13 predictions. From the 40 systems with planet predictions in this sample, 24 appeared in BL13. These pre- dictions have been updated and reprioritized. We have ordered our list of predicted planets based on each planet's geometric prob- ability to transit (Tables C1 and C2). Our new prioritized predic- tions should help on-going planet detection efforts in Kepler multi- planet systems. c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 Figure 15. The period ratios of adjacent-planet pairs in our sample of Ke- pler multiples, which can be compared to Figure 4 of HB14. Panel a) repre- sents systems which adhere to the TB relation better than the Solar System (χ2/d.o.f. < 1). Panel b) represents those systems which adhere to the TB relation worse than the Solar System and where BL13 inserted planets. This panel shows the period ratios of adjacent-planet pairs after planets are inserted. Panel c) is similar to Panel b), except that the γ value, which is used to determine the best TB relation insertion for a given system, is divided by the number of inserted planets squared. In BL13 and panel b), γ was divided by the number of inserted planets. Panel d) shows the pe- riod ratios between adjacent pairs of the same systems from panels b) and c), except before the additional planets from the predictions of BL13 have been inserted. Panel e) represents our most complete sample and contains the systems which are more likely to be dynamically full (as defined in Section 5). The mean χ2/d.o.f. value for each subset is shown on the left side of each panel. Kepler's bias toward detecting compact systems domi- nated by short period planets may explain why the Solar System's adjacent period pairs (black hatched histogram in panel e) are not representative of the histogram in panel e. The periods of predicted planets are drawn ran- domly from their TB relation predicted Gaussian distributions (Tables C1 and C2). χ2/d.o.f. (displayed on the left side of the panels), more closely resembles that of our 'most complete sample' (panel e). Since each panel in Fig. 15 represents a mixture of planetary systems with different distributions of period ratios, Figure 16 may be a better way to compare these different samples and their adher- ence to the TB relation. For each planetary system in each panel in Fig. 15, we compute the mean adjacent-planet period ratio. Figure 16 shows the distribution of the offsets from the mean period ratio of each system. How peaked a distribution is, is a good measure of how well that distribution adheres to the TB relation. A delta func- tion peak at an offset of zero, would be a perfect fit. The period ratios of adjacent-planet pairs in our dynamically full "most com- plete sample" (green in Fig. 16) displays a significant tendency to cluster around the mean ratios. This clustering is the origin of the usefulness of the TB relation to predict the existence of undetected planets. The proximity of the thick blue curve to the green distri- bution is a measure of how well our TB predictions can correct for 5:22:15:33:24:35:4Adheres to TBR betterthan Solar System(χ2/d.o.f. < 1.0):Meanχ2/d.o.f. = 0.48 051015 Na) Adheres to TBR worsethan Solar System:after insertions0.27γ ∝ nins-1 051015 Nb) Adheres to TBR worsethan Solar System:after insertions0.36γ ∝ nins-2 051015 Nc) Adheres to TBR worsethan Solar System:before insertions3.16 02468 Nd) 'Most complete' sample:Pairs with ∆ < 10 (afterinsertion of 1 M⊕planet between pair)0.401234 02468 Pouter/Pinner Ne) -1.0-0.50.00.51.00.000.050.100.15 Offset from mean period ratio Relative frequency Inclinations and Titius-Bode Predictions of Kepler Systems 13 ACKNOWLEDGEMENTS T.B. acknowledges support from an Australian Postgraduate Award. We acknowledge useful discussions with Daniel Bayliss, Michael Ireland, George Zhou and David Nataf. REFERENCES 549, 135, 1998, Icarus, Ballard S., Johnson J. A., 2014, ApJ (submitted), arXiv:1410.4192 Barnes R., Raymond S. N., 2004, ApJ, 617, 569, doi:10.1086/423419 Bovaird T., Lineweaver C. H., 2013, MNRAS, 435, 1126 Christiansen J. L. et al., 2012, PASP, 124, 1279 Dong S., Zhu Z., 2013, ApJ, 778, 53, doi:10.1088/0004-637X/778/1/53 Dressing C. D., Charbonneau D., 2013, ApJ, 767, 95, doi:10.1088/0004- 637X/767/1/95 Fabrycky D. C., Winn J. N., 2009, ApJ, 696, 1230 Fabrycky D. C. et al., 2014, ApJ, 790, 146 Fang J., Margot J.-L., 2012, ApJ, 761, 92 Fang J., Margot J.-L., 2013, ApJ, 767, 115 Figueira P. et al., 2012, A&A, 541, A139 Foreman-Mackey D., Hogg D. W., Morton T. D., . 2014, ApJ, 795, 64, doi:10.1088/0004-637X/795/1/64 Fressin F. et al., 2013, ApJ, 766, 81 Gladman B., 1993, Icarus, 106, 247 Hayes W., S., Tremaine doi:10.1006/icar.1998.5999 Howard A. W. et al., 2012, ApJS, 201, 15 Huang C. X., Bakos G. A., 2014, MNRAS, 681, 674 Jaki S. L., 1972, Am. J. Phys., 40, 1014 Johansen A., Davies M. B., Church R. P., Holmelin V., 2012, ApJ, 758, 39, doi:10.1088/0004-637X/758/1/39 Kopparapu R. K. et al., 2013, ApJ, 765, 131, doi:10.1088/0004- 637X/765/2/131 Kovács G., Bakos G., 2005, arXiv:0508081 Lillo-Box J., Barrado D., Bouy H., . 2014, A&A, 556, A103 Lissauer J. J. et al., 2011, ApJS, 197, 8 Morton T. D., Swift J., 2014, ApJ, 791, 10, doi:10.1088/0004- 637X/791/1/10 Petigura E. A., Howard A. W., Marcy G. W., . 2013a, PNAS, 110, 19273 Petigura E. A., Marcy G. W., Howard A., . 2013b, ApJ, 770, 69 Rogers L. a., 2014, ApJ (submitted), arXiv:1407.4457 Seager S., Mallén-Ornelas G., 2003, ApJ, 585, 1038 Silburt A., Gaidos E., Wu Y., . 2014, preprint (arXiv:1406.6048v2), arXiv:arXiv:1406.6048v2 Souami D., Souchay J., 2012, A&A, 543, A133 Steffen J. H., Hwang arXiv:arXiv:1409.3320v1 Tremaine S., Dong S., 2012, AJ, 143, 94, doi:10.1088/0004- 6256/143/4/94 Watson G. S., 1982, Journal of Applied Probability, 19, 265 Weissbein A., Steinberg E., 2012, arXiv:arXiv:1203.6072v2 Winn J. N., Fabrycky D. C., 2014, ARAA (submitted), arXiv:1410.4199 Wolfgang A., Lopez E., 2014, ApJ (submitted), arXiv:1409.2982 2014, MNRAS (submitted), J. A., APPENDIX A: ESTIMATION OF THE INVARIABLE PLANE OF EXOPLANET SYSTEMS A1 Coordinate System In Fig. 1a and this appendix we set up and explain the coordinate system used in our analysis. The invariable plane of a planetary system can be defined as the plane passing through the barycenter of the system and is perpendicular to the sum (cid:104)(cid:126)L(cid:105), of all planets in the system: (cid:104)(cid:126)L(cid:105) = ∑ j (cid:126)L j, (A1) c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 where (cid:126)L j = (Lx,Ly,Lz) is the orbital angular momentum of the jth planet. One can introduce a coordinate system in which the x axis points from the system to the observer (Fig.1a). With an x axis established, we are free to choose the direction of the z axis. For example, consider the vector (cid:126)L j in Fig.1a. If we choose a variety of z(cid:48) axes, all perpendicular to our x axis, then independent of the z(cid:48) axis, the quantity z(cid:48) is a constant. Thus, without loss of generality, we could choose a z(cid:48) axis such that Ly(cid:48) = 0. In Fig.1a, we have choosen the z axis such that the sum of the y−components of the angular momenta of all the planets, is zero: (cid:126)L j = ((cid:104)(cid:126)L(cid:105)x, 0, (cid:104)(cid:126)L(cid:105)z ). L2 y(cid:48) + L2 (cid:113) (A2) (cid:104)(cid:126)L(cid:105) = ∑ j In other words we have choosen the z axis such that the vector defining the invariable plane, (cid:104)(cid:126)L(cid:105), is in the x− z plane. We define the plane perpendicular to this vector as the invariable plane of the system. The angular separation between (cid:126)L j and (cid:126)(cid:104)L(cid:105) is φ j. φ j is a positive-valued random variable and can be well-represented by a Rayleigh distribution of mode σφ (Fabrycky & Winn 2009). For the jth planet, (cid:126)L j is the projection of (cid:126)L j onto the x-z plane. The angle between (cid:126)L j and the z axis is θ j. The angle between (cid:126)(cid:104)L(cid:105) and the z axis is (cid:104)θ(cid:105) where, (cid:104)θ(cid:105) = ∑ j θ jL j ∑L j . (A3) The angular separation in the x − z plane between (cid:126)L j and (cid:126)(cid:104)L(cid:105) is ∆θ j. In the x− z plane, we then have the relation (Fig.1a), (cid:104)θ(cid:105) + ∆θ j = θ j, ∆θ 2 (cid:113) (A4) where ∆θ j is a normally distributed variable centered around (cid:104)θ(cid:105) with a mean of 0. In other words, ∆θ j can be positive or neg- ative. A positive definite variable such as φ j is Rayleigh dis- tributed if it can be described as the sum of the squares of two independent normally distributed variables (Watson 1982), i.e. j,(y−z) where ∆θ j,(y−z) is the unobservable com- φ j = ponent of φ j in the y− z plane perpendicular to ∆θ j (see Fig.1a). From this relationship, the Gaussian distribution of ∆θ j has a stan- dard deviation equal to the mode of the Rayleigh distribution of φ j: σ∆θ = σφ . j + ∆θ 2 We can illustrate the meaning of the phrase "mutual inclina- tion" used in the literature (e.g. Fabrycky et al. (2014)). For exam- ple, in Fig. 1a, imagine adding the angular momentum vector (cid:126)Lm of another planet. And projecting this vector onto the x− z plane and call the projection (cid:126)Lm (just as we projected (cid:126)L j into (cid:126)L j). Now we can define two "mutual inclinations" between the orbital planes of these two planets. ψ3D is the angle between (cid:126)L j and (cid:126)Lm and ψ is the angle in the x− z plane between (cid:126)L j and (cid:126)Lm (i.e. ∆θ j − ∆θm). Since both ∆θ j and ∆θm are Gaussian distributed with mean µ = 0, their difference ∆θ j − ∆θm is Gaussian distributed with √ 2σ∆θ and µ(∆θ j−∆θm) = 0. Hence σ(∆θ j−∆θm) = ψ = ∆θ j − ∆θm is a positive-definite half-normal Gaussian with mean µψ =(cid:112)2/π σ(∆θ j−∆θm) = 2√ (cid:113) For ψ3D, the angle between (cid:126)L j and (cid:126)Lm, we have, (∆θ j − ∆θm)2 + (∆θ j,(y−z) − ∆θm,(y−z))2. (∆θ j − ∆θm) √ is Gaussian distributed with From above, σ(∆θ j−∆θm) = 2σ∆θ . Since we expect σ(∆θ j,(y−z)−∆θm,(y−z)) = σ(∆θ j−∆θm), ψ3D is Rayleigh distributed with mode σi (reported √ 2 σ∆θ . The mean of the in Table A1). That is, σψ3D = σi = π σ∆θ . ψ3D = (cid:113) σ 2 ∆θ j + σ 2 (A5) ∆θm = 14 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen Table A1. Comparison of exoplanet coplanarity studies Reference i Distribution Observables Modea of Rayleigh Distributed Mutual Inclinations Sample (quarter, multiplicity) N b p Np Np Kepler (Q2, 1-6) Weissbein & Steinberg (2012) Lissauer et al. (2011) Tremaine & Dong (2012) Figueira et al. (2012) Fang & Margot (2012) Johansen et al. (2012) σi ∼ 2.0◦ i < 4.0◦ σ c i ∼ 1.4◦ σ d i ∼ 1.4◦ σ c σi < 3.5◦ no fit σi ∼ 1.8◦ σi = 2.0◦ +4.0−2.0 Kepler M-dwarfs (Q16, 2-5) √ 2 σ∆θ . Assuming σ∆θ = 1.5◦ is equivalent to assuming σi = 2.1◦. a The mode σi is equal to the σψ3D discussed at the end of Appendix A1. Thus σi = σψ3D = b Np is the multiplicity vector for the numbers of observed n-planet systems, i.e. Np = (# of 1-planet systems, # of 2-planet systems, # of 3-planet systems,...). c Converted from the mean µ of the mutual inclination Rayleigh distribution: σi =(cid:112)2/π µ. Kepler (Q6, 1-6) Kepler (Q6, 1-3) Kepler (Q6, 1-6) Kepler (Q6, 1-6) Rayleigh, R of R uniform i + rotation f Np, ξ e Np Np Np, ξ Np Fabrycky et al. (2014) Ballard & Johnson (2014) RV & Kepler (Q2, 1-6) HARPS & Kepler (Q2, 1-3) Rayleigh Fisher Rayleigh Rayleigh Rayleigh Rayleigh d Converted from Rayleigh distribution relative to the invariable plane: σi = e ξ is the normalized transit duration ratio as given in Eq. 11 of Fang & Margot (2012). f Each planet is given a random uniform inclination between 0◦ − 5◦. This orbital plane is then rotated uniformally between 0− 2π to give a random longitude of ascending node. √ 2 σ∆θ . Rayleigh distribution of ψ3D is µψ3D = will have µψ3D /µψ = π 2 . √ π σ∆θ . On average we A2 Exoplanet invariable planes: permuting planet inclinations An N-planet system has N different values of θ j (see Fig. 1). Since observations are only sensitive to θ j, we don't know whether we are dealing with positive or negative angles. To model this uncer- tainty, we analyse the 2N−1 unique sets of permutations for pos- itive and negative θ j values. For example, in a 4-planet system consider the planet with the largest angular momentum. We set our coordinate system by assuming its inclination i is less than 90◦. We do not know whether the i values of the other 3 planets are on the same side or the opposite side of 90◦. The permutations of the +1s and −1s in Eq. A6 represent this uncertainty. There will be kmax = 23 = 8 sets of permutations for the θ j of the remaining 3 planets, defined by θ jMj,k where Mj,k is the permutation matrix,  1 1 1 1 −1 1 1 −1 1 1 −1 −1 −1 1 1 1 −1 −1 −1 −1 1 −1 −1 −1  . Mj,k = (A6) For each permutation k, we compute a notional invariant plane, by taking the angular-momentum-weighted average of the permuted angles (compare Eq. A3): (cid:104)θ(cid:105)k = ∑L jθ jMj,k ∑L j (A7) which yields 8 unique values of (cid:104)θ(cid:105)k, each consistent with the b j, θ j and i j values from the transit light curves (see Eqs.1,2,3). For each of these (cid:104)θ(cid:105)k, we compute a proxy for coplanarity which is the mean of the angular-momentum-weighted angle of the orbital planes around the notional invariant plane: σ(cid:104)∆θ(cid:105)k = ∑L j(cid:104)θ(cid:105)k − θ jMj,k ∑L j . (A8) The smaller the value of σ(cid:104)∆θ(cid:105)k , the more coplanar that permuta- tion set is. This permutation procedure is most appropriate when the system is close to edge-on since in this case the various plan- ets are equally likely to have actual inclinations on either side of 90◦. By contrast, when (cid:104)θ(cid:105) is large, these permutations exagger- ate the uncertainty since most planets are likely to be on the same side of 90◦ as the dominant planet. Thus, using this method, close- to-edge-on systems with (cid:104)θ(cid:105) (cid:46) 0.5◦ will yield the smallest and more appropriate dispersions which we find to be in the range: (cid:46) 1.5◦. Since the coplanarity of a system should not 0◦ (cid:46) σ(cid:104)∆θ(cid:105)k depend on the angle to the observer, the values of σ(cid:104)∆θ(cid:105)k should not depend on (cid:104)θ(cid:105). We find that this condition can best be met when we reject permutations which yield values of σ(cid:104)∆θ(cid:105)k less than 0.4◦ and greater than 1.5◦. (see Fig. A1). When no permutations for a given system meet this criteria, we select the single permutation which is closest to this range. Since the sign of (cid:104)θ(cid:105) is not impor- tant, when more than one permutation meets this criterion, we es- timate (cid:104)θ(cid:105) by taking the median of the absolute values of the (cid:104)θ(cid:105)k (cid:46) 1.5◦. These permutations are used in for which 0.4◦ (cid:46) σ(cid:104)∆θ(cid:105)k Figs. 11,12 & 13, where the most probable inclination ambiguities are indicated by two blue planets at the same semi-major axis, one above and one below the i = 90 dashed horizontal line. APPENDIX B: CALCULATING THE GEOMETRIC PROBABILITY TO TRANSIT: PTRANS We assume σ∆θ = 1.5◦ is constant over all systems (Section 1.1), such that the geometric probability for the jth planet to transit is the fraction of a Gaussian-weighted opening angle within the transit region, where the standard deviation of the Gaussian is σ j = a jσ∆θ (Figure B1 and Eq. B1). Ptrans((cid:104)θ(cid:105),a j) = 1 √ 2π σ j (cid:90) u2 u1 − u2 2σ 2 e j du (B1) where the integration limits u1 and u2 are defined by (see Fig. B1), c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 Inclinations and Titius-Bode Predictions of Kepler Systems 15 with the axes in Fig. B1 being the same as in panel b of Fig. 1, we have zo < 0. And since (cid:104)θ(cid:105) ∼ 0, we have cos(cid:104)θ(cid:105) ∼ 1. In the panels of Figs. 11, 12 and 13, the green area is a func- tion of the (cid:104)θ(cid:105) of the system. We can integrate Ptrans to get the size of the green area that is closer to the host star than acrit. This yields an area that is an estimate of the amount of parameter space in which planets can transit: Area((cid:104)θ(cid:105)) = (cid:90) acrit ((cid:104)θ(cid:105)) Ptrans da. (B5) 0 In the top panel of Fig. A1 the blue curve is a normalized version of Eq. B5 and represents the statistical expectation of the number of systems as a function of (cid:104)θ(cid:105) (ignoring detection biases). The histogram shows the values we have obtained. Figure A1. Top panel: The red histogram represents the final (cid:104)θ(cid:105) angles for the systems in our sample. The blue line represents the normalised area out to acrit as a function of (cid:104)θ(cid:105) (Eq. B5). The bottom panel shows the dispersion for our Kepler multiples, about the calculated (cid:104)θ(cid:105) value for each system. The smaller values of σ(cid:104)∆θ(cid:105) at (cid:104)θ(cid:105) ∼ 0◦ are probably due to the over-representation of i = 90◦ in the Kepler data. Figure B1. Illustration of the variables required to calculate Ptrans (Eq. B1), the fraction of a Gaussian within the transit region (b (cid:46) 1). We assume that a planet with semi-major axis a j, in a system whose invariable plane is the thick line, will have a position ∆θ j, whose probability will be described by a normal distribution along the red line perpendicular to the invariant plane. The integration variable u of Eq. B1 is along this perpendicular red line. The mean of the Gaussian is at the point (xo,zo) where u = 0. The standard deviation σ j = a jσ∆θ is also the half-width of the green region at a distance a j from the host star. u1 = R∗ + zo cos(cid:104)θ(cid:105) R∗ − zo cos(cid:104)θ(cid:105) u2 = z0 = −a j sin(cid:104)θ(cid:105). (B2) (B3) (B4) c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 01020304050 N 012345 012 σ〈∆θ〉 〈θ〉 (°) (x0,z0)b=1b=1b=0u1u2σ∆θR*zxajσj〈θ〉u=0u=1 16 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen APPENDIX C: TABLES OF PLANET PREDICTIONS Table C1: 77 Planet predictions with a high geometric probability to transit (Ptrans (cid:62) 0.55) in 40 Kepler systems (cid:18) χ2 (cid:19) (cid:18) χ2 (cid:19) - - - - - - b Ptrans 2 4 2 0 6 1 0 2.2 3.4 0.3 0.2 5.02 0.06 0.14 0.55 KOI-505 KOI-168 23.0 5.0 115.4 1.6 3.2 1.3 10.2 0.7 7.19 10.23 40.4 0.1 2.2 65.6 -1.0 1.1 2.2 7.6 0.9 5.2 1.9 1.9 KOI-1052 KOI-2585 0.01 0.56 0.74 0.19 1.36 8.20 0.90 0.19 0.16 0.14 3.69 1.07 1.69 2.26 KOI-1198 KOI-1082 1 2 2 1 2 1 5 0 KOI-1955 2 3 1 1 1 2 2 4 0.76 0.18 0.27 2.36 5.47 1.20 Period a KOI-2859 KOI-250 KOI-707 KOI-1336 Inserted Planet # d.o.f. i d.o.f. f KOI-952 KOI-500 KOI-4032 System Number Inserted γ ∆γa 1.00 1.00 1.00 1.00 0.99 0.94 1.00 1.00 0.72 1.00 1.00 1.00 1.00 0.99 0.91 1.00 0.98 0.64 0.98 0.96 0.91 0.83 0.95 0.87 0.76 0.64 0.95 0.87 0.78 0.67 0.57 0.94 0.69 0.92 0.78 0.62 0.91 0.89 0.89 0.89 0.62 0.87 0.75 0.87 0.74 0.60 0.87 0.85 0.78 0.67 0.56 0.74 0.60 0.73 0.57 0.72 0.60 0.71 0.62 0.70 0.56 0.70 0.56 0.69 0.57 0.69 0.58 0.69 0.68 0.66 0.57 KOI-156 0.61 KOI-137 0.60 KOI-1151 0.59 KOI-1015 0.58 KOI-2029 0.56 KOI-664 0.56 a ∆γ = (γ1 − γ2)/γ2 where γ1 and γ2 are the highest and second highest γ values for that system respectively (see Bovaird & Lineweaver (2013)). b Rmax is calculated by applying the lowest SNR of the detected planets in the system to the period of the inserted planet. See Eq. 7. c A planet number followed by "E" indicates the planet is extrapolated (has a larger period than the outermost detected planet in the system). (days) 2.1± 0.4 4.3± 0.7 2.6± 0.3 4.1± 0.5 6.4± 0.7 10.1± 1.1 1.8± 0.2 2.8± 0.3 14.7± 1.4 1.5± 0.3 1.5± 0.2 2.1± 0.2 3.4± 0.2 4.5± 0.2 6.9± 0.3 8.9± 0.8 6.7± 0.7 25.1± 2.5 2.4± 0.1 4.8± 0.4 6.6± 0.6 9.2± 0.8 22.6± 1.2 33.2± 1.4 48.6± 1.7 71.3± 2.2 14.8± 0.7 20.7± 0.8 28.9± 0.9 40.4± 1.1 56.4± 1.3 10.8± 1.2 27.1± 2.8 21.9± 2.5 34.7± 3.9 55± 7 8.2± 1.1 4.2± 0.5 11.8± 2.0 11.4± 1.2 28.0± 2.9 12.1± 1.2 18.3± 1.8 15.4± 1.7 23.9± 2.5 37.1± 3.9 8.4± 0.8 13.7± 2.1 23.4± 1.5 33.0± 1.8 46.5± 2.3 13.6± 1.0 21.0± 1.3 16.6± 1.2 27.4± 1.5 11.3± 0.6 16.4± 0.7 12.7± 0.6 16.4± 0.7 13.6± 1.0 20.6± 1.2 16.8± 0.9 26.1± 1.4 11.4± 0.6 16.4± 0.7 27.9± 1.5 39.0± 1.8 14.8± 2.0 32.7± 2.6 13.2± 0.5 16.9± 0.5 17.9± 1.0 31.2± 3.1 33.0± 2.2 36.1± 3.5 23.7± 1.6 40.3± 3.0 (AU) 0.03 0.06 0.04 0.05 0.07 0.10 0.03 0.04 0.11 0.02 0.02 0.03 0.04 0.05 0.07 0.09 0.07 0.17 0.03 0.05 0.06 0.07 0.17 0.21 0.28 0.36 0.12 0.15 0.19 0.24 0.30 0.10 0.18 0.15 0.20 0.27 0.08 0.04 0.10 0.10 0.18 0.10 0.14 0.13 0.17 0.23 0.07 0.11 0.17 0.21 0.27 0.10 0.14 0.13 0.17 0.08 0.10 0.09 0.11 0.11 0.14 0.12 0.16 0.10 0.12 0.18 0.22 0.11 0.22 0.11 0.13 0.12 0.19 0.20 0.22 0.15 0.23 1 2 1 2 3 4 1 2 3 E c 1 1 2 1 2 3 E 1 1 2 1 1 2 3 1 E 2 E 3 E 4 E 1 E 2 E 3 E 4 E 5 E 1 2 1 2 3 1 1 1 1 2 1 2 1 2 3 1 1 1 E 2 E 3 E 1 E 2 E 1 E 2 E 1 E 2 E 1 E 2 E 1 E 2 E 1 2 1 E 2 E 1 E 2 E 1 1 E 1 E 2 E 1 E 1 E 1 E 1 E 1 E 1 E Rmax (R⊕) 1.3 1.5 0.9 1.0 1.2 1.3 1.0 1.1 1.6 0.9 1.2 1.3 0.8 0.9 1.0 2.0 1.7 2.4 0.6 1.2 1.3 1.4 2.6 2.9 3.2 3.5 1.1 1.2 1.3 1.4 1.6 1.6 2.0 4.5 5.1 5.7 0.9 1.4 2.1 2.0 2.5 1.5 1.6 1.4 1.6 1.8 0.6 1.5 1.4 1.5 1.6 1.6 1.8 1.9 2.1 2.0 2.2 0.4 0.4 1.3 1.5 0.3 0.3 1.5 1.7 2.3 2.5 0.8 2.1 0.7 0.7 1.6 3.1 1.0 2.3 0.9 1.5 Teff (K) 1642 1297 1568 1347 1157 994 1184 1029 588 904 1091 960 1347 1224 1061 1162 1053 679 1242 686 616 553 909 800 704 620 967 865 774 692 619 909 669 788 676 580 739 633 761 668 494 828 720 813 702 607 621 756 774 690 615 522 451 586 497 514 455 566 520 703 612 582 502 711 630 620 554 514 703 751 692 476 539 564 590 514 618 KOI-1831 KOI-248 KOI-880 KOI-1567 KOI-1952 KOI-730 KOI-719 KOI-1060 KOI-3083 KOI-701 KOI-1306 KOI-2722 0.10 0.13 0.85 0.62 0.31 0.06 1.11 0.13 0.28 0.07 0.01 2.54 2.48 1.35 1.51 3.26 1.3 18.3 4.0 5.9 85.6 0.2 80.8 22.4 44.8 16.6 0.33 1.36 0.22 0.56 0 0 1 0 0 0 0 0 0 0 0 - 1.5 - - - 1.1 - - 4.04 4.12 0.54 18.0 5.7 - 0 0 0 0 0 3 KOI-1358 KOI-1627 KOI-1833 KOI-2055 - - - - - - - - - - - - - - - - KOI-3158 0.01 0.24 0.74 0.23 0.29 1.56 0.49 1.0 7.4 0.4 - - - - - - - - - - - - - - - - - - - - - - - - - KOI-245 KOI-749 0.65 0.01 0.62 1.0 1.3 0.17 0.66 KOI-351 0.5 5.78 c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 Table C2: All 228 Planet Predictions in 151 Systems (Table C1 is a high Ptrans subset of this table) Inclinations and Titius-Bode Predictions of Kepler Systems 17 (cid:18) χ2 (cid:19) (cid:18) χ2 (cid:19) d.o.f. i d.o.f. f 7.19 10.23 5.02 2.36 5.47 1.20 3.69 1.07 1.69 2.26 0.14 0.55 1.36 8.20 2.54 2.48 1.35 1.51 3.26 5.78 0.74 0.19 0.06 0.76 0.18 0.27 0.90 0.19 0.16 0.14 - - 0.01 0.56 1.11 0.13 0.28 0.07 0.01 0.65 γ 2.2 3.4 23.0 2.2 7.6 0.9 3.2 1.3 10.2 0.7 - - ∆γa 0.3 0.2 5.0 5.2 1.9 1.9 2.2 65.6 -1.0 1.1 - - 115.4 40.4 1.6 0.1 1.3 18.3 4.0 5.9 85.6 0.5 0.2 80.8 22.4 44.8 16.6 1.0 System KOI-1198 KOI-1955 KOI-1082 KOI-952 KOI-500 KOI-4032 KOI-707 KOI-1336 KOI-2859 KOI-250 KOI-168 KOI-2585 KOI-1052 KOI-505 KOI-1831 KOI-248 KOI-880 KOI-1567 KOI-1952 KOI-351 KOI-701 KOI-1306 KOI-2722 KOI-1358 KOI-1627 KOI-1833 KOI-3158 Number Inserted 2 4 2 1 2 2 1 2 1 5 0 0 2 3 1 1 1 2 2 4 6 1 0 0 0 0 0 Inserted Planet # 1 2 3 E c 1 2 3 4 5 E 1 2 3 E 1 2 E 1 2 3 E 1 2 3 E 1 2 E 1 2 3 E 1 2 E 1 2 3 4 5 6 E 1 E 2 E 3 E 4 E 5 E 1 E 2 E 3 E 4 E 5 E 1 2 3 E 1 2 3 4 E 1 2 E 1 2 E 1 2 E 1 2 3 E 1 2 3 E 1 2 3 4 5 E 1 2 3 4 5 6 7 E 1 2 E 1 E 2 E 3 E 1 E 2 E 1 E 2 E 1 E 2 E 1 E 2 E Period (days) 2.1± 0.4 4.3± 0.7 73± 12 2.6± 0.3 4.1± 0.5 6.4± 0.7 10.1± 1.1 62± 7 1.8± 0.2 2.8± 0.3 14.7± 1.4 1.5± 0.3 40.0± 6.2 1.5± 0.2 2.1± 0.2 14.5± 1.3 3.4± 0.2 4.5± 0.2 6.9± 0.3 8.9± 0.8 68± 7 6.7± 0.7 25.1± 2.5 60± 6 2.4± 0.1 5.1± 0.3 4.8± 0.4 6.6± 0.6 9.2± 0.8 24.1± 1.9 33.2± 2.6 63.3± 5.0 22.6± 1.2 33.2± 1.4 48.6± 1.7 71.3± 2.2 104.5± 2.7 14.8± 0.7 20.7± 0.8 28.9± 0.9 40.4± 1.1 56.4± 1.3 10.8± 1.2 27.1± 2.8 110± 20 21.9± 2.5 34.7± 3.9 55± 7 140± 20 8.2± 1.1 100± 20 4.2± 0.5 30.1± 3.3 11.8± 2.0 120± 20 11.4± 1.2 28.0± 2.9 69± 8 12.1± 1.2 18.3± 1.8 64± 7 15.4± 1.7 23.9± 2.5 37.1± 3.9 140± 20 520± 60 8.4± 0.8 26.6± 2.6 39.1± 3.7 57± 6 84± 8 180± 20 390± 40 13.7± 2.1 55± 9 23.4± 1.5 33.0± 1.8 46.5± 2.3 13.6± 1.0 21.0± 1.3 16.6± 1.2 27.4± 1.5 11.3± 0.6 16.4± 0.7 12.7± 0.6 16.4± 0.7 a (AU) 0.03 0.06 0.37 0.04 0.05 0.07 0.10 0.33 0.03 0.04 0.11 0.02 0.19 0.02 0.03 0.10 0.04 0.05 0.07 0.09 0.35 0.07 0.17 0.31 0.03 0.05 0.05 0.06 0.07 0.14 0.17 0.27 0.17 0.21 0.28 0.36 0.46 0.12 0.15 0.19 0.24 0.30 0.10 0.18 0.46 0.15 0.20 0.27 0.51 0.08 0.41 0.04 0.16 0.10 0.45 0.10 0.18 0.32 0.10 0.14 0.31 0.13 0.17 0.23 0.54 1.31 0.07 0.15 0.19 0.25 0.32 0.54 0.90 0.11 0.28 0.17 0.21 0.27 0.10 0.14 0.13 0.17 0.08 0.10 0.09 0.11 b Rmax (R⊕) 1.3 1.5 3.1 0.9 1.0 1.2 1.3 2.0 1.0 1.1 1.6 0.9 2.1 1.2 1.3 2.2 0.8 0.9 1.0 2.0 3.3 1.7 2.4 3.0 0.6 0.8 1.2 1.3 1.4 1.8 2.0 2.3 2.6 2.9 3.2 3.5 3.8 1.1 1.2 1.3 1.4 1.6 1.6 2.0 2.8 4.5 5.1 5.7 7.2 0.9 1.7 1.4 2.2 2.1 3.7 2.0 2.5 3.1 1.5 1.6 2.2 1.4 1.6 1.8 2.4 3.4 0.6 0.8 0.9 1.0 1.1 1.3 1.6 1.5 2.1 1.4 1.5 1.6 1.6 1.8 1.9 2.1 2.0 2.2 0.4 0.4 e Teff (K) 1642 1297 505 1568 1347 1157 994 541 1184 1029 588 904 299 1091 960 506 1347 1224 1061 1162 590 1053 679 507 1242 967 686 616 553 401 360 290 909 800 704 620 546 967 865 774 692 619 909 669 423 788 676 580 426 739 316 633 329 761 354 668 494 366 828 720 474 813 702 607 391 252 621 423 372 327 288 223 172 756 476 774 690 615 522 451 586 497 514 455 566 520 Ptrans d 1.00 1.00 0.41 1.00 1.00 0.99 0.94 0.42 1.00 1.00 0.72 1.00 0.36 1.00 1.00 0.50 1.00 0.99 0.91 1.00 0.50 0.98 0.64 0.39 0.98 0.54 0.96 0.91 0.83 0.53 0.44 0.29 0.95 0.87 0.76 0.64 0.52 0.95 0.87 0.78 0.67 0.57 0.94 0.69 0.31 0.92 0.78 0.62 0.36 0.91 0.25 0.89 0.30 0.89 0.27 0.89 0.62 0.37 0.87 0.75 0.38 0.87 0.74 0.60 0.27 0.12 0.87 0.52 0.41 0.33 0.25 0.15 0.09 0.85 0.43 0.78 0.67 0.56 0.74 0.60 0.73 0.57 0.72 0.60 0.71 0.62 18.0 7.4 4.04 0.01 5.7 0.4 - - - - - - - - - - 4.12 0.54 0.01 0.24 0.74 0.23 0.62 - - - - - continued c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 T. Bovaird, C. H. Lineweaver and S. K. Jacobsen 18 Table C2: All 228 Planet Predictions in 151 Systems (Table C1 is a high Ptrans subset of this table) (cid:18) χ2 (cid:19) Inserted Planet # 3 E 1 E 2 E 1 2 3 4 E 1 E 2 E 1 E 2 E 1 2 E 1 E 2 E 1 E 2 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 2 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 2 E 1 E 1 E 1 E 1 2 3 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 2 3 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E Period (days) 21.1± 0.8 13.6± 1.0 20.6± 1.2 16.8± 0.9 26.1± 1.4 32.6± 1.7 63.1± 3.3 11.4± 0.6 16.4± 0.7 27.9± 1.5 39.0± 1.8 14.8± 2.0 88± 12 32.7± 2.6 52.7± 3.5 13.2± 0.5 16.9± 0.5 17.9± 1.0 31.2± 3.1 33.0± 2.2 36.1± 3.5 23.7± 1.6 40.3± 3.0 19.1± 1.4 28.7± 3.3 56± 6 72± 7 39.1± 5.4 130± 20 34.8± 3.6 49.5± 5.6 24.6± 3.0 29.0± 2.5 40.5± 2.4 35.5± 3.5 7.6± 0.4 46.3± 3.1 45.9± 5.3 20.3± 1.2 12.7± 1.1 56.2± 3.8 17.7± 1.0 19.1± 1.6 21.4± 1.9 30.3± 2.3 38.3± 2.9 75± 8 170± 20 44.0± 4.6 20.3± 2.1 39.1± 3.6 63± 11 130± 30 580± 100 59± 7 55± 6 26.1± 2.9 73± 7 31.3± 3.6 33.1± 3.5 82± 9 53± 7 76± 8 130± 20 37.0± 2.3 18.2± 1.9 40.9± 5.6 73± 10 230± 40 72± 8 86± 10 100± 20 11.2± 1.0 15.2± 0.8 33.2± 2.2 87± 13 107± 7 84± 12 110± 20 34.9± 3.3 110± 20 20.7± 1.8 64± 6 73± 8 74± 9 75± 12 a (AU) 0.13 0.11 0.14 0.12 0.16 0.18 0.28 0.10 0.12 0.18 0.22 0.11 0.35 0.22 0.30 0.11 0.13 0.12 0.19 0.20 0.22 0.15 0.23 0.12 0.16 0.30 0.35 0.22 0.49 0.20 0.27 0.14 0.16 0.24 0.20 0.07 0.26 0.25 0.16 0.09 0.28 0.14 0.14 0.14 0.18 0.21 0.34 0.60 0.24 0.15 0.20 0.31 0.51 1.35 0.29 0.29 0.14 0.35 0.16 0.16 0.37 0.25 0.36 0.55 0.21 0.12 0.19 0.28 0.61 0.32 0.38 0.44 0.10 0.12 0.16 0.38 0.46 0.35 0.42 0.20 0.44 0.15 0.27 0.27 0.34 0.32 b Rmax (R⊕) 0.5 1.3 1.5 0.3 0.3 0.3 0.4 1.5 1.7 2.3 2.5 0.8 1.2 2.1 2.3 0.7 0.7 1.6 3.1 1.0 2.3 0.9 1.5 1.0 1.8 1.3 2.3 1.2 1.7 2.9 1.7 1.7 1.8 2.3 2.8 0.7 1.7 2.1 1.1 1.3 1.8 1.8 1.2 2.1 1.8 0.9 2.7 3.4 2.9 1.9 2.8 2.7 3.3 4.7 2.6 1.9 3.9 2.0 1.7 1.9 2.4 2.9 3.5 4.2 1.8 2.0 0.8 0.9 1.2 1.7 1.3 1.3 1.5 1.5 1.6 2.0 2.6 3.8 2.8 3.0 3.6 1.4 2.6 2.8 2.0 4.7 e Teff (K) 478 703 612 582 502 467 374 711 630 620 554 514 284 703 599 751 692 476 539 564 590 514 618 430 452 586 541 498 331 477 512 377 374 647 437 868 532 517 845 415 458 791 695 574 466 408 439 334 453 640 360 409 320 196 461 500 379 455 323 293 440 362 442 452 494 499 294 242 164 374 425 426 893 661 298 397 424 349 376 508 381 649 289 260 460 343 Ptrans d 0.55 0.70 0.56 0.70 0.56 0.50 0.33 0.69 0.57 0.69 0.58 0.69 0.24 0.68 0.53 0.66 0.57 0.61 0.60 0.59 0.58 0.56 0.56 0.53 0.53 0.53 0.52 0.52 0.24 0.51 0.49 0.48 0.47 0.47 0.46 0.46 0.44 0.44 0.44 0.44 0.44 0.43 0.42 0.42 0.41 0.41 0.41 0.24 0.40 0.40 0.40 0.39 0.25 0.09 0.39 0.38 0.38 0.38 0.37 0.37 0.37 0.37 0.37 0.36 0.36 0.35 0.35 0.24 0.11 0.34 0.34 0.34 0.34 0.33 0.32 0.32 0.32 0.31 0.30 0.30 0.30 0.30 0.30 0.29 0.29 0.29 c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19 System KOI-2055 KOI-245 KOI-749 KOI-730 KOI-719 KOI-1060 KOI-3083 KOI-156 KOI-137 KOI-1151 KOI-1015 KOI-2029 KOI-664 KOI-2693 KOI-1590 KOI-279 KOI-1930 KOI-70 KOI-720 KOI-1860 KOI-1475 KOI-1194 KOI-2025 KOI-733 KOI-2169 KOI-2163 KOI-3319 KOI-2352 KOI-1681 KOI-1413 KOI-2597 KOI-2220 KOI-1161 KOI-582 KOI-82 KOI-157 KOI-864 KOI-939 KOI-898 KOI-841 KOI-408 KOI-1909 KOI-2715 KOI-1278 KOI-1867 KOI-899 KOI-1589 KOI-884 KOI-829 KOI-94 KOI-2038 KOI-1557 KOI-571 KOI-1905 KOI-116 KOI-2732 KOI-665 KOI-1931 KOI-886 KOI-1432 KOI-945 KOI-869 KOI-111 KOI-1364 KOI-1832 KOI-658 KOI-1895 KOI-2926 KOI-1647 KOI-941 Number Inserted 0 3 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 2 0 0 0 0 0 0 0 0 0 0 0 0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 γ - 1.0 - - ∆γa - 1.3 - - 1.1 1.5 - - - - - - - - - - - - 7.4 - - - - - - - - - - - - - - - - - 3.7 - - - 7.1 - - - - - - - - - - - - 19.9 - - - - - - - - - - - - - - - - - continued - - - - - - - - - - - - 2.9 - - - - - - - - - - - - - - - - - 6.9 - - - 0.4 - - - - - - - - - - - - 5.2 - - - - - - - - - - - - - - - - - (cid:18) χ2 (cid:19) d.o.f. i d.o.f. f 0.29 1.56 0.49 0.33 1.36 0.22 0.56 0.10 0.13 0.85 0.62 0.31 0.06 0.01 0.64 0.13 0.60 3.51 0.14 0.02 0.94 0.52 0.21 0.22 0.87 0.06 0.01 0.28 0.15 0.14 0.13 0.19 0.18 0.08 0.92 3.15 0.09 0.24 0.08 4.35 0.51 0.26 0.62 0.52 0.53 0.01 0.56 0.45 0.07 0.19 0.11 0.58 4.87 0.01 0.34 0.22 0.01 0.20 0.46 0.07 0.04 0.08 0.03 0.41 0.03 0.60 0.11 0.49 0.07 0.36 - 0.17 - - 0.66 - - - - - - - - - - - - 0.42 - - - - - - - - - - - - - - - - - 0.69 - - - 0.15 - - - - - - - - - - - - 0.07 - - - - - - - - - - - - - - - - - Inclinations and Titius-Bode Predictions of Kepler Systems 19 Table C2: All 228 Planet Predictions in 151 Systems (Table C1 is a high Ptrans subset of this table) (cid:18) χ2 (cid:19) (cid:18) χ2 (cid:19) Inserted Period a b Planet # 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 2 3 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E 1 E (days) 35.3± 3.1 120± 20 26.9± 2.5 140± 20 150± 20 15.2± 0.6 32.4± 2.7 15.6± 0.6 150± 30 110± 20 110± 20 110± 20 160± 20 170± 30 16.4± 1.2 40.0± 3.4 78± 9 110± 20 78± 11 23.3± 2.0 180± 30 94± 9 26.4± 2.0 120± 20 120± 20 100± 20 74± 10 31.3± 3.6 160± 30 320± 60 1380± 250 130± 20 110± 20 230± 40 250± 50 110± 20 12.5± 0.7 42.2± 3.5 280± 50 230± 20 17.9± 1.6 260± 50 290± 30 14.2± 1.1 180± 40 200± 30 310± 40 290± 40 160± 40 640± 120 430± 70 530± 110 790± 220 770± 190 940± 190 9.5± 0.8 (AU) 0.21 0.48 0.17 0.53 0.57 0.13 0.20 0.13 0.59 0.46 0.42 0.45 0.60 0.63 0.12 0.20 0.30 0.41 0.34 0.17 0.68 0.43 0.17 0.47 0.46 0.37 0.33 0.19 0.55 0.90 2.35 0.46 0.38 0.77 0.76 0.37 0.12 0.23 0.83 0.74 0.13 0.82 0.88 0.12 0.61 0.57 0.89 0.79 0.56 1.55 1.11 1.28 1.69 1.64 1.59 0.07 Rmax (R⊕) 2.2 2.1 3.6 1.9 2.9 2.7 2.8 1.3 1.4 2.1 1.8 2.3 3.6 3.8 1.8 1.6 2.4 4.2 3.5 1.4 4.1 2.2 1.5 2.5 2.5 2.2 4.2 2.3 2.3 2.8 4.0 2.8 2.5 3.8 4.5 1.9 1.7 1.3 1.5 7.6 3.5 1.8 5.3 2.0 3.4 3.2 2.5 3.6 4.0 3.4 5.1 4.3 2.9 3.0 3.1 3.2 e Teff (K) 709 388 491 414 395 933 630 1033 391 391 314 335 381 337 575 343 253 295 413 692 374 490 614 351 334 252 357 555 321 252 155 274 224 336 309 198 998 592 299 298 779 303 294 681 301 215 247 235 282 272 218 222 199 188 124 559 Ptrans d 0.28 0.28 0.28 0.28 0.28 0.27 0.27 0.27 0.27 0.26 0.26 0.25 0.25 0.25 0.25 0.24 0.24 0.24 0.24 0.23 0.23 0.22 0.22 0.22 0.22 0.22 0.22 0.21 0.21 0.13 0.05 0.20 0.20 0.20 0.19 0.19 0.19 0.19 0.19 0.18 0.18 0.17 0.16 0.16 0.16 0.15 0.15 0.15 0.14 0.13 0.10 0.10 0.08 0.07 0.05 0.00 System KOI-3741 KOI-700 KOI-1563 KOI-2135 KOI-2433 KOI-2086 KOI-1102 KOI-3097 KOI-1445 KOI-232 KOI-520 KOI-2707 KOI-152 KOI-1332 KOI-2485 KOI-877 KOI-775 KOI-757 KOI-510 KOI-117 KOI-935 KOI-285 KOI-671 KOI-834 KOI-509 KOI-904 KOI-723 KOI-1436 KOI-435 Number Inserted 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 2 γ - - - - - - - - - - - - - - - - - - - - - - - - - - - - 1.8 ∆γa - - - - - - - - - - - - - - - - - - - - - - - - - - - - 0.3 d.o.f. i d.o.f. f 0.01 0.91 0.62 0.12 0.55 0.33 0.75 0.31 0.01 0.93 0.18 0.29 0.64 0.03 0.17 0.32 0.04 0.01 0.05 0.42 0.03 0.04 0.62 0.84 0.22 0.93 0.04 0.19 6.69 - - - - - - - - - - - - - - - - - - - - - - - - - - - - 0.84 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 KOI-2073 KOI-812 KOI-474 KOI-907 KOI-1422 KOI-710 KOI-623 KOI-282 KOI-620 KOI-3925 KOI-2167 KOI-1426 KOI-1127 KOI-191 KOI-1430 KOI-806 KOI-612 KOI-481 KOI-2714 KOI-1258 KOI-564 KOI-1922 KOI-2183 KOI-518 KOI-2842 a ∆γ = (γ1 − γ2)/γ2 where γ1 and γ2 are the highest and second highest γ values for that system respectively (see BL13). b Rmax is calculated by applying the lowest SNR of the detected planets in the system to the period of the inserted planet. See Eq. 7. c A planet number followed by "E" indicates the planet is extrapolated (has a larger period than the outermost detected planet in the system). d Ptrans values (cid:62) 0.55 are shown in bold, indicating a higher probability to transit. e Teff values between Mars and Venus (206 K to 300 K, assuming an albedo of 0.3) are shown in bold. 0.08 0.04 0.54 0.91 0.04 0.69 0.40 0.01 0.84 0.38 0.05 0.02 0.81 0.99 0.98 0.16 0.10 0.02 0.09 0.96 0.77 0.02 0.70 0.84 0.27 - - - - - - - - - - - - - - - - - - - - - - - - - c(cid:13) 2015 RAS, MNRAS 000, 1 -- 19
1306.1005
1
1306
2013-06-05T07:22:58
Microlensing by a wide-separation planet: detectability and boundness
[ "astro-ph.EP" ]
We explore the detection condition of a wide-separation planet through the perturbation induced by the planetary caustic for various microlensing parameters, especially for the size of the source stars. By constructing the fractional deviation maps at various positions in the space of microlensing parameters, we find that the pattern of the fractional deviation depends on the ratio of the source radius to the caustic size, and the ratio satisfying the observational threshold varies with the star-planet separation. We have also obtained the upper limits of the source size that allows the detection of the signature of the host star as a function of the separation for given observational threshold. It is shown that this relation further leads one to a simple analytic condition for the star-planet separation to detect the boundness of wide-separation planets as a function of the mass ratio and the source radius. For example, when 5% of the detection threshold is assumed, for a source star with the radius of ~1 R_sun, an Earth-mass planet and a Jupiter-mass planet can be recognized of its boundness when it is within the separation range of ~10 AU and ~30 AU, respectively. We also compare the separation ranges of detection by the planetary caustic with those by the central caustic. It is found that when the microlensing light curve caused by the planetary caustic happens to be analyzed, one may afford to support the boundness of the wide-separation planet farther than when that caused by the central caustic is analyzed. Finally, we conclude by briefly discussing the implication of our findings on the next-generation microlensing experiments.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 22 July 2018 (MN LATEX style file v2.2) Microlensing by a wide-separation planet: detectability and boundness Yoon-Hyun Ryu1⋆, Myeong-Gu Park1⋆, Heon-Young Chang1⋆, and Ki-Won Lee2⋆ 1Department of Astronomy and Atmospheric Sciences, Kyungpook National University, Daegu 702-701, Rep. of Korea 2Institute of Liberal Education, Catholic University of Daegu, Gyeongbuk 712-702, Rep. of Korea 22 July 2018 ABSTRACT The microlensing technique has proven sensitive to Earth-like and/or wide- separation extrasolar planets. The unbiased spatial distribution of extrasolar planets with respect to host stars is crucial in studying the planet formation processes. If one can characterize the planetary microlensing light curves whether they are produced by a wide-separation planet or a free-floating planet, it will greatly help to establish the spatial distribution of extrasolar planets without contamination by free-floating plan- ets. Previous studies have shown that the effect of the host star on the microlensing by the accompanying wide-separation planet can be significant enough to be detected by the high-frequency microlensing experiments for typical microlensing parameters. Here, we further explore the detection condition of a wide-separation planet through the perturbation induced by the planetary caustic for various microlensing parame- ters, especially for the size of the source stars. By constructing the fractional deviation maps at various positions in the space of microlensing parameters, we find that the pattern of the fractional deviation depends on the ratio of the source radius to the caustic size, and the ratio satisfying the observational threshold varies with the star- planet separation. We have also obtained the upper limits of the source size that allows the detection of the signature of the host star as a function of the separation for given observational threshold. It is shown that this relation further leads one to a simple analytic condition for the star-planet separation to detect the boundness of wide-separation planets as a function of the mass ratio and the source radius. For example, when 5% of the detection threshold is assumed, for a source star with the ra- dius of ∼ 1 R⊙, an Earth-mass planet and a Jupiter-mass planet can be recognized of its boundness when it is within the separation range of ∼ 10 AU and ∼ 30 AU, respec- tively. We also compare the separation ranges of detection by the planetary caustic with those by the central caustic. It is found that when the microlensing light curve caused by the planetary caustic happens to be analyzed, one may afford to support the boundness of the wide-separation planet farther than when that caused by the central caustic is analyzed. Finally, we conclude by briefly discussing the implication of our findings on the next-generation microlensing experiments. Key words: gravitational lensing -- planetary systems -- planets and satellites: general 3 1 0 2 n u J 5 . ] P E h p - o r t s a [ 1 v 5 0 0 1 . 6 0 3 1 : v i X r a 1 INTRODUCTION Up to the present, 19 extrasolar planets have been dis- covered towards the Galactic bulge by the microlensing experiments operating in the survey and follow-up mode (Bond et al. 2004; Udalski et al. 2005; Beaulieu et al. 2006; Gould et al. 2006; Gaudi et al. 2008; Bennett et al. 2008; ⋆ E-mail: (MGP); [email protected] (HYC); [email protected] (KWL) [email protected](YHR); [email protected] Dong et al. 2009; Sumi et al. 2010; Janczak et al. 2010; Miyake et al. 2011; Batista et al. 2011; Muraki et al. 2011; Yee et al. 2012; Bennett et al. 2012; Bachelet et al. 2012; Han et al. 2013; Kains et al. 2013). A planetary microlens- ing event shows a perturbation, lasting for a short-duration of tE,p ∼ √qtE, typically about a day and a few hours for the mass ratio q corresponding to a Jupiter-mass planet and an Earth-mass planet, respectively, to the standard light curve induced by a lens star, whose typical timescale is the Einstein timescale for the lens system tE (Mao & Paczynski 2 Ryu et al. 1991; Gould & Loeb 1992). Continuous monitoring with a high cadence is, therefore, prerequisite to avoid missing plan- etary microlensing events. To monitor a wide field of view continuously with high cadence, existing microlensing experiments, such as, the OGLE collaboration and MOA collaboration, have recently changed the camera to expand the field of view or planned to upgrade the telescope. Next-generation surveys, with a wider field of view and higher cadence such as Korea Mi- crolensing Telescope Network (KMTNet) and Wide-Field Infrared Survey Telescope (WFIRST ) are being prepared to search the extrasolar planets by the microlensing tech- nique in ground and space, respectively. The main aim of KMTNet project is to find sub-Earth-mass planets using three 1.6 m wide field (2 × 2 degrees field of view) opti- cal telescopes located in Chile, South Africa, and Australia (Kim et al. 2010). The key goal of WFIRST mission is to discover planets down to 0.1 Earth-mass in the separation range wider than 0.5 AU, consequently to dispense a com- plete statistical census of the planet in our Galaxy (Bennett 2011). Such surveys with a frequent sampling can be suffi- ciently sensitive to wide-separation and free-floating planets, and thus enable us to detect wide-separation planets as well as free-floating planets which are otherwise very difficult to find by other planet search techniques, such as, radial ve- locity technique, transit technique, direct imaging, or pulsar timing analysis (Bennett & Rhie 2002; Han & Kang 2003; Sumi et al. 2011; Bennett et al. 2012). In fact, Sumi et al. (2011) recently reported the discovery of unbound or pos- sibly distantly orbiting populations with a planetary-mass, based on microlensing survey observations on MOA-II phase between 2006 and 2007. According to their statistical es- timates the number of such populations is approximately twice that of main-sequence stars in the Galaxy. If one can tell via microlensing observations whether these planets are still bound to their host stars or freely floating without host stars, essential information can be provided on the spatial distribution of extrasolar planets with respect to host stars without contamination by unbound planets so that the core accretion theory can be tightly constrained (Laughlin, Bodenheimer, & Adams 2004; Ida, & Lin 2005; Kennedy, Kenyon, & Bromley 2006; Kennedy & Kenyon 2008). A free-floating planet itself is also important in constraining the planet formation processes since it is related to the processes, such as star- planet scattering (Holman & Wiegert 1999; Musielak et al. 2005; Malmberg et al. scattering (Rasio & Ford 1996; Lin & Ida 1997; Ford & Rasio 2008; Veras & Raymond 2012), and star death (Veras et al. 2011). 1996; Weidenschilling & Marzari 2011), planet-planet Several suggestions have been made to distin- guish wide-separation planets from free-floating planets (Di Stefano & Scalzo 1999a,b; Han & Kang 2003; Han et al. 2005; Han 2009b; Di Stefano 2012a,b). Basically, a plane- tary microlensing event can be inferred that it is caused by a wide-separation planet if the light curve shows any signature of the host star, such as, a long-term bump, blended light, or a signal of the planetary caustic in the light curve. The former two signatures depend on the source trajectory and lens flux, respectively, and need additional long-term obser- vations for confirmation, while the last signature will become important in the next-generation surveys with high survey monitoring frequency. Han & Kang (2003) investigated the microlensing properties of wide-orbit planets and found that the signature of the central star can be detected for a large fraction of Jupiter mass planetary system for a typical source size and star-planet distance. Han et al. (2005) comprehen- sively discussed these three methods for the detection of a host star, and found that one-third of all events should show signatures of its host star regardless of the planet separation through these methods. In this paper we investigate the condition in terms of lensing parameters, particularly, the size of source stars, un- der which signatures of the host star in the planetary mi- crolensing light curve can be detected. We construct mag- nification maps for various source radii by the inverse ray- shooting method, taking the limb darkening into considera- tion, since the planetary caustic located close to the wide- separation planet is so small that the finite source effect becomes critical for a given photometric accuracy. As a re- sult, we obtain a general empirical formula for the upper limit on the ratio of source size to the planetary caustic size that allows to detect the signature of the host star as a func- tion of the separation, which asymptotically approaches to a constant as the separation goes to infinity, as approximated by Chang & Refsdal (1979, 1984). We also compare the sep- aration ranges to detect the boundness of the planet to the host star through the channel of the central and planetary caustics. This paper is organized as follows. We briefly describe properties of a planetary caustic for a wide-separation plan- etary system in §2. We discuss how we construct the frac- tional deviation maps in §3. We present conditions to detect signatures of a host star in the planetary microlensing light curve in §4. Finally, we summarize and conclude our results in §5. 2 PROPERTIES OF PLANETARY CAUSTIC A planetary perturbation occurs when a source star crosses or passes by the caustic resulted from a star-planet system. Provided that an extrasolar planetary system has a single planet, the planetary system produces one, two or three dis- connected caustics depending on the separation between a planet and its host star. In a wide-separation planetary sys- tem, in which the projected separation of the planet and the host star is larger than the upper limit of the lensing zone, two disconnected caustics are generated: central and plane- tary caustics. Here, the position of the planetary caustic is so close to the planet that sometimes the planetary caustic is located within the Einstein ring due to the planet. The planetary caustic structure can be expressed by a analytic form (e.g., Witt & Mao 1995; Bozza 2000). The formulae of x− and y−components of the caustics in the polar coordinate system, x(θ) and y(θ), are given by x(θ) = (s − + (1 + m1 s m2 ε2 1 ) + (ε1 − )ε2 cos θ + m1ε1 cos θ ) cos θ + m2 s2 ε1 m1 s3 (ε2s cos θ − ε2 y(θ) = (ε1 − m2 ε1 ) sin θ − m1ε1 sin θ s2 1 cos 2θ), (1) (2) + (1 + )ε2 sin θ − where ε1 and ε2 are given by m2 ε2 1 m1 s3 (ε2s sin θ − ε2 1 sin 2θ), , !1/2 m1ε4 1 sin2 2θ m1 cos 2θ ±ps4 − m2 s2 − m1/s2 ε1 = m2 ε2 = − Here, m1 and m2 are the masses of lens star and planet, respectively, and s is the projected separation between them normalized by the Einstein ring radius. 1 cos θ + m2s2 cos 3θ) 1 cos 2θ + m2s2) m2s3(m1ε2 1(m1ε2 (3) (4) . Hence, the planetary caustic size along the planet-star axis is ∆x = x(0) − x(π) = 2√m2 s m1(2s2 + m1 + 1) p(s4 − m1)(s2 + m1) . (5) If we choose m1 = 1 and m2 = q = m2/m1 (also see Han 2006), then Equation 5 becomes 4√q s√s2 − 1 . ∆x ≃ As shown in Equation 6, the planetary caustic size ∆x is proportional to √q and inversely to s2 when s ≫ 1. Thus, a low mass planet with wide separation has a fairly small perturbation region. (6) 3 FRACTIONAL DEVIATION MAP In Figure 1, we show the fractional deviation map for vari- ous ratios of the source radius normalized by the planetary caustic size ρ⋆/∆x, where ρ⋆ = θ⋆/θE and θ⋆ being the an- gular radius of a source star, and normalized separation s by the Einstein ring radius θE. The ratio of the normalized source radius and separation are denoted on the top of each column and on the right-hand side of each row, respectively. The fractional deviation σ is defined by A − AP , AP σ ≡ where the magnification of the planetary lensing A results from a planet with its host star, and AP from a planet alone, like the case of a free-floating planet. Note that AP is scaled by the Einstein ring radius of the planet itself, θE,P, which is related to the Einstein ring radius for the total mass of (7) the star-planet system, θE, as θE,P = θEpq/(1 + q). The position of the peak magnification of the planet itself, s, is also adjusted to match the center of a planetary caustic of the star-planet system, i.e., s = s − 1/s. Since a low mass planet with wide separation has a fairly small perturbation region, the finite source effect is crucial. We construct the magnification maps using the in- verse ray-shooting method to include the finite source effect (Schneider & Weiss 1986; Kayser et al. 1986; Wambsganss 1997). In addition to the finite source effect, we also take into account the limb darkening effect by the surface of source star. On the surface of the source star, the specific intensity I with the flux F and the limb darkening coefficient of Γ is given by I = F πθ2 ⋆ [1 − Γ(1 − 1.5 cos φ)], (8) Microlensing by a wide-separation planet 3 where φ is the angle between the normal direction to the star's surface and the direction toward the observer (Milne 1921; An et al. 2002). In this particular study, we adopt a fixed value of Γ for all source stars, i.e., Γ = 0.5. We also assume that the lens system with the mass ratio of 1× 10−3 is located at Dl = 6 kpc and the source star is at Ds = 8 kpc. The panels in Figure 1 are divided by thick black lines with arrows to outline in the parameter space the domains where regions with the amplitude of fractional deviations as indicated by the arrows can be found in the deviation maps. Note that the contours are drawn at the levels of σ = ±1%, ±5%, ±10%, ±15%, and ±20% as shown in the scale-bar. As one may expect, the fractional deviation decreases as the ratio of the source radius to the planetary caustic size and/or the projected separation increases. 4 LENSING BY A WIDE-SEPARATION PLANET In Figure 2, we show the upper limits of normalized source radius ρ⋆/∆x that allow to detect the signature of the host star as a function of the separation when a threshold σth is given. Different symbols represent different threshold σth, that is, σth = 5%, 10%, 15%, and 20% for filled circles, triangles, squares, and pentagons, respectively. The error bars mean the uncertainty on the determination of the up- per limit of ρ⋆/∆x. When s becomes much larger than unity, the magnification pattern can be approximated by the Chang & Refsdal (1979, 1984) lensing, and thus ρ⋆/∆x is nearly constant independent of s. However, the planetary caustic becomes asymmetric and the size of the planetary caustic also becomes bigger as s approaches unity. This is why ρ⋆/∆x rapidly increases as s gets close to unity since the major perturbation regions around the cusps of the plan- etary caustic are located outside of the Einstein ring size of planet mass. Another point to note is that this upper limit on the source-caustic size ratio can be approximated by a fitting function, ρ⋆ ∆x (s) = C1 s√s2 − 1 + C2 for s > 3, (9) with C1 ≃ 0.3/σth√σth and C2 ≃ 0.21/√σth, respectively, which is represented by solid curves. What we have seen here has interesting implications. For instance, this plot can be read to see how the finite source effect constrain the range of the projected separation when the detection threshold is given. In other words, one may obtain for a given source radius and a planet-star mass ratio the upper limit of the star-planet separation that allows the detection of caustic structure due to a host star. Specifically speaking, combining Equations 6 and 9, one obtains (cid:19)1/2 , 2 s 6(cid:18) 1 + √1 + 4R−2 R ≡ −C2 +pC 2 where 2 + C1ρ⋆/q1/2 2C1 (10) (11) . In Figure 3, we show the upper limit of the star-planet 4 Ryu et al. separation to detect the boundness of the wide-separation planet as a function of the source radius and the mass of the planet (left panel), assuming for σth = 5%, and of the detection threshold of fractional deviation and the mass of the planet (right panel), assuming for R⋆ = 1 R⊙. Dotted contours and gray scales represent the upper limit of the separation in units of AU. The planet mass Mp is given in a log scale in units of the solar mass M⊙, the source radius R⋆ in the left panel is in units of the solar radius R⊙. In this specific example, we set Dl = 6 kpc, Ds = 8 kpc, and the lens mass of 0.5 M⊙. According to the left panel, the maximum separation one can tell whether a planetary microlensing feature is caused by a bound planet becomes larger as the radius of the source star becomes smaller and/or the planet mass larger. In other words, as the source star is large and the finite source effect becomes important, one can only detect the existence of its host star when the planet is massive and planet-star separation is small. When 5% of the detection threshold is assumed, for a source star with the radius of ∼ 1 R⊙, an Earth-mass planet (log Mp = −5.5 M⊙) and a Jupiter-mass planet (log Mp = −3.0 M⊙) can be recognized of its boundness when it is within the separation range of ∼ 10 AU and ∼ 30 AU, respectively. Similarly, according to the right panel, the upper limit of the separation becomes larger as the detection threshold becomes lower. The pho- tometric precision of KMTNet project is expected to meet 1% at 21 magnitude in V band on 10 minute monitoring frequency and that of WFIRST mission is 6 1% at 20.5 magnitude in J band on 6 15 minute sampling cadence. Therefore, we expect that the next-generation microlens- ing experiments with high survey monitoring frequency and accurate photometry will discover wide-separation planets with various separations as analyzed in this study. In Figure 4, we compare the separation ranges to de- tect the boundness of the wide-separation planet through the channel of the central and planetary caustics. The black curves represent the upper limits of the separation range in- duced by the planetary caustic for various source radii R⋆ = 1 R⊙, 3 R⊙, 5 R⊙, and 7 R⊙, while the gray curves repre- sent by the central caustic that we calculate following Han (2009a). Results are given both in units of AU and θE. We take the same values of parameters as in Figure 3 for the distances of lens and source, and the lens mass, and assum- ing 5% of the detection threshold. The separation range to detect the boundness of the wide-separation planet induced by the planetary caustic is wider than that induced by the central caustic. What it means is that, when the microlens- ing light curve induced by the planetary caustic happens to be analyzed, one may afford to detect the boundness of the wide-separation planet farther than when that caused by the central caustic is analyzed. 5 CONCLUSION The unbiased spatial distribution of extrasolar planets with respect to host stars is crucial in studying the planet forma- tion processes. Massive and/or close extrasolar planets are likely to be detected by a commonly employed method. On the other hand, the microlensing technique can be sensitive to Earth-like and/or wide-separation planets. To obtain the spatial distribution of extrasolar planets without contami- nating by free-floating planets, it is important to character- ize the planetary microlensing light curves whether they are caused by a wide-separation planet or a free-floating planet. Here, we analyze the condition in terms of lensing pa- rameters, including the size of source stars, under which sig- natures of the host star in the planetary microlensing light curve can be detected. By constructing the fractional devi- ation maps at various positions in the space of microlensing parameters, we have obtained the upper limits of ρ⋆/∆x that allow to detect the signature of the host star as a func- tion of the separation when a threshold is given. We confirm that when s ≫ 1 the Chang & Refsdal (1979, 1984) lensing well-approximate what we have found. We also note that a simple analytical function can be fit, as given in Equation 9. This relation further leads one to a simple analytic condi- tion for the star-planet separation to detect the boundness of wide-separation planets as a function of the mass ratio and source radius, as shown in Figure 3. Finally, we have com- pared the separation ranges to detect the boundness of the wide-separation planet through the channel of the central and planetary caustics. As a result, we conclude that when the microlensing light curve caused by the planetary caus- tic happens to be analyzed, one may afford to support the boundness of the wide-separation planet farther than when that caused by the central caustic is analyzed. Therefore, we conclude that the next-generation microlensing experiments with high survey monitoring frequency are expected to add the number of wide-separation planets through the channel of the planetary caustic. ACKNOWLEDGMENTS We thank the anonymous referee for critical comments which clarify and improve the original version of the manuscript. This research was supported by Basic Science Research Pro- gram through the National Research Foundation of Ko- rea (NRF) funded by the Ministry of Education, Science and Technology (2012R1A6A3A01013815) for YHR and (2012R1A1A4A01013596) for MGP. MGP was also sup- ported by the National Research Foundation of Korea to the Center for Galaxy Evolution Research (NO.2010-0027910). HYC was supported by the National Research Foundation of Korea Grant funded by the Korean government (NRF- 2011-0008123). REFERENCES An J. H., et al., 2002, ApJ, 572, 521 Bachelet E., et al., 2012, ApJ, 754, 73 Batista, V., et al., 2011, A&A, 529, A102 Bennett D. P., 2011, AAS, 43, #318.01 Bennett, D. P., & Rhie, S. H., 2002, ApJ, 574, 985 Bennett D. P., et al., 2012, ApJ, 757, 119 Bennett, D. P., et al., 2008, ApJ, 684, 663 Beaulieu, J.-P., et al., 2006, Nature, 439, 437 Bond, I. A., et al., 2004, ApJL, 606, L155 Bozza, V., 2000, A&A, 355, 423 Chang K., Refsdal S., 1984, A&A, 132, 168 Chang K., Refsdal S., 1979, Natur, 282, 561 Microlensing by a wide-separation planet 5 Di Stefano, R., 2012a, ApJS, 201, 20 Di Stefano, R., 2012b, ApJS, 201, 21 Di Stefano, R., & Scalzo, R. A., 1999a, ApJ, 512, 564 Di Stefano, R., & Scalzo, R. A., 1999b, ApJ, 512, 579 Dong, S., et al., 2009, ApJ, 698, 1826 Ford, E. B., & Rasio, F. A., 2008, ApJ, 686, 621 Gaudi, B. S., et al., 2008, Science, 319, 927 Gould, A., & Loeb, A., 1992, ApJ, 396, 104 Gould, A., et al., 2006, ApJL, 644, L37 Han, C., 2006, ApJ, 638, 1080 Han, C., 2009a, ApJ, 691, 452 Han, C., 2009b, ApJ, 700, 945 Han, C., & Kang, Y. W., 2003, ApJ, 596, 1320 Han C., et al., 2013, ApJ, 762, L28 Han C., et al., 2005, ApJ, 618, 962 Holman, M. J., & Wiegert, P. A., 1999, AJ, 117, 621 Ida, S., & Lin, D. N. C., 2005, ApJ, 626, 1045 Janczak, J., et al., 2010, ApJ, 711, 731 Kains N., et al., 2013, A&A, 552, A70 Kayser, R., Refsdal, S., & Stabell, R., 1986, A&A, 166, 36 Kennedy, G. M., & Kenyon, S. J., 2008, ApJ, 673, 502 Kennedy, G. M., Kenyon, S. J., & Bromley, B. C., 2006, ApJ, 650, L139 Kim S.-L., et al., 2010, SPIE, 7733 Laughlin, G., Bodenheimer, P., & Adams, F. C., 2004, ApJ, 612, L73 Lin, D. N. C., & Ida, S., 1997, ApJ, 477, 781 Malmberg, D., Davies, M. B., & Heggie, D. C., 2011, MN- RAS, 411, 859 Mao, S., & Paczynski, B., 1991, ApJL, 374, L37 Milne, E. A., 1921, MNRAS, 81, 361 Miyake N., et al., 2011, ApJ, 728, 120 Muraki Y., et al., 2011, ApJ, 741, 22 Musielak, Z. E., Cuntz, M., Marshall, E. A., & Stuit, T. D., 2005, A&A, 434, 355 Rasio, F. A., & Ford, E. B., 1996, Science, 274, 954 Schneider, P., & Weiss, A., 1986, A&A, 164, 237 Sumi, T., et al., 2011, Nature, 473, 349 Sumi, T., et al., 2010, ApJ, 710, 1641 Udalski, A., et al., 2005, ApJL, 628, L109 Veras, D., & Raymond, S. N., 2012, MNRAS, 421, L117 Veras, D., Wyatt, M. C., Mustill, A. J., Bonsor, A., & El- dridge, J. J., 2011, MNRAS, 417, 2104 Wambsganss, J., 1997, MNRAS, 284, 172 Weidenschilling, S. J., & Marzari, F., 1996, Nature, 384, 619 Witt, H. J., & Mao, S., 1995, ApJL, 447, L105 Yee J. C., et al., 2012, ApJ, 755, 102 6 Ryu et al. Figure 1. Maps of the fractional deviation in magnifications for various ratios of the source radius normalized by the planetary caustic size ρ⋆/∆x, where ρ⋆ = θ⋆/θE and θ⋆ being the angular radius of a source star, and normalized separation s by the Einstein ring radius θE. The ratio of the source radius to the planetary caustic size ρ⋆/∆x and the separation s are marked on the top of each column and the right-hand side of each row, respectively. The panels are divided by thick black lines with arrows to outline in the parameter space the domains where regions with the amplitude of fractional deviations as indicated by the arrows can be found in the deviation maps. In each map, the contours are the levels of σ = ±1%, ±5%, ±10%, ±15%, and ±20%, and the colors of blue and red tones are the regions with negative and positive σ, respectively, where lighter shades represent smaller σ levels. We show σ levels corresponding to each color in the scale-bar. The black curve at the center in each panel represents the caustic shape. In this particular plot, we set the Galactic distances of lens Dl = 6 kpc and that of source star Ds = 8 kpc. The lens star is located at the origin of coordination, i.e., (x, y)=(0,0). Microlensing by a wide-separation planet 7 Figure 2. Upper limits on the ratio of the source radius to the planetary caustic size for σth = 5%, 10%, 15%, and 20%. The filled circles, triangles, squares, and pentagons represent the maximum values of ρ⋆/∆x as a function of s for different thresholds σth, that is, σth = 5%, 10%, 15%, and 20%, respectively. The error bars mean the uncertainty on the determination of the upper limit of ρ⋆/∆x. The black curve is the fitting function, that is, Equation 9. See text for detailed discussions of this plot. 8 Ryu et al. Figure 3. Upper limit of the star-planet separation to detect the boundness of the wide-separation planet as a function of the source radius and the mass of the planet (left panel ), assuming for σth = 5%, and as a function of the detection threshold of fractional deviation and the mass of the planet (right panel ), assuming for R⋆ = 1 R⊙. Dotted contours and gray scales represent the upper limit of the separation in units of AU. The employed parameters in this example are Dl = 6 kpc, Ds = 8 kpc, and lens mass of 0.5 M⊙. Microlensing by a wide-separation planet 9 Figure 4. Comparison of the separation ranges to detect the boundness of the wide-separation planet through the channel of the central caustics with the planetary caustics. The upper limits of the separation ranges are plotted with the black and gray curves for the planetary and central caustics, respectively. We assume that the the distances of lens and source and the lens mass are Dl = 6 kpc, Ds = 8 kpc, and 0.5 M⊙, and the detection threshold is 5%.
1202.2799
1
1202
2012-02-13T17:29:08
Extrasolar Planet Transits Observed at Kitt Peak National Observatory
[ "astro-ph.EP" ]
We obtained J-, H- and JH-band photometry of known extrasolar planet transiting systems at the 2.1-m Kitt Peak National Observatory Telescope using the FLAMINGOS infrared camera between October 2008 and October 2011. From the derived lightcurves we have extracted the mid-transit times, transit depths and transit durations for these events. The precise mid-transit times obtained help improve the orbital periods and also constrain transit-time variations of the systems. For most cases the published system parameters successfully accounted for our observed lightcurves, but in some instances we derive improved planetary radii and orbital periods. We complemented our 2.1-m infrared observations using CCD z'-band and B-band photometry (plus two Hydrogen Alpha filter observations) obtained with the Kitt Peak Visitor's Center telescope, and with four H-band transits observed in October 2007 with the NSO's 1.6-m McMath-Pierce Solar Telescope. The principal highlights of our results are: 1) our ensemble of J-band planetary radii agree with optical radii, with the best-fit relation being: (Rp/R*)J = 0.0017 + 0.979 (Rp/R*)optical, 2) We observe star spot crossings during the transit of WASP-11/HAT-P-10, 3) we detect star spot crossings by HAT-P-11b (Kepler-3b), thus confirming that the magnetic evolution of the stellar active regions can be monitored even after the Kepler mission has ended, and 4) we confirm a grazing transit for HAT-P-27/WASP-40. In total we present 57 individual transits of 32 known exoplanet systems.
astro-ph.EP
astro-ph
EXTRASOLAR PLANET TRANSITS OBSERVED AT KITT PEAK NATIONAL OBSERVATORY BY PEDRO V. SADA 1, 8 , DRAKE DEMING 2, 8 , DONALD E. JENNINGS 3, 8, BRIAN K. JACKSON 3, 8 , CATRINA M. HAMILTON 4, 8 , JONATHAN FRAINE 2, 8 , STEVEN W. PETERSON 5 , FLYNN HAASE 5 , KEVIN BAYS 5 , ALLEN LUNSFORD 3, 6, 8 , EAMON O’GORMAN 7 1 Universidad de Monterrey, Departamento de Física y Matemát icas, Av. I. Morones Prieto 4500 Pte., San Pedro Garza García, Nuevo León, 66238, México [email protected] 2 Department of Astronomy, University o f Maryland, College Park, MD 20742, USA [email protected], [email protected] 3 Planetary Systems Laboratory, Goddard Space Flight Center, Mail Code 693, Greenbelt, MD 20771, USA donald.e. [email protected], brian.k. [email protected] 4 Dickinson Co llege, Carlisle, PA 17013, USA [email protected] 5 Kitt Peak National Observatory, National Opt ical Astronomy Observatory, 950 N Cherry Ave., Tucson, AZ, 85719, USA [email protected], [email protected], [email protected] 6 The Catho lic University o f America, Washington, DC 20064, USA allen.w. [email protected] 7 Tr inity Co llege, Dublin, Dublin 2, IRELAND [email protected] 8 Visit ing Astronomer, Kitt Peak National Observatory, National Opt ical Astronomy Observatory, which is operated by the Association of Universit ies for Research in Astronomy under cooperative agreement with the National Science Foundation. ABSTRACT We obtained J-, H- and JH-band photometry of known extrasolar planet transit ing systems at the 2.1-m Kitt Peak National Observatory Telescope using the FLAMINGOS infrared camera between October 2008 and October 2011. From the derived lightcurves we have extracted the mid-transit t imes, transit depths and transit durations for these events. The precise mid-transit t imes obtained help improve the orbital periods and also constrain transit-t ime variat ions of the systems. For most cases the published system parameters successfully accounted for our observed lightcurves, but in some instances we derive improved planetary radii and orbital periods. We complemented our 2.1-m infrared observations using CCD z’-band and B-band photometry (plus two Hydrogen Alpha filter observations) obtained with the Kitt Peak Visitor’s Center telescope, and with four H-band transits observed in October 2007 with the NSO’s 1.6-m McMath- Pierce So lar Telescope. The principal highlights of our results are: 1) our ensemble of J-band planetary radii agree with optical radii, with the best-fit relation being: (Rp/R*)J = 0.0017 + 0.979 (Rp/R*)optical, 2) We observe star spot crossings during the transit o f WASP-11/HAT-P- 10, 3) we detect star spot crossings by HAT-P-11b (Kepler-3b), thus confirming that the magnet ic evo lut ion of the stellar act ive regions can be monitored even after the Kepler mission has ended, and 4) we confirm a grazing transit for HAT-P-27/WASP-40. In total we present 57 individual transits of 32 known exoplanet systems. KEYWORDS: Extrasolar Planets 1. INTRODUCTION Many exoplanet systems contain Jupiter-mass planets on close-in orbits. These planets are strongly irradiated by their host stars, and emit significant radiat ion in the infrared (Charbonneau et al. 2005, Deming et al. 2005). Characterization o f their atmospheres using transit and secondary eclipse techniques has become a very act ive field (Seager and Deming 2010). Atmospheric observat ions using secondary eclipse are also sensit ive to the orbital dynamics, specifically the eccentricity of the orbit, via the phase o f the eclipse (Deming et al. 2007). Consequently, interpreting secondary eclipse observat ions requires knowing the ephemeris o f the transits to high precision. Cont inuing explorations of discoveries by transit surveys have given us a sample of more than 70 hot Jupiters transit ing systems brighter than V=13, and the increasing sample size makes it d ifficult to maintain accurate parameters for all systems. Some systems already require addit iona l transit observat ions in order to attain sufficiently precise ephemeredes to interpret secondary eclipse phase (e. g., Todorov et al. 2011). Moreover, the discovery photometry for transit ing planets typically provides only relat ively coarse photometric precision, and fo llow-up photometry with larger aperture telescopes is needed to determine the giant planet radius to a precision limited only by knowledge of the stellar mass (e. g., Winn et al. 2007a). For these reasons, several groups are monitoring known transit ing planets using moderate to large aperture telescopes (e. g., Southworth et al. 2009). In addit ion to the motivations discussed above, we are interested in transit monitoring at near-IR (JHK) wavelengths. Near-IR wavelengths o ffer reduced stellar limb-darkening, and thereby provide an independent check on planetary radii inferred using optica l photometry. Moreover, the composit ion o f exoplanetary atmospheres may cause rea l differences in transit radii with wavelength that could eventually be detected with sufficient ly precise observat ions. The relat ively large sample o f known transit ing systems means that a general transit survey is well matched to classical observing and telescope scheduling methods. In this paper, we present results from a series of classical observing runs, producing high-precision photometry o f several known exoplanet transit ing systems observed in the near infrared (J-, H- and JH-bands), and several less precise transits in the optical. Our photometry was obtained at the Kitt Peak National Observatory using several telescopes. In total, we present 57 lightcurves o f 32 transit ing exoplanetary systems. In sect ion 2 we present the observat ions and data analysis methods used, while in sect ion 3 we describe fitt ing to the photometry using Markov Chain Monte Carlo (MCMC) methodology (Ford, 2005). Section 4 discusses results derived from the model fit s, by comparing our results for planetary radii w ith results in the optical, and we discuss details of individual systems such as improvements to the orbital per iods. 2. OBSERVATIONS AND PHOTOMETRY 2.1 OBSERVATIONS Our primary observat ional system is the Kitt Peak Nat ional Observatory 2.1-meter reflector with the FLAMINGOS 2048×2048-pixel infrared imager and a J-band (1.25 μm), an H-band (2.50 μm), and a J- and H-band combinat ion (JH) filters (E lston 1998). The 0.6 arc-sec per pixel scale yielded a FOV of ~20×20 arc-min with suffic ient comparison stars available for different ial photometry. Fo llowing the conclusion o f night ly public programs, we also had access to the 0.5-meter telescope at the Kitt Peak Visitor Center (the VC telescope). With this telescope we used a 3072x2048 CCD camera at 0.45 arc-sec per pixel and typically a z’-band filter (although in a few instances the data was acquired through a B filter, and twice through a Hydrogen Alpha filter). Most of the observations presented here were obtained between October 2008 and October 2011. We also present four other transits obtained in October 2007 using the National So lar Observatories’ (NSO) 1.6-meter McMath-P ierce Solar Telescope and their NSO Array Camera (NAC). The NAC is a cryogenically cooled 1K×1K InSb Aladdin-III array (Ayres et al. 2008). For these observat ions only one quadrant of the chip was used with only one comparison star available due to the small FOV (~5×5 arc-min). Observations at all three telescopes used various degrees o f defocus to improve the photometric precision, and all used automatic off-axis or manual guiding to maintain point ing stability: Exposure times varied between 20 seconds and 120 seconds, depending on the system used and the stellar magnitude, so that pixel values stayed well within saturation levels. The optical CCD exposures were binned 2×2 to facilitate rapid readout. Flat-field observat ions were acquired at all three observatories using either twilight sky flats, dome flats, or a series o f night-sky exposures that incorporated point ing offsets to allow removal o f stars via a median filter. Standard, dark-field corrections were also applied. 2.2 PHOTOMETRY Subsequent to dark current subtraction and division by a flat-field frame, we performed aperture photometry on the target star and the comparison stars using standard and custom IDL routines. In all cases, except for the McMath-P ierce telescope observat ions, between 2 and 8 stars of similar magnitude to the target star were used for comparison. This allowed for inter-comparison between these stars to make sure no variability was detected in them. Due to the characterist ics o f the heliostat (image rotation) and the small FOV available at the so lar telescope only the comparison star nearest to the target star was of use. The apertures selected to measure the stars and background varied depending on the degree of defocus and seeing condit ions for each observing session. These were chosen such that they minimized the scatter on the final lightcurve. The defocus on the 2.1-m telescope in particular was sensit ive to changes throughout the night, due to mechanical flexure and temperature variat ions. We eventually learned to actively adjust the defocus setting gradually during the observat ions, so as to maintain image stability. For those data that exhibit variable defocus, we adjusted the numerical apertures accordingly in the data analysis process. Best results were also obtained by averaging the ratios of the target star to each comparison star. This produced similar or smaller scattering than the method of ratio ing the target star to the sum o f all the comparison stars. In most cases the comparison stars were of similar brightness (± ~1.5 magnitudes) as the target star. Uncertaint ies for each photometric po int were estimated as the standard deviat ion o f the ratio to the individual comparison stars, divided by the square root of their number (error of the mean). In all cases the observed scatter in the photometry was larger than the estimated formal uncertaint ies, suggesting that the errors in our photometry procedure may be underestimated. This may be due perhaps to inadequate estimat ion o f the uncertainty in the background level in the IDL/ASTROLIB/APER photometry algorithm we use. After normalizing the target star to the comparison stars some gradual variat ions as a funct ion of t ime were found in some instances. In the case of the optical observat ions the variat ion was removed by using a linear airmass-dependent function fit to the baseline before and after the transit. Most transits have at least one hour’s worth of baseline observat ions before the transit ingress and after the transit egress for this purpose. However, the near-IR observat ions exhibited a more complex baseline variat ion that could not be attributed to simple airmass-dependent comparison star different ia l ext inct ion. These are most likely due to telluric waver vapor absorption variat ions, and/or to other instrumental effects. For these cases po lynomial functions of order 2-5 were used to fit the baseline photometry. Most of the near-IR lightcurves included longer pre- ingress and post-egress observations which allowed for improved baseline fit s. Figures 1a and 1b show the near-IR transits observed with the 2.1-m telescope, while Fig. 2a (left panel) shows the four transits observed with the McMath-P ierce Telescope, and Figs. 2a and 2b show the optical transits observed with the Visitor Center Telescope. During the observing runs other transits were recorded as well, but they are either incomplete (show the ingress or egress only) or suffered from clouds, and are therefore of limited use and not included in this work. 3. MODELING In order to fit the observed transit lightcurves we first created init ial standard mode l lightcurves. These were constructed numerically as a tile-the-star procedure using the Binary Maker II software (Bradstreet 2005). The init ial system parameters used were obtained from the latest literature available. Linear limb-darkening function coefficients were taken from Claret (2000). For most cases the init ial model lightcurves yielded very good agreement with the observed ones and only small adjustments to the duration and depth of the model transits were necessary to optimize the fits. This was done by applying small (< 2% on average) mult iplicat ive factors to both the depth and durat ion of the model transit. In mo st cases these small corrections fall within the published uncertaint ies. However, a few o f the transits exhibited larger model deviat ions and further study was required. These are explained for the individual systems in sect ion 4.2. Table 1 presents all the observed transits and the principal lightcurve parameters (mid- transit time, depth, and durat ion) derived from fitt ing the models as explained above. Because the scatter of the photometry is larger than the formal errors suggest, for each lightcurve we used the scatter to estimate the uncertaint ies. In order to derive improved orbital periods for the transit ing systems we ut ilized the transit timings reported in the published literature, including the observat ions from this paper, and we implemented a least-squares linear fit to the data, weight ing the individual transit times by their uncertaint ies. When relevant we have converted reported Heliocentric Julian Dates (HJD) to Barycentric Julian Dates (BJD) (Eastman, Siverd & Gaudi 2010) and have used the Dynamical T ime-based system (BJD_TDB) instead of the Coordinated Universal T ime-based system (BJD_UTC). The difference between heliocentric and barycentric times can be up to about four seconds, but most often it is less than this value and well within the individual timing uncertaint ies reported, so it has limited effect on the derived periods. However, the difference between UTC-based and TDB-based t imings is a systemat ic offset which depends on recent addit ions of leap seconds to UTC. To convert BJD_TDB to BJD_UTC subtract 0.000766 days for transits observed after 1 January 2009 (JD 2,454,832.5), 0.000754 days from transits between 1 January 2006 (JD 2,453,736.5) and 1 January 2009, and 0.000743 days from transits between 1 January 1999 (JD 2,451,179.5), and 1 January 2006. The result ing system periods are presented in Table 2 along with the number of mid-transit timings used and the time span between the first and last observat ion reported. The reference epoch presented (JD0) is the result from the fit and generally corresponds with the first reported transit found in the literature. For some of the best lightcurves obtained with the KPNO 2.1m telescope (J- and JH- band) we proceeded to fit theoretical transit curves to the transit data using the Transit Analys is Package (TAP) which uses Bayesian probability distribut ions with Markov Chain Monte Carlo (MCMC) techniques (Gazak, Tonry & Johnson, 2011). For this modeling we fixed the period and the eccentricity o f each system, and used quadratic limb darkening coefficients for the stars from Claret (2000). We ran 5 chains with 105 samples for each system. During testing of the software we found out that extending the modeling to 106 samples did not yield significant improvement on the fit, not just ifying the larger comput ing requirements. There was also no significant difference in using either a single linear (u) or two quadratic (a & b) coefficients to describe the stellar limb- darkening, in accordance with Southworth (2008) who found out that linear limb darkening was adequate for the analysis o f high quality ground-based data. The derived lightcurve parameters of interest are reported in Table 3: the orbital inclinat ion (i), the planet-to-star radius ratio (Rp/R*) and the scaled semimajor axis (a/R*). 4. RESULTS AND DISCUSSION We here discuss the results o f modeling our 2.1-meter J-band transit lightcurves. Sec. 4.1 discusses the overall comparison to planetary radii derived at optical wavelengths and Sec. 4.2 discusses individual systems. While we generally find good agreement with optical results, some o f our J-band lightcurves do suggest variations from the published models which require further study. However, some o f the systems in which we detect differences in the inclinat ion and the scaled semimajor axis may have degenerate solut ions since these two parameters trade off against each other, particularly for IR data where the stellar disk has litt le limb darkening. All the observations presented in this paper will be made available to the public through the NASA/IPAC/NExSci Star and Exoplanet Database (NStED) (or are available from the main author upon request). Both raw data and fitted data will be made available since we realize that the system parameters are sensit ive to the baseline modeling method chosen. The simple method selected in this paper was pr imary aimed at establishing an accurate mid-transit t ime. 4.1 COMPARISON WITH OPTICAL PLANETARY RADII Stellar limb darkening is much less prominent in the near-infrared as compared to optica l wavelengths, so the depth of an infrared transit is closely proportional to (Rp/R*)2. One significant source of uncertainty in planetary radii from IR transit curves is the definit io n of the photometric baseline, which can vary in broad-band IR observat ions due to differences in spectral type between the target star and the comparison stars (e.g., Deming et al. 2011). However, since this source of error should vary independent ly from one transit ing system to another, it represents random - not systemat ic - error when comparing an ensemble o f planetary radii at optical and IR wavelengths. Figure 3 plots Rp/R* from our 2.1-m J-band transits (Table 3) against published optical radii. Potentia l sources of systemat ic differences in near-IR versus optical planetary radii include errors in stellar limb darkening, stellar act ivity (which affects the near-IR data less than the optical), as well as potential real d ifferences in p lanetary radii w ith wavelength. The latter could be produced, for example, by high alt itude haze (S ing et al. 2011) which increases transit radii at the shortest wavelengths. The large stellar photon flux available to optical observers usually produces significant ly smaller random errors than for our J-band transits. Consequently, we make the approximat ion that all of the error lies in our J-band values for Rp/R*, and we perform an error-weighted linear least squares fit, shown by the solid line in Fig. 3. This fit yields: The 1-sigma error on the slope of this relat ion is ±0.025, so the best-fit relation differs from the null hypothesis by less than 1-sigma. If we omit the three seemingly discrepant systems having the largest J-band errors (HAT-P-4, WASP-2 and WASP-48), the slope becomes 0.9765 ± 0.025, still better than 1-sigma agreement with the null hypothesis. We conclude that the optical radii are not likely to be affected at the approximately 5-percent level (>2-sigma) by errors in optical limb darkening. As for real differences in exoplanetary radii as a funct ion of wavelength, we note that the variat ion in Rp/R* seen by Sing et al. (2011) for HD189733b is a 1.3-percent effect from the near-IR to 400 nm. That variat ion remains beyond the sensit ivity o f our statist ical relation, especially since many of the optical radii apply to wavelengths longward of 400 nm (e.g., R-, I-, or z’- band). Nevertheless, we are encouraged by the prospect that further improvements in our (Rp/R*)J = 0.0017 +0.979 (Rp/R*)optical J-band photometry (higher precision, and more transits per system), combined with improved ground-based precis ion in the bluer near 400 nm, could potentially provide ground-based statist ical detection of haze in giant exoplanetary atmospheres, whereas such detections have so far required space-borne observat ions. 4.2 DISCUSSION OF INDIVIDUAL SYSTEMS 4.2.1 COROT-1 The brightness of the CoRoT-1 (V=13.6, K=12.1) star system was beyond our practica l observing limit and thus resulted in a lightcurve with considerable scatter (Fig. 1a), and thus no attempt was made to derive the system parameters. It is however st ill o f use in confirming its previously published period because of the extended time coverage. For this we used the original discovery transit s o f Barge et al. (2008) as timed by Bean (2009), a low-precision prediscovery transit of Rauer et al. (2010), and a high-precision transit observed by Gillon et al. (2009b) at the VLT. The Gillon et al. ephemeris predicts a midtransit time for our observat ion o f about BJD 2,455,162.91696 ± 0.00024. Our observed midtransit t ime o f 2,455,162.91621 ± 0.00059 falls somewhat short of the predict ion, but our relat ively large uncertainty st ill allows for a match. Bean (2009) and Csizmadia et al. (2010) had found no significant periodic timing variations with a period shorter than the original observat ional window of 55 days. The Gillon et al. (2009b) observat ion and the present one extend this t ime period and are also consistent with a fixed period for the planet. Including our observation we find a period for CoRoT-1b of 1.5089682 ± 0.0000005 days. 4.2.2 COROT-2 CoRoT-2 is a transit ing system which exhibits clear evidence o f starspots that have been used to estimate the rotation period of the star (Lanza et al. 2009). Alonso et al. (2008) in their discovery paper present an ephemeris which summarizes the 78 transits observed by the CoRoT mission. We combined this ephemeris with prediscovery transits reported by Rauer et al. (2010), one ephemeris reported by Vereš et al. (2009) and our current measurement (Fig. 2a) to calculate the per iod o f the system. Our result o f P=1.7429971 ± 0.0000011 days agrees with the original period of 1.7429964 ± 0.0000017 days reported by Alonso et al. (2008) and further reduces the uncertainty due to the longer time baseline. However, the earliest prediscovery transit of Rauer et al. (2010) exhibits a large deviat ion (O-C ~ 23.3 minutes) from its predicted transit time, well outside its own 3- sigma uncertainty range, and is thus considered suspect. Eliminat ing this single anomalous observat ion results in an alternate preferred solut ion o f P=1.7429981±0.0000011 days. 4.2.3 GJ 1214 The first two of the observat ions presented here have already been analyzed and discussed in Sada et al. (2010). See this reference for a more thorough modeling analys is of this system. However, since then other transits have been reported. Here we assemble the original midtransit observat ions o f Charbonneau et al. (2010), as re-evaluated in Berta et al. (2011), other observat ions also reported in Berta et al. (2011) including two high- precision VLT transit s, and those o f Sada et al. (2010). We also include in the period solut ion 12 new full transit s (ingress and egress recorded) presented by Carter et al. (2011), three transits reported by Kundurthy et al. (2011) (their best result: chain003a), four near- infrared transits observed from Hawaii (Croll et al. 2011), plus two recent unreported transit we observed simultaneously at KPNO using the 2.1m telescope (J- band, Fig. 1a) and the VC telescope (z’-band, Fig. 2a). From these 33 transits we derive a period of P=1.5804048±0.0000002 days, in complete agreement with other recent calculat ions and, within observat ional uncertaint ies, with no evidence o f variat ion during the first two observing seasons. Modeling o f the system parameters from the first 2.1m lightcurve, confirming the reported planet radius, is also described in Sada et al. (2010). We have not yet attempted to fit the second high-precision lightcurve, pending further observat ions of the system. 4.2.4 HAT-P-1 Two HAT-P-1 transits were observed with the NAC through an H-band filter on the NSO/KPNO McMath-P ierce So lar Telescope on 2007 October 08 and October 17 (Fig. 2a). These were part icularly difficult observat ions because the image field rotation inherent to a helio stat slowly changed the reflect ing mirror surface areas throughout the long observing period. Because of this, only one star was of use as a comparison source. Fortunately HAT-P-1 has a close companion. Even so, on the night of 2007 October 17 high winds hitt ing the main heliostat mirror introduced severe no ise on the data. We used our data along with the low-precision Bakos et al. (2007) discovery paper observat ion, the midtransit t imes of Winn et al. (2007b) as corrected in Winn et al. (2008), and the Johnson et al. (2008) reported transits to derive a period of 4.4653054±0.0000069 days. In this calculat ion we did not include our 2007 October 17 observat ion since it deviated more than 3-sigma from its predicted value, probably as a result of the severe wind problem. 4.2.5 HAT-P-3 We observed two transits of HAT-P-3 on 2009 May 15 and 2010 May 27. Each transit was observed with the KPNO 2.1m telescope with a JH filter (Figs 1a) and also with the KPNO VC telescope through a z’ filter the first night and a B filter on the second (Fig. 2a). We combined our derived midtransit t imings with those o f Torres et al. (2007), Gibson et al. (2010), Chan et al. (2011) and Nascimbeni et al. (2011) to derive a period of 2.8997382±0.0000009 days, in agreement with recent calculat ions but with improved uncertainty. The 2009 May 15 lightcurve in part icular exhibit s a gap in the data during egress and was not modeled. The 2010 May 27 lightcurve on the other hand yields a model with slight ly lower orbital inclination (~85.7o±0.55o vs ~87.1o±0.55o) than the one reported in Chan et al. (2011) and Torres et al. (2008). We also obtain a smaller scaled semimajor axis (~9.2±0.5 vs ~10.4±0.5), which combined probably explains the slight ly shorter duration (~5 min) observed for this particular transit compared with Chan et al. (2011). 4.2.6 HAT-P-4 We observed one transit for HAT-P-4 on 2011 May 22 with the KPNO 2.1m telescope using a J-band filter (Fig. 1a). We combined our midtransit timing with the two found in the discovery paper (Kovács et al. 2007), one reported observat ion by Winn et al. (2010), and ten addit ional EPOXI observat ions (Christiansen et al. 2011) to derive a period of 3.0565254±0.0000012 days. This differs from 3.0565114±0.0000028 days reported by Christ iansen. In this part icular case we can trace the difference to a registered discontinuity in our data (possible due to a temporary gain fluctuation in the camera amplifiers) just at egress that had to be corrected manually. The depth of the transit is so shallow that any variat ion at the ingress/egress portions o f the lightcurve is crit ical in determining the transit durat ion. We attempted to compensate empirically for it by matching the comparison star brightness after egress. However, this makes our egress portion of the lightcurve suspect and could account for our unusual short transit duration and shift of the midtransit time. For this system we also derive a smaller orbital inclinat ion (~86.0o vs ~89.8o) and scaled semimajor axis (~5.6 vs ~6.0) from those reported by Torres et al. (2008) and Christiansen et al. (2011) which is also reflected by the observed smaller (by ~10 min) transit duration. 4.2.7 HAT-P-6 We observed one transit of HAT-P-6 on 2009 November 25 with the KPNO 2.1m telescope through a J-band filter (Fig. 1a). We have combined our midtransit t iming with those of the discovery paper (Noyes et al. 2008), as reevaluated in Szabó et al. (2010), and a newer transit also reported by Szabó et al. (2010), to derive an improved period of 3.8530018±0.0000015 days. The only reported model parameters for this system correspond to the original discovery paper by Noyes et al. (2008). Our results do vary slight ly from those published init ially and may be an alternat ive, though addit iona l observat ions of this system are needed to improve on our uncertainties. 4.2.8 HAT-P-11 We observed one transit o f HAT-P-11 with both the KPNO 2.1m telescope (J-band) and VC telescope (B-band) on 2010 June 01 (Figs. 1a and 2a). The depth of this transit is rather shallow but well defined by the good observing condit ions at the time. These observat ions are detailed further in Deming et al. (2011). In addit ion we also observed a transit of HAT-P-11 with the VC telescope (B-band again) on 2011 May 14. We used our derived timings along with those reported by Bakos et al. (2010b), Ditmann et al. (2009), Hirano et al. (2011), and the Kepler mid-transit observat ions analyzed in Deming et al. (2011) to derive an improved period o f 4.8878056±0.0000015 days. Although our reported mid-transit timings for June 2010 may be suspect (see Deming et al. 2011), they have litt le weight against the high-quality Kepler data. Of particular interest is the VC telescope transit observed on 2011 May 14 (Fig. 2a). We chose to observe through the B- filter so as to obtain greater sensit ivity to possible stellar activity due to the shorter effect ive wavelength of the B filter. Our intent was to observe possible star spots as the planet, which crosses the stellar disk nearly perpendicular to its equator, covered the lat itudes of interest on the surface of this known active star (see Deming et al. 2011, and Sanchis-Ojeda et al. 2011, for further details). From careful observation of the lightcurve we do see a definite feature at ~+0.019 days which corresponds with one of the locat ions where starspots were seen in the Kepler data. This confirms that the lat itudes of stellar activity on this star can be monitored with ground-based observations using short- wavelength filters. Even after the end of the Kepler mission, ground-based observat ions of the transits can be used to monitor the evo lut ion of magnet ic act ivity on this active star (Deming et al. 2011). 4.2.9 HAT-P-12 The only o ther midtransit ephemeris for HAT-P-12 available in the literature corresponds to the discovery paper by Hartman et al. (2009). Combined with our 2.1m telescope J- band observat ion presented here (Fig. 1a) we obtain an improved period o f 3.2130553 ± 0.0000010 days. More transit timings need to be reported in order to further improve the period of this system. Our derived model parameters from Table 3 are in agreement with those reported in Hartman et al. (2009) although our uncertaint ies are larger. 4.2.10 HAT-P-27 / WASP-40 We combined our single recent observat ion of this transit ing system (KPNO 2.1m telescope, J-band, Fig. 1a) with the epoch reported in the discovery papers by Béky et al. (2001) and Anderson et al. (2011) to derive an improved period o f 3.0395824 ± 0.0000035 days. Our derived model parameters from Table 3 are in agreement with those reported in the discovery papers although our uncertaint ies are larger. Anderson et al. (2011) report a 40% probability that the transits occur in a grazing configurat ion. The minimal limb darkening in the near-IR normally yields transits that are quite flat- bottomed. In contrast to this normal behavior, the noticeable roundness of our J-band transit curve (Fig. 1a) confirms that the transit is grazing. 4.2.11 HAT-P-32 One ephemeris for this system is reported in the literature result ing from the observat ion of several transits in the discovery paper (Hartman et al. 2011). We combined our three KPNO 2011 observat ions (two 2.1m telescope J-band transit s on October 09 and 11 – Fig. 1a, and a VCT z’-band one on October 11 – Fig. 2a) to derive a period of 2.1500103 ± 0.0000003 days with improved uncertainty. Our simultaneous observat ions of 2011 October 11 through a z’-band and a J-band filters both exhibit a small brightness increase just after midtransit that is not evident in the J- band observat ion from the previous planetary transit observed two days before through a J-band filter, and it is probably associated with starspot activity. The model result s reported in Table 3 correspond only to modeling the 2011 October 09 lightcurve since it had a higher S/N ratio and did not have the starspot. Our result ing model inclinat ion and scaled stellar radius agree with those reported by Hartmann et al. (2011), but our planet- to-star size ratio seems to be a slight ly larger. 4.2.12 HD 17156 HD 17156 is a system with a relat ively long orbital period (~21.2 days) and thus transit opportunit ies are infrequent and observat ions are valuable. We registered one transit o f this system with the KPNO VC telescope on 2009 Nov. 24 (Fig. 2a). This star has a high northern declinat ion, beyond the limit of the 2.1m telescope, and reachable only using the german equatorial mount of the VC telescope. We also gather the first reported transit observat ion by Barbieri et al. (2007), the midtransit s reported by Irwin et al. (2008), Narita et al. (2008), Gillon et al. (2008), Winn et al. (2009) and the high-quality HST observat ions analyzed in Nutzman et al. (2011) to report a period of 21.216384±0.000016 days for the system. The so lut ion is dominated by the three high-quality HST observat ions and our result deviate by about 150 seconds from the predict ion based on the derived ephemeris. However, this difference is st ill within our measurement uncertainty. 4.2.13 HD189733 This is a well-observed transit ing system with ample reported transits. We observed HD 189733 with the 2.1m telescope (J-band, Fig. 1a) and the VC telescope (Fig. 2a) on two different occasions. The latest VC telescope observat ion is o f particular interest since, due to a filter wheel error, we observed it through an Hα filter. Careful analys is o f the observed lightcurve reveals variations from a symmetric lightcurve due to chromospheric stellar features as the planet transits the disk of the active star. For this system we obtained all ground-based (Bouchy et al. 2005, Bakos et al. 2006, Winn et al. 2007b, Hrudková et al. 2010) and spacecraft (Pont et al. 2007, Knutson et al. 2007, Miller-Ricci et al. 2008, Knutson et al. 2009, Agol et al. 2010) midtransit times reported in the literature to derive a period of 2.2185754 ± 0.0000001 days, consistent with the latest Agol et al. (2010) estimate based solely on Spitzer observations. This is an o ften-studied bright system that has been observed from space and its lightcurve parameters are well constrained. Our single lightcurve observat ion cannot improve on those results, but it was o f particular help in refining the data analys is and modeling techniques used throughout this work. Within our much larger uncertaint ies, our lightcurve model parameters correspond with those of the literature. 4.2.14 Qatar-1 Our single KPNO 2.1m J-band observat ion of this system (Fig. 1a) was combined with the discovery article ephemeris (Alsubai et al. 2010) to derive an improved period of 1.4200227 ± 0.0000012 days. The results o f our modeling are in fair agreement with those reported by Alsubai et al. (2010) assuming a circular orbit. 4.2.15 TrES-1 There are few midtransit t imes reported in the professional literature for this system despite this being one o f the earliest exoplanet systems discovered and announced. We managed to observe two consecutive transits in 2007 using the NAC array at the KPNO NSO McMath-P ierce Solar Telescope (H-band, left panel Fig. 2a). We only used one comparison star in the photometry because of problems o f image field rotation due to the design nature of the helio stat telescope used. Fortunately this system has a close comparison star of similar magnitude that yielded usable lightcurves. Using our two observat ions and those reported in the literature (Alonso et al. 2004, Charbonneau et al. 2005, Narita et al., 2007, Winn et al, 2007c, Hrudková et al, 2009, Raetz et al. 2009b, and Rabus et al. 2009) we obtain a period o f 3.0300724±0.0000004 days, slight ly improving on the Rabus et al. (2009) latest value. However, we note that a longer timeline of transit s observat ions, like the large amateur collect ion found in the Exoplanet Transit Database (ETD - http://var2.astro.cz/ETD/ ), is needed and might st ill yield a slight ly different period. 4.2.16 TrES-2 We obtained two transits o f this well-observed and characterized exoplanet system in 2011 with the both KPNO telescopes on May 20 (VCT z’-band, Fig. 2b) and October 13 (2.1m J-band, Fig. 1a). Combining our midtransit t imes with those reported from ground- based observat ions (O’Donovan et al. 2006, Holman et al. 2007, Raetz et al. 2009a, Mislis & Schmitt 2009, Mislis et al. 2010, Rabus et al. 2009, Scuderi et al. 2010 & Co lón et al. 2010) and Kepler (Kipping & Bakos 2011) and EPOXI (Christ iansen et al. 2011) spacecraft data we obtain a per iod of 2.4706128±0.0000003 days. Our period matches the one der ived from combining several observat ions from d iverse sources over a long per iod (Christ iansen et al. 2011) but is less than the period derived using short-term, but very high-quality, Kepler spacecraft observat ions (Kipping & Bakos 2011). Our single near-IR lightcurve modeling cannot compare with better quality data available for this well studied system. However, our results do agree very well with those published for this system, w ithin our larger uncertaint ies. 4.2.17 TrES-3 We observed one transit o f TrES-3, with the KPNO VC telescope (z’-band, Fig. 2b). We gathered the midtransit times reported in the literature by Sozzetti et al. (2008), which includes a reevaluation o f the midtransit t ime from the O’Donovan et al. (2006) announcement paper, Gibson et al. (2009), Colón et al. (2010), Christ iansen et al. (2011) and Woo-Lee et al. (2011) to derive a period of 1.3061865±0.0000002 days for this system. This period agrees with the one presented by Christiansen et al. (2011) and Vaňko et al. (2011), but is slight ly shorter than the 1.30618700±0.00000015 day period reported by Woo-Lee et al. (2011) that includes a large number of amateur observat ions. 4.2.18 TrES-4 We observed one transit of TrES-4 with both the KPNO 2.1m telescope (J-band, Fig. 1b) and VC telescope (B-band, Fig. 2b) on 2010 May 30. The depth o f this transit is shallow but well defined by the good observing condit ions at the t ime. Unfortunately there was also a timing issue that night with the 2.1m telescope software and we cannot derive a trustworthy midtransit time for that observat ion. We therefore use our single VC telescope timing along with those reported by Sozzetti et al. (2009), which are re- analyzed original observat ions from Mandsushev et al. (2007), and those of Chan et al. (2011) to derive an improved period of 3.5539303±0.0000019 days. Our derived lightcurve model parameters from Table 3 are also in agreement, within our uncertaint ies, with those recently reported by Chan et al. (2011). 4.2.19 WASP-1 We observed WASP-1 on one occasion with the KPNO 2.1m telescope (J-band, Fig. 1b). This system has few midtransit observat ions reported in the literature. We used the discovery ephemeris (Co llier Cameron et al. 2007) along w ith the midtransit reports from Charnonneau et al. (2007), Shporer et al. (2007), Wang et al. (2008) and Albrecht et al. (2011) to derive a period of 2.5199425±0.0000014 days. In our analysis we did not include the midtransit report from Szabó et al. (2010) because it deviates by over 14 minutes from the predicted time, well outside of its reported uncertainty. Thus we consider the time suspect. Our derived parameters for this system fall just within those lately reported by S impson et al. (2011). 4.2.20 WASP-2 We observed WASP-2 on one occasion with the KPNO 2.1m telescope (J-band, Fig. 1b). This system also has few midtransit observat ions reported in the literature. We used these (Collier Cameron et al. 2007, Charbonneau et al. 2007, Hrudková et al. 2009 and Southworth et al. 2010) to derive a period of 2.1522213±0.0000004 days, in agreement with the report of Southworth et al. (2010) that also includes a large number of amateur observat ions. Southworth et al. (2010) also do a thorough job of comparing various model parameters reported in the literature, and our results agree within our larger uncertaint ies. 4.2.21 WASP-3 WASP-3 is a well observed system with ample mid-transit timings found in the literature. Here we gathered the earlier observat ions from Pollaco et al. (2008), Gibson et al. (2008), a low-precision measurement by Damasso et al. (2009), well observed events by Triapathi et al. (2010), Maciejewski et al. (2010), Littlefield (2011), and space-based observat ions by Christ iansen et al. (2011) and combine them with our four observat ions (three from the VC telescope and one from the 2.1m telescope – Figs. 1b and 2b) to derive an improved period of 1.8468332±0.0000004 days. In particular Maciejewski et al. (2010) propose that the observed mid-transit time deviat ions from a constant period may be due to the possible presence of a ~15 earth-mass planet located close to the 2:1 mean mot ion resonance. The observations from Littlefield (2011) seem to support in part this conclusion. Our transits do not directly confirm these t iming variat ions. The measured midtransit uncertaint ies for lightcurves, obtained using different the methodologies, accompanied with t ime synchronizat ion and standardizat ion issues, are just large enough to confuse the issue. Our modeling results also agree with those reported in Christ iansen et al. (2011) from EPOXI observat ions. 4.2.22 WASP-6 We have combined our 2011 October 12 J-band KPNO 2.1m midtransit t ime (Fig. 1b) with the one reported on the discovery paper (Gillon et al. 2009a) and the observat ion from (Dragomir et al. 2011) to derive an improve period of 3.3609998 ± 0.0000011 days. Our model for this system agrees with the inclination and semimajor axis reported by Gillon et al. (2009a) and Dragomir et al. (2011). However, we obtain a smaller planet-to- star ratio in our near-IR observat ions. 4.2.23 WASP-10 WASP-10 was our first system observed transit ing at both the KPNO 2.1m telescope (J- band, Fig. 1b) and the VC telescope (z’-band, Fig. 2b). Our midtransit times differ from one another primarily due to the relative scarcity o f data during egress at the VC telescope, although they st ill overlap at the 3-sigma level. We gathered other exist ing midtransit times from the discovery paper by Christ ian et al. (2009), from the high- quality observat ion by Johnson et al. (2009) corrected in Johnson et al. (2010), from Dittman et al. (2010), and from Maciejewski et al. (2011a) and Maciejewski et al. (2011c). Maciejewski et al. (2011a) also reanalyzed four midtransit t imes presented in Krejčová et al. (2010). All these observat ions yield a period of 3.0927297±0.0000003 days. Maciejewski et al. (2011a) report midtransit t iming variat ions which can be explained by the presence of an addit ional planet about a tenth of the mass of Jupiter orbit ing close to the outer 5:3 mean motion resonance with a period of about 5.23 days. Our single observat ions at epoch 129 (see their Figure 2) would yield an O-C of about +2.35 minutes from their linear ephemeris. Our observed deviat ions are ~+0.7±0.3 minutes for the 2.1m and ~+2.5±0.6 minutes for the VC telescope. Although the VC telescope observat ion might agree with the predict ion, we have a larger confidence in the mid- transit time derived from the 2.1m telescope lightcurve. Thus we are unable to either support or deny the claim of a second planet in the system. In fact, our higher qualit y 2.1m observat ion matches the Maciejewski et al. (2011c) four high-quality observat ions very well if we use the per iod der ived here. An alternat ive, and more general, explanat ion could be that most midtransit times are susceptible to larger variat ions than their forma l uncertaint ies reflect because o f either incomplete lightcurves and baseline trends due to stellar variability and/or different ial ext inct ion o f the field comparison stars which have not been ident ified and accounted for properly. There has been some discussion in the literature concerning the radius of WASP-10b (Johnson et al. 2009, Dittman et al. 2010). Our J-band transit result yields a planetary to stellar radius ratio in close agreement with Johnson et al. (2009). 4.2.24 WASP-11/HAT-P-10 We observed four transits o f this system on 2009 November 26 with the KPNO 2.1m telescope, on 2010 November 07 with the KPNO VC telescope, and on 2011 October 08 with both telescopes (See Figs. 1b and 2b). We combined our midtransit t imings wit h those reported on both discovery papers (West et al. 2009 and Bakos et al. 2010) to derive an improved per iod of 3.7224793±0.0000007 days. Our higher quality J-band lightcurve from 2009 November 26 shows an overall depth o f 1.8±0.2%, which is smaller than the reported depth for this system. In addit ion, the shape of the bottom of the transit is not smooth and does not match the fitted model well. Specifically, our observed depth seems to be greater just before the beginning of transit egress, which suggests that the real depth is indeed larger and that spots could be present on the stellar surface. To test this theory we ran several models in which we adopted the best system parameters and then proceeded to spot the stellar surface during the first part of the transit. The result is presented in Fig. 4. The so lid line represents the standard literature model while the dotted line is our best overall fit to the observed depth of the lightcurve. The dashed line is the standard literature model with part of the stellar surface covered by starspots with temperatures ~6% lower than the effect ive temperature of the star (4625 K vs 4920 K). Although we po int out that this so lut ion is not unique and no effort was made to find the best fit possible. Note how the middle part of the transit and the brightness drop at phase ~0.025 is better matched by the spotted model. Spectroscopy of the star by Knutson et al. (2010) showed that it exhibited the 5th most intense K-line emission of 50 stars in their sample, a clear indicator of stellar act ivity. Further evidence that the star may exhibit starspots is found in the Exoplanet Transit Database (Poddany, Brat & Pejcha 2010), a compilat ion of pro fessional and amateur planetary transit lightcurves. Reported observat ions for this system show variat ions in recorded transit depths that go from about 16 to 26 milimagnitudes, Our observed transit corresponds to their epoch #108, close in time to a reported depth of ~16 milimagnitudes on transit #106. Our KPNO VC telescope z’-band transit one observing season later shows a deeper 2.1% transit, although with higher uncertainty; and both transit lightcurves from 2011 October 08 show a depth of ~1.9%. From this we conclude that the star exhibits starspot activit y that affects the transits and thus the derived model parameters. On Table 3 we report our model results from both of our J-band lightcurves which agree very well with each other, and fall somewhere in between those reported by both discovery papers. 4.2.25 WASP-12 We observed one transit o f WASP-12 with the KPNO VC telescope (Fig. 2b, z’-band). We have combined our observat ion with those published by Hebb et al. (2009), Chan et al. (2011) and Maciejewski et al. (2011b) to derive a period o f 1.0914224±0.0000003 days in agreement with the last two publications. 4.2.26 WASP-24 We observed one transit o f WASP-24 with the KPNO 2.1m telescope (J-band, Fig. 1b) and combine our midtransit t ime with those reported in the discovery paper by Street et al. (2010) to derive a period of 2.3412162 ± 0.0000014 days. Simpson et al. (2011) publishes model parameters that match those of Street et al. (2010). Our derived result s seem to deviate significant ly, showing a larger system inclinat ion (~86.4o vs. 83.6o) compensated by a larger scaled semimajor axis (~7.1 vs. 6.0), although our larger uncertaint ies might still allow for a match. The planet size is similar in both studies. 4.2.27 WASP-32 We observed one transit of WASP-32 through a J-band filter with the KPNO 2.1m telescope on 2011 October 15 (Fig. 1b). We combine our result with the ephemeris presented in the discovery paper (Maxted et al. 2010) to derive an improved period of 2.7186591 ± 0.0000024 days. Our model results agree with those reported by Maxted et al. (2010) with respect to the inclinat ion and scaled semimajor axis o f the system. Our derived planet-to-star radius ratio (0.1030) is smaller than their reported radius ratio (0.1113) by about 3-sigma. 4.2.28 WASP-33 WASP-33 is the hottest known hot Jupiter (Smith et al. 2011) closely orbit ing a bright delta Scuti variable host star (Herrero et al. 2011). Only two epochs are reported in the literature: in the discovery paper (Collier Cameron et al. 2010) and by Smith et al. (2011). We combine these midtransit times with our observat ions: a z’-band lightcurve from the KPNO VC telescope on 2010 November 03 and two other lightcurves obtained on 2011 October 13 (a J-band one at the 2.1m telescope, and a Hydrogen Alpha observat ions with the VCT), shown in Figs. 1b and 2b, to derive a period of 1.2198721±0.0000003 days. Both observat ions obtained on 2011 October 13 clearly show short-period variability that interferes with a clean determinat ion of the lightcurve parameters on Table 1 and the near-IR model parameters of Table 3. This is evidenced as an offset of the baseline before ingress compared with the egress baseline, and there is also an increase in brightness affect ing the first half o f the transit depth. These effects are enhanced on the shorter wavelength and narrower bandpass Hydrogen Alpha lightcurve that shows the transit o f the planet against the chromosphere of the star compared with the J-band one which shows the same transit against a smoother and less limb-darkened stellar disk. Simultaneous observat ions o f the same transit at different wavelengths can help limit stellar surface characterist ics ( like wavelength-dependent brightness amplitude variabilit y in this case). Because the above ment ioned stellar variability was not accounted for, the result ing model parameters in Table 3 are suspect. 4.2.29 WASP-48 There is only one discovery epoch reported for this system (Enoch et al. 2011). We combine this midtransit time with our KPNO 2.1m J-band observat ion (Fig. 1b) to derive a period of 2.1436283±0.0000041 days. For this system we also derive parameters that seem to deviate from the ones reported in the discovery paper by Enoch et al. (2011). Our inclinat ion seems to be larger (~85.1o vs. ~80.1o) also fo llowed by a larger scaled semimajor axis (~5.4 vs. ~4.2). However, the planet size seems to be the same between both studies. 4.2.30 WASP-50 We combined our two consecut ive 2011 observations of this system from KPNO (z’- band with the VCT on October 15 – Fig. 2b, and J-band with the 2.1m on October 17 – Fig. 1b) with the reported ephemeris on the discovery article (Gillon et al. 2011) to derive an improved period of 1.9550905 ± 0.0000022 days thanks to lengthening the t ime baseline of observat ions. Our modeling result s are also consistent with those reported by Gillon et al. (2011). 4.2.31 XO-1 We observed one transit of this system with the KPNO VC telescope (z’-band, Fig. 2b). We combined our midtransit timing with those reported by McCullough et al. (2006), Holman et al. (2006), Wilson et al. (2006), Vaňko et al. (2009), Raetz et al. (2009b), Cáceres et al. (2009), and the HST observat ions by Burke et al. (2010) to derive a period of 3.9415052±0.0000008 days in agreement with the latest report by Burke et al. (2010). 4.2.32 XO-5 We observed one transit of XO-5 with the KPNO 2.1m telescope (J-band, Fig. 1b). We combined our midtransit timing with the low precision observat ions reported by Burke et al. (2008), the better precision observat ions o f Pál et al. (2009), and two addit ional observat ions from Maciejewski et al. (2011c) to der ive a per iod of 4.1877544±0.0000016 reported values. fall within previously system parameters also days. The TABLE 1 EXTRASOLAR PLANET TRANSIT INFORMATION Name Date Tel. Filter # C.S. Base. Fit Mid-Transit Time Duration Depth (1) (2) (3) (4) (5) (BJD_TDB) (min) (%) Nov. 27, 2009 CoRoT-1 May 09, 2009 CoRoT-2 May 29, 2010 GJ 1214 May 29, 2010 May 18, 2011 May 18, 2011 Oct. 08, 2007 HAT-P-1 Oct. 17, 2007 May 15, 2009 HAT-P-3 May 15, 2009 May 27, 2010 May 27, 2010 May 22, 2011 HAT-P-4 Nov. 25, 2009 HAT-P-6 Jun. 01, 2010 HAT-P-11 Jun. 01, 2010 May 14, 2011 HAT-P-12 May 31, 2010 HAT-P-27/W40 May 21, 2011 Oct. 09, 2011 HAT-P-32 5162.91698 ± 0.00059 4960.88065 ± 0.00044 5345.82204 ± 0.00011 5345.82221 ± 0.00032 5699.83288 ± 0.00034 5699.83283 ± 0.00014 4381.81060 ± 0.00077 4390.74563 ± 0.00139(6) 4966.89186 ± 0.00025 4966.89323 ± 0.00047 5343.85846 ± 0.00021 5343.85867 ± 0.00048 5703.77929 ± 0.00060 5160.75292 ± 0.00034 5348.83923 ± 0.00027(6) 5348.83574 ± 0.00055 5695.87244 ± 0.00105 5347.76929 ± 0.00021 5702.74876 ± 0.00039 5843.75341 ± 0.00019 147.2 ± 2.4 140.4 ± 1.8 52.7 ± 0.5 55.1 ± 1.3 54.6 ± 1.4 52.7 ± 0.6 174.8 ± 3.1 181.5 ± 5.7 123.6 ± 1.0 115.4 ± 1.9 119.5 ± 0.9 129.1 ± 2.0 244.2 ± 2.5 204.4 ± 1.4 146.4 ± 1.1 145.6 ± 2.3 158.7 ± 4.3 138.3 ± 0.9 97.5 ± 1.6 187.1 ± 0.8 1.96 ± 0.65 3.23 ± 0.51 1.47 ± 0.16 1.74 ± 0.39 1.42 ± 0.37 1.44 ± 0.16 1.68 ± 0.40 1.47 ± 0.66 1.27 ± 0.20 1.36 ± 0.28 1.25 ± 0.23 1.35 ± 0.28 0.73 ± 0.21 0.96 ± 0.16 0.45 ± 0.11 0.50 ± 0.23 0.50 ± 0.21 2.27 ± 0.28 1.28 ± 0.15 2.59 ± 0.20 2.1m VC 2.1m VC VC 2.1m MP MP 2.1m VC 2.1m VC 2.1m 2.1m 2.1m VC VC 2.1m 2.1m 2.1m JH z’ J B z’ J H H JH z’ JH B J J J B B J J J 3 8 7 8 4 8 1 1 3 4 3 6 2 8 6 2 5 5 7 7 P5 X P5 X --- P1 P1 P1 P3 X P5 X P1(7) P5 P2 --- X P5 P2 P1 Oct. 11, 2011 Oct. 11, 2011 Nov. 24, 2009 HD 17156 Oct. 19, 2008 HD 189733 May 13, 2011 Oct. 16, 2011 Qatar-1 Oct. 10, 2007 TrES-1 Oct. 13, 2007 May 20, 2011 TrES-2 Oct. 13, 2011 May 06, 2009 TrES-3 May 30, 2010 TrES-4 May 30, 2010 Oct. 22, 2008 WASP-1 Oct. 18, 2008 WASP-2 May 12, 2009 WASP-3 Jun. 02, 2010 Jun. 02, 2010 May 17, 2011 Oct. 12, 2011 WASP-6 Oct. 17, 2008 WASP-10 Oct. 17, 2008 WASP11/HP10 Nov. 26, 2009 Nov. 07, 2010 Oct. 08, 2011 Oct. 08, 2011 2.1m VC VC 2.1m VC 2.1m MP MP VC 2.1m VC 2.1m VC 2.1m 2.1m VC 2.1m VC VC 2.1m VC 2.1m 2.1m VC 2.1m VC J z’ z’ J Ha J H H z’ J z’ J B J J z’ J z’ z’ J z’ J J z’ J z’ 6 5 4 5 2 4 1 1 5 8 7 5 4 8 7 8 5 4 6 7 6 5 7 5 6 5 P1 X --- P2 X P3 X --- X --- X P3 X P2 P3 X P2 X P1 P2 X P2 P3 --- --- P1 5845.90287 ± 0.00024 5845.90314 ± 0.00040 5159.84014 ± 0.00068 4758.64910 ± 0.00011 5694.88813 ± 0.00025 5850.69628 ± 0.00019 4383.68588 ± 0.00072 4386.71733 ± 0.00078 5701.88715 ± 0.00068 5847.65447 ± 0.00029 4957.86698 ± 0.00048 5346.83509 ± 0.00061(6) 5346.84102 ± 0.00051 4761.73558 ± 0.00033 4757.70492 ± 0.00032 4963.84563 ± 0.00055 5349.83457 ± 0.00039 5349.83182 ± 0.00039 5698.88358 ± 0.00060 5846.72540 ± 0.00045 4756.82125 ± 0.00039 4756.81997 ± 0.00018 5161.71529 ± 0.00021 5507.90419 ± 0.00042 5842.92921 ± 0.00021 5842.92952 ± 0.00044 185.5 ± 1.0 183.2 ± 1.6 182.3 ± 2.7 109.8 ± 0.4 108.5 ± 1.0 101.3 ± 0.8 150.3 ± 2.9 147.5 ± 3.2 110.4 ± 2.8 110.2 ± 1.2 77.9 ± 1.9 217.1 ± 2.5 215.8 ± 2.1 226.0 ± 1.4 109.8 ± 1.3 160.9 ± 2.3 161.6 ± 1.6 156.8 ± 1.6 150.4 ± 2.4 157.8 ± 1.8 132.5 ± 1.6 134.2 ± 0.7 158.6 ± 0.9 163.1 ± 1.7 154.3 ± 0.8 158.1 ± 1.8 2.21 ± 0.22 2.39 ± 0.21 0.72 ± 0.30 2.49 ± 0.16 2.68 ± 0.35 2.46 ± 0.19 1.85 ± 0.47 1.58 ± 0.53 1.43 ± 0.28 1.63 ± 0.19 2.49 ± 0.48 0.91 ± 0.17 0.89 ± 0.18 1.24 ± 0.23 1.41 ± 0.17 1.09 ± 0.26 1.01 ± 0.14 1.15 ± 0.21 1.09 ± 0.29 2.18 ± 0.24 3.05 ± 0.26 2.92 ± 0.28 1.78 ± 0.19 2.09 ± 0.35 1.87 ± 0.25 1.94 ± 0.24 Oct. 29, 2010 May 20, 2011 Oct. 15, 2011 Nov. 03, 2010 Oct. 13, 2011 Oct. 13, 2011 May 15, 2011 Oct. 15, 2011 Oct. 17, 2011 May 08, 2009 Nov. 24, 2009 VC 2.1m 2.1m VC 2.1m VC 2.1m VC 2.1m VC 2.1m z’ J J z’ J Ha J z’ J z’ J 8 3 3 6 4 2 8 2 2 2 7 X P2 P2 P1 --- --- --- X P2 --- P5 5498.89590 ± 0.00079 5701.80338 ± 0.00049 5849.75000 ± 0.00037 5503.86346 ± 0.00035 5847.86796 ± 0.00032 5847.86974 ± 0.00072 5696.81358 ± 0.00057 5849.92131 ± 0.00060 5851.87634 ± 0.00028 4959.83598 ± 0.00039 5159.89907 ± 0.00038 177.2 ± 3.2 152.6 ± 2.0 139.5 ± 1.5 167.5 ± 1.4 169.9 ± 1.3 155.3 ± 2.9 190.3 ± 2.3 108.6 ± 2.5 108.8 ± 1.2 177.6 ± 1.6 188.3 ± 1.5 1.39 ± 0.45 1.12 ± 0.31 1.16 ± 0.23 1.14 ± 0.20 0.95 ± 0.14 1.16 ± 0.32 1.07 ± 0.19 2.11 ± 0.32 1.84 ± 0.19 2.11 ± 0.38 1.19 ± 0.27 (1) Telescope Used: 2.1m = KPNO 2.1m telescope, VC = KPNO V isitor Center 0.5m telescope, MP = KPNO National So lar Observatory McMath-P ierce 2.0m Telescope. (2) Filter Used: B = Johnson Blue (0.44 μm), Ha = Hydrogen Alpha (0.656 μm), z’ = S loan DSS z’ (0.90 μm), J = mid-IR J-band (1.25 μm), H = mid-IR H-band (1.64 μm), JH = mid-IR J- and H-bands combined (1.25 and 1.64 μm) (3) # C.S.: Number of field comparison stars used to derive the lightcurve. (4) Base. Fit: Type of fit used on the out-of-transit baseline. Pn = polynomial o f order n, X = airmass dependence, --- = none used. (5) Barycentric Julian Date based on Dynamical T ime (2,450,000+). To convert to BJD_UTC (Coordinated Universal T ime) subtract 0.00075 days from the 2008 times, and 0.00077 days from the 2009 and 2010 times. (6) Timing issue with telescope. Not trustworthy. (7) Severe d iscont inuity in data after egress. Fixed manually. WASP-12 WASP-24 WASP-32 WASP-33 WASP-48 WASP-50 XO-1 XO-5 Name Number of Time Coverage Period Observat ions (years) (days) CoRoT-1 CoRoT-2 GJ 1214 HAT-P-1 HAT-P-3 HAT-P-4 HAT-P-6 HAT-P-11 HAT-P-12 HAT-P-27/WASP-40 HAT-P-32 HD 17156 HD 189733 Qatar-1 TrES-1 TrES-2 TrES-3 TrES-4 WASP-1 WASP-2 WASP-3 WASP-6 WASP-10 1.5089682 ± 0.0000005 1.7429981 ± 0.0000011 1.5804048 ± 0.0000002 4.4653054 ± 0.0000069 2.8997382 ± 0.0000009 3.0565254 ± 0.0000012 3.8530018 ± 0.0000015 4.8878056 ± 0.0000015 3.2130553 ± 0.0000010 3.0395824 ± 0.0000035 2.1500103 ± 0.0000003 21.216384 ± 0.000016 2.2185754 ± 0.0000001 1.4200227 ± 0.0000012 3.0300724 ± 0.0000004 2.4706128 ± 0.0000003 1.3061865 ± 0.0000002 3.5539303 ± 0.0000019 2.5199425 ± 0.0000014 2.1522213 ± 0.0000004 1.8468332 ± 0.0000004 3.3609998 ± 0.0000011 3.0927297 ± 0.0000003 3.0 3.8 2.0 1.1 3.1 4.0 1.9 6.8 2.5 1.4 3.9 2.2 5.7 0.9 5.0 5.2 3.2 3.1 2.3 3.2 4.3 3.4 3.0 38 5 33 11 19 14 4 32 2 3 4 10 59 2 31 54 28 9 6 7 33 3 22 TABLE 2 DERIVED SYSTEM EPHEMERIDES JD0 (BJD_TDB) (2,450,000+) 4138.32807 ± 0.00006 4237.53639 ± 0.00014 4964.94469 ± 0.00006 3984.39735 ± 0.00026 4218.75959 ± 0.00026 4245.81521 ± 0.00020 4035.67618 ± 0.00025 4605.89123 ± 0.00013 4419.19631 ± 0.00020 5186.01982 ± 0.00032 4420.44637 ± 0.00009 4353.61930 ± 0.00034 3629.39489 ± 0.00003 5518.41097 ± 0.00020 3186.80703 ± 0.00012 3957.63574 ± 0.00011 4185.91110 ± 0.00008 4230.90575 ± 0.00043 3912.51531 ± 0.00032 3991.51536 ± 0.00018 4143.85194 ± 0.00017 4596.43342 ± 0.00013 4664.03803 ± 0.00006 WASP-11/HAT-P-10 WASP-12 WASP-24 WASP-32 WASP-33 WASP-48 WASP-50 XO-1 XO-5 8 6 10 2 5 2 3 25 22 6.0 2.7 2.1 1.9 4.6 0.9 0.8 5.0 2.8 3.7224793 ± 0.0000007 1.0914224 ± 0.0000003 2.3412162 ± 0.0000014 2.7186591 ± 0.0000024 1.2198721 ± 0.0000003 2.1436283 ± 0.0000041 1.9550905 ± 0.0000022 3.9415052 ± 0.0000008 4.1877545 ± 0.0000016 4759.68753 ± 0.00011 4508.97683 ± 0.00019 5081.38033 ± 0.00010 5151.05460 ± 0.00050 4163.22465 ± 0.00022 5364.55120 ± 0.00027 5558.61277 ± 0.00020 3808.91777 ± 0.00011 4485.66876 ± 0.00028 TABLE 3 MODELED SYSTEM PARAMETERS Name Limb Darkening Coeff. a b i [o] 85.72+0 .58 86.01+2 .74 83.89+0 .84 88.46+0 .99 84.23+0 .88 88.16+1 .17 86.05+0 .43 84.81+0 .82 84.08+0 .61 82.36+1 .77 85.90+2 .62 85.18+1 .26 86.33+2 .51 88.60+0 .91 88.64+0 .91 89.24+0 .52 89.08+0 .64 85.00+1 .58 83.24+3 .71 86.42+2 .37 85.06+3 .05 85.31+0 .82 87.24+1 .41 -0.51 -3.24 -0.74 -0.93 -0.88 -1.03 -0.37 -0.74 -0.67 -1.16 -2.37 -1.03 -2.73 -0.88 -0.91 -0.69 -0.86 -1.29 -2.14 -2.83 -3.13 -0.68 -1.14 a / R* Rp / R* 9.21+0 .56 5.64+0 .42 6.77+0 .51 11.22+0 .45 9.11+1 .01 5.98+0 .10 9.12+0 .38 6.56+0 .36 7.93+0 .57 5.79+0 .84 5.24+0 .32 8.22+1 .08 5.48+0 .33 10.44+0 .27 11.81+0 .44 12.11+0 .18 12.28+0 .28 7.71+1 .08 3.31+0 .20 7.08+0 .66 5.44+0 .57 7.90+0 .56 10.12+1 .01 -0.48 -0.74 -0.41 -0.69 -0.71 -0.15 -0.32 -0.32 -0.54 -0.47 -0.46 -0.82 -0.55 -0.46 -0.72 -0.38 -0.60 -0.82 -0.16 -1.12 -0.76 -0.46 -1.01 0.1093+0 .0020 0.0804+0 .0049 0.0970+0 .0023 0.1404+0 .0026 0.1344+0 .0389 0.1531+0 .0012 0.1536+0 .0025 0.1499+0 .0023 0.1295+0 .0048 0.0942+0 .0030 0.1069+0 .0034 0.1135+0 .0053 0.0951+0 .0043 0.1395+0 .0014 0.1598+0 .0040 0.1255+0 .0020 0.1256+0 .0027 0.1030+0 .0031 0.1022+0 .0027 0.0998+0 .0039 0.0988+0 .0051 0.1347+0 .0037 0.1022+0 .0033 -0.0019 -0.0051 -0.0023 -0.0026 -0.0174 -0.0012 -0.0026 -0.0030 -0.0039 -0.0035 -0.0033 -0.0056 -0.0045 -0.0013 -0.0038 -0.0018 -0.0028 -0.0033 -0.0028 -0.0040 -0.0049 -0.0037 -0.0031 0.095 0.398 HAT-P-3 0.102 0.366 HAT-P-4 0.047 0.360 HAT-P-6 0.221 0.313 HAT-P-12 0.170 0.341 HAT-P-27/WASP-40 HAT-P-32 (1) 0.064 0.366 0.187 0.332 HD 189733 0.195 0.340 Qatar-1 0.087 0.362 TrES-2 0.065 0.372 TrES-4 0.076 0.374 WASP-1 0.170 0.341 WASP-2 0.055 0.366 WASP-3 0.119 0.359 WASP-6 0.231 0.311 WASP-10 WASP-11/HAT-P-10 (2) 0.222 0.320 WASP-11/HAT-P-10 (3) 0.222 0.320 0.071 0.362 WASP-32 WASP-33 (4) 0.016 0.359 0.076 0.365 WASP-24 0.252 0.301 WASP-48 0.128 0.358 WASP-50 XO-5 0.137 0.356 (1) Parameters for the 2011 October 09 lightcurve. (2) Parameters for the 2009 November 26 J-band lightcurve which exhibits starspot activity. (3) Parameters for the 2011 October 08 J-band lightcurve. (4) Exhibits stellar variability and uneven baseline outside the transit. Figure 1a: Exoplanet Transits observed in the near-IR with the KPNO 2.1m telescope Figure 1b: Exoplanet Transits observed in the near-IR with the KPNO 2.1m telescope Figure 2a: Exoplanet Transits observed in the near-IR with the KPNO Nat ional So lar Observatory McMath-P ierce 2.0m telescope ( left column) and in the visible with the KPNO V isitor Center 0.5m telescope. Figure 2b: Exoplanet Transits observed in the visible with the KPNO V isitor Center 0.5m telescope. Figure 3: Ratio of planetary to stellar radius, Rp/R*, from MCMC fitt ing to our J-band transit curves (ordinate, from Table 3), compared to published optical radii for the same planets. We exclude HAT-P-27/WASP-40 due to its much larger error. The solid line is a least squares fit (see text), and it differs negligibly from the null hypothesis that the derived planetary radii are independent of wavelength (see text). Figure 4: Observed transit of HAT-P-10/WASP-11 on 2009 November 26 at the KPNO 2.1m telescope through a J-Band filter. The solid line is the model der ived from the literature parameters. The dotted line is our fit to the entire depth of the transit. The dashed line is the literature model but with part of the stellar surface covered by starspots with temperature ~6% lower than the effect ive temperature of the star (4920 K). REFERENCES Ago l, E., et al. 2010, ApJ, 721, 1861 Albrecht, S. et al. 2011, ApJ accpeted (arXiv; 1106.2548A) Alonso, R. et al. 2004, ApJ, 613, L153 Alonso, R. et al. 2008, A&A, 482, L21 Alsubai, K. et al. 2010, MNRAS 417, 709 Anderson, D. R. et al. 2011, PASP, 123, 555 Ayres, T., Penn, M., Plymate, C. & Keller, C. 2008, ESPM-12, 2.74 Bakos, G. Á. et al. 2006, ApJ, 650, 1160 Bakos, G. Á. et al. 2007, ApJ, 670, 826 Bakos, G. Á. et al. 2010, ApJ, 696, 1950 Bakos, G. Á. et al. 2010b, ApJ, 710, 1724 Barbieri, M. et al. 2007, A&A 476, L13 Barge, P. et al. 2008, A&A, 482, L17 Bean, J. L. 2009, A&A 506, 369. Béky, B. et al. 2011, ApJ 734, 109 Berta, Z. A. et al. 2011, ApJ, 736, id. 12 Bradstreet, D. H. 2005, SASS-24, 23 Bouchy, F., et al. 2005, A&A, 444, L15 Burke, Ch. J. et al. 2008, ApJ, 686, 1331 Burke, Ch. J. et al. 2010, ApJ, 719, 1796 Cáceres, C. et al. 2009, A&A, 507, 481 Carter, J. A. et al. 2011, ApJ, 730, 82 Chan, T. et al. 2011, AJ, 141,179 Charbonneau, D. et al. 2005, ApJ, 626, 523 Charbonneau. D. et al. 2007, ApJ, 658, 1322 Charbonneau, D. et al. 2009, ApJ, 626, 523 Charbonneau, D. et al. 2009, Nature, 462, 891 Christ ian, D. J. et al. 2009, MNRAS, 392, 1585 Christ iansen, J. L. et al. 2011, ApJ, 726, 94 Claret, A. 2000, A&A 363, 1081 Collier Cameron, A. et al. 2007, MNRAS, 375, 951 Collier Cameron, A. et al. 2010, MNRAS, 407, 507 Colón, K. D., Ford, E. B., Lee, B., Mahadevan, S. & Blake, C. H. 2010, MNRAS, 408, 1494 Croll, B. et al. 2011, ApJ, 736, 78 Csizmadia, Sz. et al. 2010, A&A 510, A94 Deming, D. et al. 2005, Nature, 434, 740 Deming, D. et al. 2007, ApJ, 667, L199 Deming, D. et al. 2011, ApJ, 740, 33 Dittmann, J. A. et al. 2009, ApJ, 699, L48 Dittmann, J. A., Close, L. M., Scuderi, L. J., & Morris, M. D. 2010, ApJ, 717, 235 Dragomir, D., et al. 2011, AJ, 142, 115 Eastman, J., Siverd, R. & Gaudi, B. S. 2010, PASP, 122, 935 Elston, R. 1998, Proc. SPIE, 3354, 404 Enoch, B. et al. 2011, AJ, 142, id. 86 Ford, E. B. 2005, AJ, 129, 1706 Gazak, J. Z., Tonry, J. & Johnson, J.A. 2011, Earth and Planetary Astrophysics, preprint (arXiv; 1102.1036) Gibson, N. P. et al. 2009, ApJ, 700, 1078 Gibson, N. P. et al. 2010, MNRAS, 401, 1917 Gillon, M. et al. 2008, A&A, 485, 871 Gillon, M. et al. 2009a, A&A, 501, 785 Gillon, M. et al. 2009b, A&A, 506, 359 Gillon, M. et al. 2011, A&A, 533, A88 Hartman, J. et al. 2009, ApJ, 706, 785 Hartman, J. et al. 2011, ApJ, 742, 59 Hebb, L. et al. 2009, ApJ, 693, 1920 Herrero, E., Morales, J. C., Ribas, I., & Naves, R. 2011, A&A, 526, L10 Hirano, T. et al. 2011, PASJ, 63, S531 Holman, M. J. et al. 2006, ApJ, 652, 1715 Holman, M. J. et al. 2007, ApJ, 664, 1185 Hrudková, M. et al. 2009, IAU Symposium, 253, 446 Hrudková, M., et al. 2010, MNRAS, 403, 2111 Irwin, J. et al. 2008, ApJ 681, 636 Johnson, J. A. et al. 2008, ApJ, 686, 649 Johnson, J. A., Winn, J. N., Cabrera, N. E. & Carter, J. A. 2009, ApJ, 692, L100 Johnson, J. A., Winn, J. N., Cabrera, N. E. & Carter, J. A. 2010, ApJ, 712, L122 Kipping, D. & Bakos, G. 2011, ApJ, 733, 36 Knutson, H. A., et al. 2007, Nature, 447, 183 Knutson, H. A., et al. 2009, ApJ, 690, 822 Knutson, H. A., et al. 2010, ApJ, 720, 1569 Kovács, G. et al. 2007, ApJ, 670, L41 Krejčová, T., Budaj, J. & Krushevska, V. 2010, CoSka, 40, 77 Kundurthy, P. et al. 2011, ApJ, 731, id. 123 Lanza, A. F. et al. 2009, A&A, 493, 193 Littlefield, C. 2011, Scient ia, 2, (arXiv: 1106:4312) Maciejewski, G. et al. 2011a, MNRAS, 411, 1204 Maciejewski, G. et al. 2011b, MNRAS, 528, A65 Maciejewski, G. et al. 2011c, A&A, 535, A7 Maciejewski, G., Seeliger, M., Adam, Ch., Raetz, St. & Neuhäuser, R. 2011c, ACA, 61, 25 Mandushev, G. et al. 2007, ApJ, 667, L195 Maxted, P.F.L. et al. 2010, PASP 122, 1465 McCullough, P. R. et al. 2006, ApJ, 648, 1228 Miller-Ricci, E., et al. 2008, ApJ, 682, 593 Mislis, D. & Schmitt, J. H. M. M. 2009, A&A, 500, 45 Mislis, D., Schroter, S., Schmitt, J. H. M. M., Cordes, O. & Reif, K. 2010, A&A, 510, 107 Narita, N. et al., 2007, PASJ, 59, 763 Narita, N., Sato, B., Ohshima, O., & Winn, J. N. 2008, PASJ, 60, L1 Nascimbeni, V., Piotto, G., Bedin, L. R. & Damasso, M. 2011, A&A, 527, A85 Noyes, R. W. et al. 2008, ApJ, 673, L79 Nutzman, P. et al. 2011, ApJ, 726, id. 3 O’Donovan, F. T. et al. 2006, ApJ, 651, L61 O’Donovan, F. T. et al. 2007, ApJ, 663, L37 Pál, A. et al. 2009, ApJ, 700, 783 Poddany, S., Brat, L. & Pejcha, O. 2010, New Astronomy, 15, 297. (http://var2.astro.cz/ETD/index.php) Pont, F., et al. 2007, A&A, 476, 1347 Rabus, M. et al. 2009, A&A, 508, 1011 Rabus, M., Deeg, H. J., Alonso, R., Belmonte, J. A. & Almenara, J. M. 2009, A&A, 508, 1011 Raetz, St. et al. 2009a, ASNA, 330, 459 Raetz, St. et al. 2009b, ASNA, 330, 475 Rauer, H. et al. 2010, AJ, 139, 53 Sada, P. V. et al. 2010, ApJL, 720, L215 Sanchis-Ojeda, R. et al. 2011, ApJ, 743, id. 61 Scuderi, L. J., Dittmann, J. A., Males, J. R., Green, E. M. & Close, L. M. 2010, ApJ, 714, 462 Seager. S. and Deming, D. 2010, ARAA, 48, 631 Shporer, A., Tamuz, O., Zucker, S., & Mazeh, T. 2007, MNRAS, 376, 1296 Sing, D. K. et al. 2011, MNRAS, 416, 1443 Smith, A. M. S., Anderson, D. R., Skillen, I., Collier Cameron, A., & Smalley, B. 2011, MNRAS, 416, 2096 Simpson, E. K. et al. 2011, MNRAS, 414, 3023 Southworth, J. 2008, MNRAS, 386, 1644 Southworth, J. et al. 2009, MNRAS, 399, 287 Southworth, J. et al. 2010, MNRAS, 408, 1680 Sozzetti, A. et al. 2009, ApJ, 691, 1145 Street, R. A. et al. 2010, ApJ, 720, 337 Szabó, Gy. M., Haja, O., Szatmáry, K., Pál, A. & Kiss, L. L. 2010, IBVS, 5919, 1 Todorov, K. O. et al. 2011, in press for ApJ, (arXiv: 1111:5858) Torres, G. et al. 2007, ApJ, 666, L121 Torres, G., Winn, J. N. & Holman, M. J, 2008, ApJ, 677, 1324. Vaňko, M. et al. 2009, IAU Symposium 253, 440 Vaňko, M. et al. 2011, IAU Symposium 282, (arXiv: 1108:5255) Vereš, P., Budaj, J., Világum J., Galád, A. & Kornoš, L. 2009, Contrib. Astron. Obs. Skalnaté Pleso, 39, 34 Wang, X.-B., Collier Cameron, A., Gu, Sh.-H., & Zhang, L.-Y. 2008, Proceedings IAU Symposium, 249, 85 West, R. G. et al. 2009, A&A 502, 395 Wilson, D. M. et al. 2006, PASP, 118, 1245 Winn, J. N. et al. 2007a, AJ, 133, 1828 Winn, J. N. et al. 2007b, AJ, 134, 1707 Winn, J. N. et al. 2007c, ApJ, 657, 1098 Winn, J. N. et al. 2008, AJ, 136, 1753 Winn, J. N. et al. 2009, ApJ, 693, 794 Winn, J. N. et al. 2010, AJ, 141, 63 Woo-Lee, J., Youn, J.-H., Kim, S.-L., Lee, Ch.-U. & Koo, J.R. 2011, PASJ, 63, 301
1204.5187
1
1204
2012-04-23T20:00:05
Interactions Between Moderate- and Long-Period Giant Planets: Scattering Experiments for Systems in Isolation and with Stellar Flybys
[ "astro-ph.EP" ]
The chance that a planetary system will interact with another member of its host star's nascent cluster would be greatly increased if gas giant planets form in situ on wide orbits. In this paper, we explore the outcomes of planet-planet scattering for a distribution of multiplanet systems that all have one of the planets on an initial orbit of 100 AU. The scattering experiments are run with and without stellar flybys. We convolve the outcomes with distributions for protoplanetary disk and stellar cluster sizes to generalize the results where possible. We find that the frequencies of large mutual inclinations and high eccentricities are sensitive to the number of planets in a system, but not strongly to stellar flybys. However, flybys do play a role in changing the low and moderate portions of the mutual inclination distributions, and erase dynamically cold initial conditions on average. Wide-orbit planets can be mixed throughout the planetary system, and in some cases, can potentially become hot Jupiters, which we demonstrate using scattering experiments that include a tidal damping model. If planets form on wide orbits in situ, then there will be discernible differences in the proper motion distributions of a sample of wide-orbit planets compared with a pure scattering formation mechanism. Stellar flybys can enhance the frequency of ejections in planetary systems, but auto-ionization is likely to remain the dominant source of free-floating planets.
astro-ph.EP
astro-ph
Interactions Between Moderate- and Long-Period Giant Planets: Scattering Experiments for Systems in Isolation and with Stellar Flybys Aaron C. Boley, Matthew J. Payne, and Eric B. Ford Department of Astronomy, University of Florida, 211 Bryant Space Science Center, PO Box 112055, Gainesville, FL 32611 ABSTRACT The chance that a planetary system will interact with another member of its host star's nascent cluster would be greatly increased if gas giant planets form in situ on wide orbits. In this paper, we explore the outcomes of planet-planet scat- tering for a distribution of multiplanet systems that all have one of the planets on an initial orbit of 100 AU. The scattering experiments are run with and without stellar flybys. We convolve the outcomes with distributions for protoplanetary disk and stellar cluster sizes to generalize the results where possible. We find that the frequencies of large mutual inclinations and high eccentricities are sen- sitive to the number of planets in a system, but not strongly to stellar flybys. However, flybys do play a role in changing the low and moderate portions of the mutual inclination distributions, and erase dynamically cold initial conditions on average. Wide-orbit planets can be mixed throughout the planetary system, and in some cases, can potentially become hot Jupiters, which we demonstrate using scattering experiments that include a tidal damping model. If planets form on wide orbits in situ, then there will be discernible differences in the proper motion distributions of a sample of wide-orbit planets compared with a pure scattering formation mechanism. Stellar flybys can enhance the frequency of ejections in planetary systems, but auto-ionization is likely to remain the dominant source of free-floating planets. 1. Introduction Observations have revealed a rich distribution of planetary system architectures1. Mas- sive Jovian planets can be found on orbits with periods ranging from a few days to thousands 1exoplanets.org (Wright et al. 2011) -- 2 -- of years (e.g., Marcy et al. 1997; Kalas et al. 2008; Marois et al. 2008; Lafreni`ere et al. 2010). Super-Earths and Neptune-size planets are abundant within stellar separations of 0.5 AU (Borucki et al. 2011). Multi-planet systems are common, including densely-packed orbital configurations (e.g., Kepler-11, Lissauer et al. 2011a). This diversity demonstrates that plan- ets cannot be thought of as isolated objects slowly growing within their respective feeding zones. Even in the Solar System, the Late Heavy Bombardment, Kuiper Belt orbital struc- ture, asteroid belt composition, and the mass of Mars suggest that the Solar System planets experienced substantial migration (e.g., Walsh et al. 2011). The architectures of large bodies in planetary systems are sculpted by at least two general mechanisms: n-body dynamics and disk-planet interactions. Neither is mutually exclusive. Examples of the former include planet-planet scattering (e.g., Rasio & Ford 1996; Lin & Ida 1997), interactions with a stellar companion (e.g. Holman et al. 1997), and planet- planet-stellar perturber excitation (Adams & Laughlin 2001; Zakamska & Tremaine 2004; Malmberg et al. 2011). The resulting scattering could explain the planet eccentricity dis- tribution, for which the median eccentricity e ≈ 0.14 (Wright et al. 2011), and could even explain highly inclined and in some cases retrograde systems (Chatterjee et al. 2008; Naga- sawa et al. 2008; Triaud et al. 2010). The second mechanism, disk-planet interactions, can cause planets to move throughout the nebula (Kley & Nelson 2012). Detailed planetary type I migration studies that include proper thermodynamics (Paardekooper & Mellema 2006) and radiation hydrodynamics (Kley et al. 2009) show that migration can be inward or outward for a range of conditions, with zero-torque radii possible as well. If two massive planets open a mutual gap in the disk, then their migration can also be inward or outward, depending on the details of a given disk's structure and the planet mass ratios (Snellgrove et al. 2001; Crida et al. 2009). Disk- planet interactions typically lead to eccentricity and inclination damping (Bitsch & Kley 2010, 2011) for the majority of planet masses and disk conditions (Moorhead & Adams 2008). Eccentricity excitation may also be possible for large planet masses or specific disk conditions (e.g., Goldreich & Tremaine 1980; Ogilvie & Lubow 2003). While excited systems may be best explained by n-body interactions, densely-packed systems like Kepler-11 or systems in or near resonances (e.g., Lissauer et al. 2011b) likely require a phase of planet- disk interactions. These mechanisms can, either separately or in combination (Moorhead & Adams 2005), turn a planetary system with planets on moderate-period orbits, e.g., between ∼ 1 to 10 AU, into a system with planets on short- and long-period orbits. Nonetheless, it remains to be seen whether scattering and/or migration can match the constraints set by multi-planet systems (e.g., Veras et al. 2009; Dodson-Robinson et al. 2009). It is also possible that formation at -- 3 -- moderate periods is not the only mode of planet formation, with at least some wide-orbit planets forming in situ by disk instability during the earliest stages of disk evolution (Boss 1997; Boley 2009). If a massive planet can form at large stellar separations, regardless of the mechanism, then the cross section for significant perturbation of the planetary system by stellar flybys would be much larger than for the solar system, and stellar flybys may be more important in shaping planetary orbits. For example, a distant stellar flyby could cause otherwise stable systems to grow to instability due to a cascade of eccentricity pumping (Zakamska & Tremaine 2004) or to decrease the decay timescale of the system (Malmberg et al. 2011). Wide-orbit planets that are placed on highly inclined orbits could also induce Kozai oscillations with other system members (e.g., Naoz et al. 2011). Finally, just as planets that form at short periods may have scattered onto on wide-orbits, e.g., Veras et al. (2009), planets that form on wide-orbits may be placed on short periods through multiple scattering events. In this paper, we explore outcomes for planet-planet scattering under the assumption that planetary architectures can begin with planets on wide-orbits. We compare isolated systems with systems that experience stellar flybys. In Section 2, we discuss encounter likelihoods, and using rates from the literature, estimate the fraction of field stars that have had an encounter with a pericenter less than some value q. We then describe our base set of scattering experiments in Section 3. We present the results in Section 4 and in Section 5, use those results to determine expectation values for median inclination and eccentricities among distributions of field star planetary systems. We also demonstrate that proper motion distributions can be used to discriminate between formation modes, and show that planets on initial 100 AU orbits can become hot Jupiter candidates. We conclude with a summary of the results in Section 6. A summary of the symbols used in this manuscript is given in Table 1. 2. Encounter Frequency Proszkow & Adams (2009), hereafter PA2009, characterized the encounter rates for stars in cluster sizes between N = 100 and 3000 for a wide range of parameters (see also Adams et al. 2006). In their study, they focused on both virial (Q = 0.5) and subvirial (Q = 0.04) velocity dispersions, where Q = Total Kinetic Energy/Total Potential Energy. For some clusters, they explored the sensitivity of the interaction rate to the star cluster core radius rc by varying Q. Here, we use the results from their Q = 0.04 initial conditions (ICs) with a cluster core radius scaling rc = 1 pc (N/300)1/2, which we choose for three principal reasons. (1) The velocity dispersion among prestellar cores is observed to be small (e.g., Andr´e 2002), -- 4 -- subvirial ICs give an effective core radius that is ∼ √ suggesting that star clusters are out of virial equilibrium at birth. (2) Star cluster cores during their gas-embedded phase are initially compact (Bastian et al. 2008), and expand to the sizes found by Lada & Lada (2003) as they evolve, with ambient gas removal likely playing a role in the cluster's expansion (Bastian & Goodwin 2006). PA2009 found that their 2 smaller than the initial rc, which is more inline with the Bastian et al. results. (3) We are specifically interested in clusters that have short lifetimes and are the dominate contributors to the field population, which are the targets for most planet discovery surveys. This limits cluster sizes to be (cid:46) 104. Within this parameter space, star cluster core radii follow the Lada & Lada scaling rc ∝ N 1/2. To proceed, we first make a simple estimate as to whether close encounters could be important for producing highly inclined outer planets in the field star population. Let Γ(q, N ) be the time-averaged rate for all encounters with a stellar flyby pericenter ≤ q in a nascent cluster of size N . We roughly model the PA2009 results (their Table 8) using the following: (cid:18)100 (cid:19)1/2(cid:16) (cid:17)γ(N ) q Γ(q, N ) ≈ 0.26 N 1000 AU encounters per star per Myr. (1) We determined the functional form for γ using the tabulated results of PA2009, and the value of γ represents the typical degree of gravitational focusing. As γ → 1, gravitational focusing becomes strong, and when γ → 2, focusing becomes weak. We set the rate exponent to γ(N ) = 2− exp (−N/782), forcing the value of γ to be between 1 and 2. Let ∆t represent the time period in a cluster during which close encounters remain important, which gives us the number of encounters with a closest approach distance < q per star for a given cluster size N as Γ(q, N )∆t. Because Γ is averaged over 10 Myr in PA2009, we will typically take ∆t ∼ 10 Myr unless noted otherwise. Next, we assume that all field stars come from dissolved clusters with member num- In this case, we use the canonical star cluster mass function bers between N0 and N1. (mdξm/dm ∼ m−1) to write the star cluster number function; namely, dξN /dN = AN−1, (2) where A is set to normalize the function to unity. Finally, we write the average number of encounters per field star for flyby pericenter < q as (cid:90) N1 N0 η = dξN dN Γ (q, N ) ∆t dN. (3) With this definition for η = η(q, N0, N1, ∆t, dξN dN ), extra weight will be given to stars that have multiple encounters for pericenters < q. We account for this weighting by introducing -- 5 -- η(cid:48), which has the same form as η, but forces Γ∆t ≤ 1. The value of η(cid:48) thus represents the fraction of field stars that have had at least one encounter. The average number of encounters among field stars that have had at least one encounter is given by the ratio of η to η(cid:48). In Table 2, we give the results of integrating equation (3) over several values of N0 and N1 for q = 100, 200, 300 AU, and 1000 AU, with ∆t = 10 Myr. The results are fairly sensitive to N0, owing to the increased likelihood of a star to have a close encounter in small N clusters, but we do find that 20-40% of field stars should have experienced at least one encounter within 300 AU. Next, we discuss the effects of the core cluster size. 2.1. Sensitivity of Results to Assumptions for Nascent Cluster Core Sizes The dominant source of uncertainty in the results for the following calculations is the nascent cluster stellar density during which most collisions occur. To understand this sen- sitivity, we consider a general flyby rate Γ = nvσ, where σ is the cross section for the for a star to pass within a distance q of another star, n is the typical stellar density in the cluster, and v is the typical speed of a star in the cluster. Let us approximate n ≈ 3N/(4πr3 c ), where rc is the cluster core radius, and v2 ≈ GN m/rc, where m is the characteristic stellar mass. We also assume that the Lada & Lada (2003) cluster size relation holds, simply scaled to higher densities, where rc ≈ r0(N/300)1/2. Finally, gravitational focusing must be included in the definition of the cross section, such that σ = πq2(cid:2)1 + 0.23r0/(qN 1/2)(cid:3), where q is the largest pericenter considered. Combining these relations, we find Γ ≈ 3 × 10−3 (cid:32) × 1 + 3 (cid:18)1 pc (cid:18) r0 r0 (cid:19)7/2(cid:18)300 (cid:19)1/4(cid:16) (cid:19)(cid:18)300 (cid:19)(cid:18)1000 AU N q (cid:17)2 (cid:19)1/2(cid:33) 1000 AU 1 pc q N (4) per star per Myr. The above rate is consistent to within a factor of three of the PA2009 rate for their virial N = 300 cluster. It shows the limiting behavior of γ(N ) and that the overall dependence of Γ on r0 is quite strong. As discussed in Section 2, we use the encounter rates from the PA2009 ICs that begin out of virial equilibrium, which reduces the effective r0 from 1 to about 0.7 pc. Using the above arguments, we would expect that an initial cluster core scaling of 0.7 pc would have encounter rates that are about 3.5 times the r0 = 1 pc rates for weak gravitational focusing. In comparison, PA2009 found that their non-virial (cold) simulations were enhanced by a factor ∼ 8 over their virial conditions, which is larger than we would expect. While starting with dynamically cold ICs leads to a smaller effective cluster size, it is not strictly the same as starting with a more compact cluster. For example, PA2009 attribute the additional enhancement to a larger fraction of bound cluster members in the -- 6 -- runs with non-virial ICs, ultimately boosting the encounter rate. Altogether, their cold ICs give a similar boost to the encounter rate for a cluster size of 1 pc that one would expect for r0 ∼ 0.55 pc, assuming weak focusing. 3. Numerical Experiments The effect that a close encounter will have on a planetary system depends on the peri- center of the encounter (e.g., Adams & Laughlin 2001; Adams et al. 2006). The odds of making significant changes to a planet with an orbit of, say, 1 AU due to a stellar encounter alone are very small, as a small pericenter is required for the planet to be strongly perturbed. However, as shown in Table 2, planets on wide orbits, i.e., with semi-major axes a ∼ 100 AU, stand a reasonable chance of being strongly perturbed. A close encounter with a system that has a planet or substellar companion on a wide orbit could cause a scattering cascade in a multiple-planet system, in the same spirit as investigated by Zakamska & Tremaine (2004), or produce an inclined outer gas giant/brown dwarf that could then cause Kozai oscillations on an inner planet. For these reasons, we have designed seven sets of simulations to explore the consequences of close encounters on the inclinations of planets on wide orbits and how these planets interact with other system members. 3.1. Scattering Experiment Design 3.1.1. N-Body Method We use the Bulirsch-Stoer (e.g., Press 2002) integrator in the Mercury package (Cham- bers 1999) to evolve realizations of five different system ensembles. Simulations are evolved for 108 yr and the typical energy error is ∼ 10−8(see appendix). The results obtained using these Bulirsch-Stoer integrations were also independently verified using a hybrid integrator from the same Mercury package, as well as using the GPU-based SWARM2 integrator using an Hermite integration scheme. The results obtained using all methods were qualitatively similar, but the Bulirsch-Stoer results were ultimately preferred due to their overall ability to conserve energy during the multiple close planetary-scattering events over the course of the 108 year integration. 2www.astro.ufl.edu/∼eford/code/swarm -- 7 -- 3.1.2. N-Body Initial Conditions All systems have a 1 M(cid:12) primary star and a wide-orbit planet with an initial semi- major axis of 100 AU. Two ensembles have three planets distributed inside the wide-orbit planet, and two ensembles have two planets inside the wide-orbit planet. For each of these cases, one ensemble has an incoming perturbing star. We set the perturber to have a stellar mass of 0.3 M(cid:12), set its initial velocity to 1 km/s, and place it randomly on the sky, as seen from the given planetary system, at a distance of 0.1pc. The perturber reaches its pericenter after ∼ 105 yr of evolution. To distinguish between the four cases, we adopt the following nomenclature: 3P1F1 refers to three planets interior to one wide-orbit planet with a perturber (flyby). 3P1F0 refers to a similar system, but with no flyby. 2P1F1 and 2P1F0 follow the same pattern. For reference, these names, as well as three additional simulations to be described later in the manuscript (3P1F1C, 2P1F1TD, and 3P1F1TD), can be found in Table 1. We perform 1,000 realizations of each of the 2P1F0 and 3P1F0 cases, and 3,000 real- izations of each of the 2P1F1 and 3P1F1 cases. In the flyby simulations, the larger sample sizes ensure that we can accurately probe both large- and small-pericenter flybys. All planets have masses drawn uniformly in log space between 1 and 10 MJ . While disk instability may produce a mass distribution that is more top heavy than assumed here, we are not strictly requiring that the formation mechanism must be disk instability. Moreover, the outcome of disk fragmentation is an active area of research, and the distribution of fragments that survive to become gas giants, brown dwarfs, or even stars, is not yet known (Boley et al. 2010; Kratter et al. 2010; Nayakshin 2010; Zhu et al. 2012). Planets interior to the wide-orbit planet are given a random inclination, uniformly distributed between zero and 0.1◦, and all eccentricities are less than 10−3. All wide-orbit planets have zero initial inclination with respect to the x − y plane. This plane is also taken to be normal to the stellar spin. Planet positions interior to 100 AU are placed randomly, uniform in a and in phase, but with the constraint that any new planet must be more than three mutual Hill radii from any neighboring planet and must have a semi-major axis a > 10 AU. We take the mutual Hill radius RH = 0.5(a1 + a2) [(m1 + m2) /3]1/3 for semi-major axes a and masses m, in stellar mass units, for planets 1 and 2. Figure 1 displays cumulative distributions for the initial semi-major axes, planet masses, and K, the number of mutual Hill radii between any two planets. Three-planet systems that have initial planet-planet spacings K < 3 exhibit strong interactions on timescales comparable to ∼ 10 orbits of the innermost planet orbit. The long-term stability of a system rises sharply for K > 3 (see Appendix B of Chatterjee et al. 2008). Because all systems have the same inner and outer bounds, the 2P1F0 and 2P1F1 systems are less tightly packed than the 3P1F0 and 3P1F1 systems. -- 8 -- The target distribution of pericenters for the perturber is set to be flat between 0 and 1500 AU. A flat distribution is biased, overall, toward more frequent close encounters than given by the results of PA2009. We will account for this bias in Section 5. This sampling is intended to provide better statistics for rare events. We do note that a roughly flat distribution is expected for small clusters (see γ functional form), so this biasing is largest for the largest of clusters. The distributions were extended to 1500 AU to ensure that we capture weak effects of flybys on systems. To verify that our calculations properly account for the bias in the q distributions, we also run one set of simulations with a q distribution given by γ = 1.3 between 0 and 1000 AU, which we call 3P1F1C. Recall that the shape and magnitude of the flyby frequency is dependent on the cluster member number N . For γ ≈ 1.3, the distribution corresponds to an N = 300 cluster, with η = 1 at 1000 AU after about 8 Myr. The actual pericenter distributions for the perturbers in simulations in 3P1F1 and 3P1F1C are given in Figure 2. The distribution for 2P1F1 is very similar to 3P1F1, so it is not shown. There are a few systems in which the flyby has a large pericenter, but as we will show, their evolution will not be altered significantly compared with flybys at 1500 AU. Adams & Laughlin (2001) found that the Solar System gas and ice giants will have their eccentricity or mutual inclinations doubled for flybys of binaries within a cross section of (400 AU)2, where the perturber masses and binary orbits were drawn from measured distributions (see paper for details). They also found that the cross section for ejections due to flybys is ∼ (73 AU)2. Scaling the cross section of the Solar System to the size of systems studied here, we expect to sample a full range from weak to very strong interactions. 4. Results In this section, we present results from the scattering experiments. We investigate the inclination distribution for each ensemble, and then explore the degree of radial mixing of planetary orbits, i.e., large changes in planets' semi-major axes. We also explore rare but non-negligible results, such as making hot Jupiters from planets that started at a ∼ 100 AU, as well as the effects of flybys on the dynamical stability of these systems. We compare the simulations with and without flybys in several ways. First, we investigate differences between the simulations with and without flybys, keeping in mind the following caveats: (1) interactions with low q will in general be overemphasized in the raw distributions. This is partially offset, however, by (2) the extension of the flyby distribution well beyond η = 1 for ∆t = 10 Myr. For example, an N = 300 cluster with γ ≈ 1.3, has an η = 1 surface at about q ≈ 750 AU after ∆t = 10 Myr. Giving equal weight to flybys out to 1500 AU will -- 9 -- tend to deemphasize the effects of flybys. Second, we show raw distributions with the data clipped to include only q > 300 AU because we expect flybys to have their strongest effects on a system when q (cid:46) 2.5Rsys, where Rsys is the orbital distance of the outer planet in a given system (see section 4.3). This selects moderate to weak interactions. Third, we weight each system's contribution to the cumulative distribution by the corresponding q, using qγ−1 from equation (1), to address results for realistic flyby frequencies. We assume N = 300 and ∆t = 10 Myr, and select only flybys with q < 750 AU. Finally, we include the results of 3P1F1C, which experienced a realistic flyby distribution for clusters of N ∼ 300. 4.1. Inclination Variations Raw cumulative distributions for the mutual inclinations of all planets at the end of the simulations (108 yr) are shown in Figure 3, where we show the maximum mutual in- clination between each pair of planets in each system. Note that these plots do not show the inclination relative to some fixed plane as might be considered in measurements of the Rossiter-McLaughlin (RM) effect. We will address RM measurements in Section 4.3. The distributions for simulations 2P1F0 and 3P1F0 are the same in each panel because no stellar flyby occurred. Perturbations from passing stars increase the fraction of planets on mutual inclinations i > 40◦ by an additional 2% of all systems for 3P1F1 and an additional one percent of all systems for 2P1F1. While these changes are relatively small compared with the entire distribution, the fraction of systems with large inclination planets is doubled to tripled in 2P1F1 compared with 2P1F0 and is increased by 40% in 3P1F1 compared with 3P1F0. When only encounters with pericenters q > 300 AU are considered, the fraction of planets with i > 40◦ is indistinguishable between simulations. For retrograde orbits, there is the possibility of an enhancement by flybys for the 4-planet systems, as the fraction of systems doubles from ∼ 1% to ∼ 2%. Caution must be taken, though, as the results rely on variations that are similar to the expected noise between the samples. Figure 3 also demonstrates that an extra planet in the system (2P1F0 compared with 3P1F0) raises the high-inclination distribution by a factor of 10 (from ∼ 0.5% to ∼ 5.5%) at 40◦. While this is a relatively small increase compared with the entire population, the fraction of planets that could effect Kozai oscillations is increased significantly. An extra planet in the system is much more important for producing planets with high mutual inclinations than are flybys, even if the system has a planet on an initial semi-major axis of 100 AU. In Figure 4 we replot Figure 3 using a logarithmic inclination axis, showing the full cumulative distributions for the mutual inclinations at the end of the simulations. As in In the ensembles without a perturber, the Figure 3, two different cuts for q are shown. -- 10 -- distribution has two clear components, with one reflecting the initial conditions of these models and another representing a broad, scattered population. The medians for the entire distributions are 0.073 and 0.085 degrees for 2P1F0 and 3P1F0, respectively. The median values for the inclination distributions with respect to the x− y plane (not shown) are about a factor of two lower, with 0.024 and 0.034 degrees for 2P1F0 and 3P1F0, respectively. For the subset of the distributions i > 0.3◦, which selects systems that have had strong planet- planet interactions, the median mutual inclinations are 3.5 and 11 degrees for 2P1F0 and 3P1F0, respectively, while they are 2.4 (2P1F0) and 5.9 (3P1F0) degrees with respect to the x − y plane. Table 4 summarizes these results and lists the initial median inclinations for comparison. In contrast, the distributions for simulations with flybys have a broad distribution from low to high inclinations. Even selecting only q > 300 AU does not erase this difference. The median mutual inclinations are 0.19 and 0.45 deg for 2P1F1 and 3P1F1, respectively. The distributions in Figures 3 and 4 can be weighted to reflect the inclination distri- butions for the expected perturber pericenter distributions of an N -member cluster. As discussed above, this weighting is dependent on the assumed cluster size, which we take to be N = 300. The results are shown in Figure 5. The distributions reflect the end-state incli- nations. As seen for the q > 0 AU cuts in Figures 3 and 4, flybys have a noticeable but small effect on the frequency of planets with mutual inclinations i > 40◦ when compared with the entire distribution. For smaller inclinations, in contrast, the cluster environment has a much stronger effect, and will tend to erase very cold initial conditions. Strict coplanarity should not be expected even in the absence of planet-planet scattering. The results for 3P1F1C are plotted along with the weighted distributions. The actual cluster distribution is very similar to the weighted one, both of which are similar to the full, flat distribution. In Figure 6, the medians (right) of the eccentricity (bottom) and of the mutual inclina- tion (top) as a function of q are shown. For each planet, the maximum mutual inclination is used, as done in the previous distributions. In addition, we show the maximum (left) eccentricity and mutual inclination for all systems in a given q bin. Bin widths are deter- mined by holding the number of systems per bin constant. The most distant q that do not form a full bin are not included. The symbols on the curves correspond to the median q for each bin and are placed along the abscissa at the center of the bin width. The maxi- mum inclination is sensitive to the number of planets/planet orbital density, and shows no dependence on q for 2P1F1 or 3P1F1. The median mutual inclination, in contrast, is not strongly dependent on the density of planets, for the cases studied here, but is dependent on q. The 2P1F1 and 3P1F1 distributions for the median mutual inclinations both follow roughly 25 exp(−x) + 2.15/(0.1 + x) degrees, where x = q/(100 AU), over the range shown (shown in red). The median inclination rises above one degree for flybys that are within -- 11 -- 4 times the radial extent of the planetary system, and all median inclinations for all q are larger than the medians in the simulations without flybys (see Table 4). As seen in the inclinations, the maximum eccentricity is not obviously influenced by flybys. Extreme outcomes (inclination and eccentricity) can be explained by planet-planet scattering alone. The median eccentricity is affected by the stellar birth cluster, with a profile that follows roughly max(0.4 exp(−x/1.1), 0.024) (shown in red). Broadly, the eccentricity results are consistent with the results of Heggie & Rasio (1996), who found that the change in the eccentricity of a binary due to a distant encounter transitions to an exponential form for sufficiently small q. A detailed comparison is difficult to make because the systems studied here all are multi-planetary systems, with the base level of eccentricity excitation higher than what is expected for most flybys. One noticeable difference between our results and those of Heggie & Rasio is that the maximum eccentricity remains below 0.5. This may be due to ejections of highly excited planets. Both the eccentricy and inclination profiles will be discussed further in Section 5.1. In the next section, we change focus from dynamical heating to major changes in the orbits of the planets as a result of scattering. 4.2. Major Changes in Planetary Orbits Planet-planet scattering, with and without close encounters, can lead to radial mixing of planetary orbits, bringing outer planets inward by several orders of magnitude and placing inner planets on very wide orbits. In Figure 7 we show the distribution of planet semi- major axes and inclinations relative to the x − y plane at the end of each simulation. Black crosses represent planets that were originally interior to the 100 AU planet, and blue circles represent the planets that were initially at 100 AU. Because the simulations with flybys have three times the number of systems as the simulations without flybys, we randomly select a third of the systems in 2P1F1 and 3P1F1 to show on the plots. Planets that are initially on wide-orbits can be scattered to a semi-major axis that is interior to the initial innermost planet. Flybys do increase the amount of this radial transport, but planet-planet scattering alone will lead to large scale mixing. It should thus be stressed that observing a planet at a given location in a disk is not by itself indicative of how and where the planet formed. This will be addressed again in Section 5.4. The connection between planet pericenters and eccentricity is shown in Figure 8. Most of the planets on small pericenters are the result of high-eccentricity orbits. Nevertheless, some planets do have small pericenters with eccentricities that are not near unity. In particular, an initially wide-orbit planet (a = 100 AU) has a q ∼ 1 AU and e ∼ 0.5 at end of the 3P1F0 simulation. The weighted eccentricity distribution is shown in Figure 9. The median -- 12 -- eccentricities are 0.015, 0.038, 0.019, and 0.047 for 2P1F0, 3P1F0, 2P1F1, and 3P1F1, respectively. Flybys do influence the distributions, but the dominant effect for producing large eccentricities remains the number of planets/planetary orbital density of the system. The large degree of radial mixing seen in these simulations, with and without flybys, gives rise to a small but non-negligible population of extreme systems. We explore these outcomes in the next section. 4.3. Extreme Outcomes In Figure 10, we show each planet's pericenter and inclination relative to the x − y plane at the time when the host system has any one planet reach the maximum inclination that ever occurs during the system's evolution, as well as at the time when the system has a planet reach the minimum pericenter that ever occurs. Only 3P1F0 and 3P1F1 are shown, and all planets in a given system are plotted. Parameter space is filled much more evenly than what is seen from the end state of each system only, with many wide-orbit planets spending time in the inner nebula or on highly inclined orbits. Many of the planets on retrograde orbits in the maximum inclination plot represent snapshots just before they become ejected. We emphasize that this does not represent the "end state" of the system, arbitrarily defined here as 108 yr. The purpose of selecting the systems at times of minimum pericenter and maximum inclination is to demonstrate that planets in any given planetary system could have occupied a much larger fraction of the disk than inferred from the end state. Furthermore, if some of these planets enter within 0.1 AU, then their evolution might be altered by tidal damping (the effects of which are not included in the plotted simulation ensembles), possibly locking the inclination of the planet into place while its semi-major axis is reduced (see also Section 5.4 Nagasawa et al. 2008, Payne et al. 2012, in prep). The fraction of systems with a planet that has penetrated q < 0.1 AU is 0.003, 0.03, 0.009, 0.04 for 2P1F0, 3P1F0, 2P1F1, and 3P1F1, respectively. Figure 11 shows unweighted cumulative distributions of the inclination relative to the x−y plane for all planets that have a minimum pericenter q < 0.1 AU. We use the inclination relative to the x − y plane because we are interested in the orbital-spin alignment between the planet and star for RM measurements. The unweighted distributions are shown because the fraction of planets is too small to apply weights with a reasonable degree of confidence. Nonetheless, the 3P1F0 simulation will serve as a baseline. In 3P1F0 and 3P1F1, the median inclination for planets that have q < 0.1 AU is between 40 and 60 degrees. Very small pericenters do occur for the 2P1F0 and 2P1F1 simulations, but these are rarer and are much more influenced by flybys than the 3P1F1 simulation. Flybys do little to change the distributions for the 3P1F1 simulation, but have -- 13 -- very strong consequences for sparsely populated planetary systems. Figure 10 shows the initial masses and semi-major axes for all planets that are included in the minimum pericenter distribution. Planets with a minimum q < 0.1 AU come from a range of locations, including very wide orbits, and have a range of masses. The large degree of radial mixing that arises from multiple scattering events in the same system can result in planet-planet collisions, even at these large separations. We define a merger event in the Mercury code to occur whenever planets pass within 2 RJ of each other. This gives an optimistic limit, as some of such collisions could only be hit and runs. In some cases, this places the planet over the deuterium burning threshold, assuming almost all of the combined planets mass is retained. The frequency of such collisions is (cid:46) 1%. The afterglows of such collisions may be observable, and in at least one case, may have already been observed (Mamajek & Meyer 2007; Miller-Ricci et al. 2009). In addition, debris trails may be a relic of collisions long after the afterglow has faded (Wyatt & Dent 2002). Another extreme outcome of planet-planet interactions is ejection of a planet from the system. Figure 12 shows the fraction of systems per q bin for which all planets remain bound to their system at the end of the simulations. The bin width is allowed to vary to ensure that there are equal systems per bin. The final bin that does not make the cutoff is ignored. For comparison, the fraction of all systems that have kept all of their planets is 0.93 and 0.47 for 2P1F0 and 3P1F0, respectively. Flybys are directly responsible for ejections for q (cid:46) 2.5 times the radial extent of the planetary system, Rsys, which is consistent with previous work (e.g., Adams & Laughlin 2001). No significant wing is seen for q > 250 AU, demonstrating that if flybys that are greater than 2.5 Rsys contribute to planetary ejections, the effect can be at most a few percent of the total fraction of systems that are destabilized. At first, these results may seem at odds with those of Malmberg et al. (2011), who find that flybys lead to a significant decrease in the long-term stability (108 yr) of a system, increasing the the fraction of systems that have had at least one ejection by factors ∼ 3- 9, depending on the mass distribution of planets. However, their measurements are taken for flybys inside q < 3.3Rsys. By this measure, we would also conclude that flybys are an important ejection impetus, so the results are consistent. However, we will show in Section 5.2 that the frequency of flyby-induced ejections for field stars is limited to a few to 10%. The formation mechanism for free-floating planets is primarily auto-ionization. Each system that experiences ejections can produce multiple free-floating planets. Using the results without applying any weightings, we find that about 0.08, 0.21, 1.1, and 1.2 free- floating planets per system are generated for 2P1F0, 2P1F1, 3P1F0, and 3P1F1, respectively. While flybys have a strong effect on sparsely populated planetary systems, having four instead of three planets in the system is more important for contributing to multiple ejections, -- 14 -- as there are more planets in the system that are able to decay (Juri´c & Tremaine 2008). The scenarios explored here do not produce ∼ 2 free-floating planets per system, as suggested to exist by (Sumi et al. 2011). However, these microlensing observations only probe separations (cid:38) 10 AU, and not strictly whether the planets are unbound. In this case, there is no contradiction between the observations and our results. It is also possible that systems have initially more giant planets than envisaged in these simulations. Most of the 2P1F0 systems have only one ejection, leaving a two-planet system. For 3P1F0, two planets are ejected, leaving a two-planet system. If we extrapolate this rate and assume that half of all systems experience an ejection cascade, six-planet gas giant systems would need be to be a common formation scenario for the microlensing results to correspond to true, free-floaters (although see Veras & Raymond 2012, for a more detailed estimate). 5. Discussion 5.1. Excitation due to Stellar Flybys: Rsys = 100, 30 AU Figure 6 demonstrates that the maximum mutual inclination that we can expect for a given system is most sensitive to the number of planets in the system, and not to the closest approach of a stellar flyby. In contrast, it also demonstrates that the median mutual inclination among planets in a given system is sensitive the pericenter of a stellar flyby, and not strongly to the number of planets in a system. We can use this result to estimate the median mutual inclination for planets in planetary systems that come from a range of initial cluster sizes. We find that imedian ≈ max(25 exp(−x) + 2.15/(0.1 + x), 0.1) degrees, where x = q/Rsys degrees, where Rsys is the system's initial radial extent. The floor of 0.1 degrees takes into account the base-level inclination excitation from planet-planet scattering alone, which we take to be the average of 2P1F0 and 3P1F0 from Table 4. In the simulations presented here, Rsys = 100 AU. The expectation median inclination for a distribution is (cid:90) N1 N0 (cid:104)imedian(cid:105) = (cid:18)(cid:90) qlarge dξN dN 0 (cid:19) ∂ξ ∂q imediandq dN/η(cid:48). (5) We restrict the number of encounters to be less than unity for a given N and q because we do not want to add extra weight to very distant encounters. As such, q is only integrated until ev- ery system on average has one encounter (qlarge). A similar approach can be taken for the me- dian eccentricity in systems by replacing imedian with emedian ≈ max(0.4 exp(−x/1.1), 0.026). We set the basal eccentricity to 0.026, which is the average of the median eccentricities for the 2P1F0 and 3P1F0. The results are shown in Table 3. The median mutual inclination for a distribution of planetary systems that had a planet at 100 AU initially will be between -- 15 -- about two and five degrees. Likewise, the median eccentricity will be between 0.04 and 0.08, depending on the cluster size. The relationships for imedian and emedian depend on Rsys. If we take Rsys = 30 AU, analogous to the Solar System's size for the major planets, then the expected median inclination is only ∼ 1 degree and the (cid:104)emedian(cid:105) is between 0.03 and 0.04. We also use our results to explore the frequency of ejections. From Figure 12, flybys trigger additional ejections for q/Rsys (cid:46) 2.5. The ejection distribution is sensitive to the number of planets, with the frequency of ejections about 8 times larger in 3P1F0 than in 2P1F0. As with the other profiles, we take the average between these basal levels. The functional form for the fraction of systems that retain all of their initial planets is then fsurvive = 0.7(q/Rsys) for q/Rsys < 2.5 and 0.7 otherwise. Table 5 gives the results. Flybys have a 4-15% effect on systems with Rsys = 100 AU and, based on scaling, we predict a 1-4% effect on systems with Rsys = 30 AU. The flyby rate and the distribution of cross sections of planetary systems is a determining factor for the efficacy of flybys on shaping planetary system architectures. In the next section, we place constraints on this distribution and build on the results from the above Rsys = 100 and 30 AU cases. 5.2. Excitation due to Stellar Flybys: Integration Over Expected Rsys The fraction of highly inclined planets that are produced by flybys is dependent on the effective cross section of the planetary system. In general, the effective cross section is not necessarily equivalent to the geometric cross section, but because we are assuming that the orbits for the outermost planets are initially circular, we take the cross section to be πa2 outer , where aouter is the outermost planet's semi-major axis. While the distribution of planetary sizes is at this time unknown, we can use the distribution of specific angular momentum among low-mass cloud cores as a proxy. Tables 3, 4, and 5 from Caselli et al. (2002) provide data for 20 sources that have enough information for estimating a given core's angular momentum, J (see also, e.g., Shu et al. 1987; Myers & Benson 1983; Goodman et al. 1993; Barranco & Goodman 1998, for additional sources and discussion regarding the internal kinematics of cloud cores). We convert from the Caselli et al. derived velocity gradient values G to J by assuming virial equilibrium and constant density cores, which allows us to write J ≈ 0.4R2 core(4.86 × 102G/nvir)1/2 (see, e.g., Goodman et al. 1993). Here G is in km s−1 pc−1 and the average number density of the core nvir is in cm−3. Based on these assumptions, the initial size of the protoplanetary disk is RJ ≈ J 2/(GMvir), i.e., the radius we expect collapse to be halted by the system's angular momentum. In this estimate, we use the virial cloud core mass Mvir as derived by Caselli et al. As shown in Table 5, about 30% of systems could -- 16 -- possibly form a substellar companion at semi-major axes 80 AU or greater. This estimate is admittedly crude, but allows us to extend our estimates from section 5.1 by accounting for disk sizes. We repeat the calculations in the previous section, but for Rsys = 20, 60, 100, 140 and 180 AU, where the result for each bin is weighted by the corresponding fraction. Under these assumptions, (cid:104)imedian(cid:105) ∼ 1 to 3 degrees, (cid:104)emedian(cid:105) ∼ 0.03 to 0.06, and fsurvive ∼ 0.6 to 0.7. At least for the systems studied here, most massive planets remain bound to their host star, with about 30-40% of systems generating free-floating planets. We conclude that planet formation and subsequent planet-planet interactions, not stellar encounters, determine the final eccentricities and inclinations of the typical planetary system. If, on average, the outermost planet in a system forms at a fraction of RJ, then the effect of flybys is further marginalized than what we find here. Even if planets form at RJ, they will not necessary stay there in a rapidly evolving, young, massive disk (Baruteau et al. 2011; Michael et al. 2011; Zhu et al. 2012). This would reduce planetary system sizes and again the effects of flybys would be further marginalized. In contrast, it may be possible to move massive planets outward (Crida et al. 2009) as disks transition away from their initial, massive state. This would increase the influence of flybys. The relative importance of these effects at this time is unknown, so although only approximate, we find our above calculation to be reasonable with the information available. 5.3. Observationally Constraining Formation Scenarios for Wide-Orbit Planets The formation mechanism of wide-orbit planets can potentially leave distinct dynamical signatures. First, consider the fraction of planets with semi-major axes between 80 and 200 AU. If we assume that every disk forms a planet in situ at its nascent cloud's angular momentum barrier and we further assume that this planet does not migrate, then about 18% of systems should have a planet with a semi-major axis in the chosen annulus (30% of all systems with a survival rate of about 60%) . We do note that even if planets form at large radii, they will not necessarily stay at this location (Baruteau et al. 2011; Michael et al. 2011). Nonetheless, the rate of occurrence may be much larger than the 3% that is found in scattering simulations (Veras et al. 2009, see their Fig. 4). If migration or the rate of planet formation on wide-orbits produces a frequency that is indistinguishable from pure scattering, we still expect to observe significant differences in the orbital distributions between in situ formation and scattering only. From the scattering experiments of Veras et al. (2009), planets with semi-major axes between ∼ 100 and 200 AU have eccentricities that are roughly evenly distributed between 0.4 and 1.0, with the -- 17 -- lower bound increasing with a such that the lower limit on the eccentricity is about 0.8 by a ∼ 1000 AU. In contrast, the median eccentricity for the outermost planet in the simulations presented here is ∼ 0.13, where we have averaged the median eccentricities for a ∼ 100 and 200 AU. These eccentricity differences lead to highly distinguishable features in observable distributions. We explore the effect that different eccentricity distributions have on observables by making the following assumptions and cuts in parameter space: We limit the semi-major axes of the planets to be between 100 and 200 AU. In this regime, the planet semi-major axis distribution due to planet-planet scattering, based on the simulations of Veras et al. (2009), is dNpl/da ∝ a−2.75, where the power law profile is taken directly from their simulation data. The a distribution for in situ formation is not so cleanly defined from core velocity gradients, so we adopt the same profile for the following comparisons. The underlying a distribution will have an effect on the profiles, but strong features are insensitive to this assumption. We produce distributions of planet radial separations from the star and of star-planet relative radial and azimuthal proper motions by using Keplerian orbits with a range of eccentricities and semi-major axes. Each orbit's contribution to the distribution is weighted by a−2.75 and by the corresponding inverse orbital period. The inverse period weighting is necessary to keep planets on longer orbits from having extra weight, as only the fraction of the planet's orbit that is within the chosen annulus matters. Radial and tangential velocity components are transformed into proper motions relative the host star to highlight observable constraints, with the tangential proper motion given as a deviation from a circular orbit at the given location. We assume a distance of 10 pc, and the distributions are assumed to be face-on. Figure 13 shows the distributions of the radial separation, R, the radial proper motion, VR, and the deviation of the tangential proper motion, Vphi − Vc, for e = 0.13, e = 0.40, and an integration over an even distribution of eccentricity in the range [0.4, 0.95]. Proper motions are strongly peaked near VR ∼ (µ/a(outer))1/2 e and Vφ− Vc ∼ (µ/a(outer))1/2 e/2, where a(outer) is the largest semi-major axis included in the distribution (200 AU here) and µ = G(mstar + mplanet). The radial distribution goes as r−2.75 in regions where the pericenter and apocenter are both within the 100-200 AU annulus, which is expected from the assume semi-major axis profile. The peak in the proper motion distributions for the range of e is broadened by high eccentricities, with a tail extending to large e, but remains near the distribution with only e = 0.4 because the lower eccentricity planets are more likely to be seen within a fixed-width annulus. The radial distributions between a fixed e and the range of e have very different shapes. A single component distribution of the proper motion can strongly distinguish between scattering only and a significant in situ formation population, even if the underlying semi- -- 18 -- major axis distribution is unknown. Pure scattering will exhibit a peak in the VR distribution that is at proper motions that are three times larger than a peak that corresponds to only in situ formation on initially circular orbits. As a result, a limited search may be able to test whether a large component of wide-orbit in situ formation planets exist. The model distributions cannot be used alone to constrain the formation mechanism of any single planet, but can be used to comment on how typical a given system is under the premise of the model. As an example, consider Fomalhaut b. Based on Kalas et al. (2008), the proper motion of the planet candidate is about 100 mas/yr, largely North and in the projected tangential direction. Taking this at face value, the tangential proper motion deviation is about 7 mas/yr for R = 119 AU, which would correspond to e = 0.13 for Fomalhaut b's motion to lie on the most probable deviation of tangential proper motion. This eccentricity is the same as that already derived by Kalas et al., and is at the median value for in situ planet formation based on our scattering experiments. 5.4. From Wide-Orbits to Hot Jupiters In Section 4.3, we found that a small fraction of our systems have a planet with a pericenter q < 0.1 AU during some point of the system's evolution. The fraction of systems with this outcome varies between tenths and several percent, depending on the total number of initial planets in the system and whether the system ever experiences a flyby. We refer to these planets as hot Jupiter candidates, as many approach the star with a pericenter that is small enough for tidal friction to be important. Some of these candidates are initially interior to the 100 AU planet, but some are initially on very wide wide-orbits (see Fig. 11). When wide-orbit planets scatter close to the star, they do not do so directly. Instead, they have multiple encounters with the inner planets, giving the planet a small pericenter. To explore this possibility, we integrate 1000 systems from the 2P1F1 and 3P1F1 ensembles (500 each) using the same Mercury integration scheme that is used in the other ensembles, but with the inclusion of a tidal damping model (Nagasawa et al. 2008). We refer to these new ensembles as 2P1F1TD and 3P1F1TD to distinguish them from 2P1F1 and 3P1F1, which are run without tidal damping. One system in 2P1F1TD and three in 3P1F1TD form a hot Jupiter from a planet that is originally at 100 AU. While this is a small fraction of all planetary systems, it is potentially a non-negligible fraction of hot Jupiters. Figure 14 shows an example of a hot Jupiter that forms from an inward scattering cascade of a planet that is initially at 100 AU. Note that the planet is also highly inclined, with an end inclination relative to the x − y plane i ∼ 70◦, a semi-major axis a = 0.019, and an e = 0.013. This emphasizes again that a planet's observed location alone is not indicative of where or by what mechanism the planet formed. -- 19 -- The frequency of hot Jupiter candidates increases with decreasing planet mass (Fig. 11), and at least half are expected to have an inclination i > 40◦ relative to the x − y plane. The mass cutoff in our simulations is 1 MJ , but continuing this trend to the masses of the observed hot Jupiter population suggests that scattering should produce even more hot Jupiters at masses less than those studied here. Whether any given planet can ultimately become tidally captured will depend on the properties of that planet, with a strong, but nontrivial dependence on planetary radius at fixed mass (e.g., Ivanov & Papaloizou 2004; Mardling 2007). If the initial radius of a planet is dependent its formation mechanism, then hot Jupiters might be used a tool for exploring different modes of planet formation. For example, if planets that form by direct instability have higher specific entropy on average than core accretion planets of comparable composition and mass, then the radius evolution of these planets could be appreciably different (e.g., Spiegel & Burrows 2011), which could have consequences for tidal capture of these objects. 5.5. Outer Planets from Inner Planets Veras et al. (2009) have shown that a population of eccentric, wide-orbit planets can be produced by planet-planet scattering. Their simulations included a high-density of planets between 3 and 7 AU. In the simulations presented here, planets are already on moderate and wide orbits, which can give rise to very different scattered population. First, consider the fraction of systems that have a planet at the end of the simulation with a radial separation from the star > 90 AU. For each ensemble, those fractions are 0.97, 0.92, 0.85, and 0.83 for 2P1F0, 2P1F1, 3P1F0, and 3P1F1, respectively, which demonstrates that at least a few to ten percent of systems with a planet initially at 100 AU will not be observed to have one, even in the absence of flybys owing from planet-planet scattering and auto-ionization. If we exclude all planets that are initially on wide orbits, the fraction of wide-orbit planets is 0.016, 0.038, 0.17, and 0.20 for 2P1F0, 2P1F1, 3P1F0, and 3P1F1, respectively. In many cases, the outermost planet is significantly farther out from the star at the end of the simulations than the outermost planet at the beginning of the simulations. Between about one and ten percent of systems with wide-orbit planets at the end of the simulation do not have the initial wide-orbit planet at such large separations. Figure 15 shows a population of very wide orbit planets for 3P1F0. All planets that are bound and have a separation > 90 AU at the end of the simulation are shown except for a small population that extend beyond the plot limits. Circles represent planets that were not initially on wide orbits, while triangles represent the initial 100 AU planets. The colorbar represents the log of the initial semi-major axis in AU. Most of the triangles are clustered around a, r = 100, 100 AU. Planets on wide orbits can come from a very wide range of initial semi-major axes. The widest planets r > 1000 AU -- 20 -- are preferentially, but not entirely, the planets that began farther out in the disk. They are not, however, preferentially the initial 100 AU planet. Recall that 3P1F0 does not include flybys. For planets that were not initially at 100 AU, a population of high pericenter, very long period planets is also possible. In this context Sedna-like orbits can be produced by a wide-orbit planet that is lost from the system. Such a loss can be through ejection, or if planet formation can take place during the earliest stages of outer disk evolution, the scatterer could be a transient clump (Boley et al. 2011). 5.6. Limitations of Current Study The current study explores a single mass for the perturbing stars (Mp = 0.3M(cid:12)) and for the host stars of the planetary systems (MH = 1.0M(cid:12)). In addition, it excludes encounters with binaries. Relaxing any one of these restrictions will have an effect on our results, which we discuss here. First, the potential perturbation from a flyby on a planetary system scales with the mass of the perturber, so we expect more massive stars to lead to more ejections and to produce a larger median inclination among planetary systems than less massive perturbers. Because we set Mp to be near the median stellar mass, half of the flybys will lead to more energetic perturbations than captured in our simulations and half of the flybys will lead to less energetic perturbations. We therefore do not expect the choice of perturber mass to be a major source of error in the study. Second, the results presented here are most relevant to planetary systems around solar- type stars, resulting from our choice of MH. Before the results can be extended to planetary systems around a distribution of host stars, we must consider the following: (1) Low-mass host stars will have planets that can be more easily ionized by a perturber compared with higher-mass host stars, ceterus paribus. (2) The specific angular momentum distribution of cloud cores may depend on cloud core mass, affecting the size distribution of planetary systems. (3) Independent of angular momentum distributions, the formation of gas giants on wide orbits may depend on host star mass. While, including a distribution for MH is necessary for understanding the role of flybys on planetary systems in general, it is difficult at this time to assess how our results will change for a realistic distribution of host star masses. Third, flybys by binaries can cause stronger perturbations to planetary orbits than flybys by single stars, owing to increased interaction cross sections and to resonant interactions between the planets and the binary whenever their periods are comparable. Early studies of -- 21 -- the multiplicity of solar-type stars suggest that about half of solar-type stars are in binaries (Duquennoy & Mayor 1991), so it would therefore seem most relevant to focus on interactions between binaries and planetary systems instead of single stars and planetary systems as done here. However, recent studies show that multiplicity is a strong function of stellar mass (e.g., see Fig. 12 of Raghavan et al. 2010), with the implication that most field stars (∼ 70%) are single (Lada 2006). In addition, only a fraction of binaries will have separations/periods that will significantly alter the interaction. For example, both very short-period binaries and very long-period binaries will appear approximately as a single perturber. To estimate the relevant fraction of binaries that will impact the planetary systems explored in this study, we integrate over the binary period distribution for solar-type stars given by Raghavan et al. (2010). The fraction of binaries with periods between 10 and 3000 yr (∼ 5 and 200 AU) is 0.41. Combining this fraction with the frequency of binaries among all stars reveals that strong interactions with binaries, compared with single stars for the same closest approach, should be expected for ∼ 12% of encounters. Only a fraction of these encounters will have relevant close approaches to the planetary system, so the results we present here are not obviously affected by the exclusion of binaries. There are several additional caveats that should be mentioned. (1) The binary period distribution is valid for solar-type stars in the field, while what is desired is multiplicity of stars while they are in their natal cluster. (2) Observations suggest that the semi-major axis distribution of binaries is dependent on binary mass, with low-mass binaries being much more compact (Siegler et al. 2005). This will tend to decrease the influence of binaries compared with single stars by reducing the interaction cross sections of the binaries. Finally, (3) we note that a very short-period binary will have a stronger perturbation on a planetary system than a single perturber on average by virtue of the binary being two stars. This effect will give a slight skew of the effective median mass of perturbers to higher mass, but we do not expect this to be a major source of error in our results. Overall, we do not find that the results of this study should be significantly affected by our choice to focus on single, 0.3 M(cid:12) perturbers. 6. Conclusions We have presented the results of a series of scattering experiments that investigate the dynamical outcomes for multi-planet systems that have planets initially on wide-orbits. The experiments compare system architectures and scattering histories between chiefly four ensembles: two of the ensembles do not include stellar flybys, while two include flybys by a 0.3 M(cid:12) perturber. All systems contain 1 planet at 100 AU. Two of the ensembles are made -- 22 -- of systems with two planets interior to the 100 AU planet, and two ensembles with three planets interior to the 100 AU planet. The four-planet systems fill the same semi-major axis range as the three-planet systems, so the four-planet systems are more densely packed. The importance of flybys on these system architectures is then evaluated by direct comparisons between the non-flyby and flyby ensembles. When possible, our results are rescaled for a distribution of planetary system sizes that are derived from literature values of cloud core velocity gradients and/or integrated over a range of natal cluster sizes for comparisons with field star populations. We find the following key results: 1. High mutual inclinations within planetary systems are more likely to be due to planet- planet interactions than due to stellar flybys. We find that for mutual i > 40◦, flybys increase the fraction total fraction by about 1%. Although a small increase overall, flybys can double the number of highly perturbed planets. 2. Low mutual inclinations are strongly affected by stellar flybys. Even if planets are born perfectly coplanar, the system's natal cluster will seed a substantial inclination dispersion. We find the median inclinations for the three and four-planet systems at the end of the simulations to be about 0.24 and 0.86 degrees, respectively. The same systems without flybys have a mutual inclination of about 0.08 and 0.12 degrees. Figure 5 shows a clear separation between the mutual inclinations of systems with and without flybys. Initial conditions for planet-planet scattering studies with very small initial inclinations may not be realistic. 3. Both high and low eccentricities are affected by the presence of flybys, although the effects of flybys remains small compared with the initial density of planets. The median eccentricities are 0.015, 0.038, 0.019, and 0.047 for 2P1F0, 3P1F0, 2P1F1, and 3P1F1, respectively. 4. Radial mixing of planetary orbits takes place in all simulations. Wide-orbit planets can be placed on moderate orbits, and moderate-period planets can be placed on very long-period orbits. Observing a planet at a given location in a disk is not by itself indicative of where and/or how it formed. Moreover, the scattering history of a planet can be complex, with the possibility that some planets will spend time in both short and long-period orbits during the system's evolution. 5. The scattering process can lead to very extreme outcomes, including turning a wide- orbit planet into a hot Jupiter. In the four-planet simulations without flybys, nearly 3% of the systems have a planet that at some point has a pericenter inside 0.1 AU. In all cases the planets are initially at distances greater than 10 AU, and several are -- 23 -- planets that are initially at distances of 100 AU. The planets that are scattered to such small pericenters are preferentially the lower mass planets in the simulations. We run a subset of the flyby simulations using a tidal damping model (2P1F1TD and 3P1F1TD), and show an example of a hot Jupiter that is formed from the inward scattering of a planet that is initially at 100 AU. The planet is also highly inclined, with an end inclination relative to the x − y plane i ∼ 70◦, a semi-major axis a = 0.019, and an e = 0.013. 6. The inclination distribution relative to the x − y plane, here assumed to be normal to the stellar spin axis, is large for the planets that penetrate 0.1 AU at the time of smallest pericenter. At least half of these planets have inclinations greater than 40◦. 7. Stellar flybys can directly cause ejections of planets for q that are within ∼ 2.5 times the outermost planet's semi-major axis, which is consistent with results in the literature. The frequency of ejections is strongly dependent on the initial density of planets in the system. After weighting the effects of flybys to account for the expected encounter rates and including a distribution of planetary system sizes, we find that ∼ 30 to 40% of systems experience at least one ejection, with the typical ejection outcome leading to two-planet systems, consistent with previous work. A few to about 8% of systems have ejections that are induced by flybys, demonstrating that auto-ionization is the dominant mechanism for forming free-floating planets. In 2P1F0, the fraction of planets ejected per system is 0.08, while this number jumps to ∼ 1 for 3P1F0. Flybys do increase the total number of ejected planets, with a perturber's influence being greatest on 2P1F1. The total number of ejected planets per system is 0.21 and 1.2 for 2P1F1 and 3P1F1, respectively. 8. The dynamical signatures for long-period planets that are born in situ versus those that are scattered onto long periods are distinct due to differences in the expected eccentricity distributions. Limited observations of relative proper motions between a companion and the host star may be able to constrain the contribution of in situ formation of planets on wide orbits, even using one component of the proper motion. 9. Planet-planet scattering in systems where one planet was originally on a wide orbit can give rise to planets on very long-period orbits (a ∼ 1000 AU) with pericenters ∼ 100 AU. In the simulations explored here, the fraction of systems that have a planet with a radial separation > 90 AU at the end of the simulation are 0.97, 0.92, 0.85, and 0.83 for 2P1F0, 2P1F1, 3P1F0, and 3P1F1, respectively. If planets that were initially on a wide orbit are excluded, the frequency of planets with radial separations > 90 AU ranges from 1 to 20 % from 2P1F0 to 3P1F1. -- 24 -- 7. Acknowledgement We thank Dimitri Veras for sharing his scattering results. We thank Fred Rasio and Smadar Naoz for helpful discussions, and we thank Fred Adams and the anonymous referee for comments that improved this manuscript. A.C.B.s support was provided by a contract with the California Institute of Technology (Caltech) funded by NASA through the Sagan Fellowship Program. M.J.P.'s and E.B.F.'s contributions are supported by the National Science Foundation under grant no. 0707203 and is also based on work supported by NASA Origins of Solar Systems grant NNX09AB35G. T The authors acknowledge the University of Florida High-Performance Computing Center for providing computational resources that contributed to the research results reported here. A. Energy Conservation In Figure 16, we show cumulative histograms of the integrator energy conservation for the simulations. The median energy error for all simulations is < dE/E ∼ 10−7, and almost all systems have energy conservation < 10−5. There are seven systems in the simulations with flybys (2P1F1 and 3P1F1 combined) that have energy conservation > 10−4, but they do not change the general results. To make sure that systems with very poor energy conservation are not biased toward systems of interest, we list the median energy errors for all systems that have a planet with q < 0.1 AU at some time during the system's evolution, which are 2 × 10−7, 1 × 10−6, 7 × 10−7, and 1 × 10−6 for 2P1F0, 2P1F1, 3P1F0, 3P1F1, respectively. We focus on these simulations because they all have experienced strong scattering events. The maximum tolerable error can be estimated from the most extreme mass ratio in the problem, which implies that we require error conservation to be << 10−3. The median error for systems with some of the strongest scattering events meets this criterion. -- 25 -- Fig. 1. -- Cumulative distributions of the initial conditions for semi-major axis (left), mass (center), and planet orbital separation in number of mutual Hill radii K (right). Fig. 2. -- Histogram for the pericenters of the perturber q. A few systems extend beyond the cutoffs of 1500 and 1000 AU for the flat (γ = 1) and cluster (γ = 1.3) distribution, respectively. These systems are included in the rightmost bin shown here. The histogram for 2P1F1 (not shown) is very similar to 3P1F1. -- 26 -- Fig. 3. -- Raw cumulative distributions for the end-state mutual inclinations of all planets for all systems that had a q > 0 and 300 AU in the left and right panels, respectively. Simulations 2P1F0 and 3P1F0 are the same in each plot because no stellar flyby occurred. The maximum mutual inclination is taken for each planet. Fig. 4. -- Similar to Figure 3, but with logarithmic inclinations bins to allow all inclinations to be compared. Simulations without flybys have two components in the distribution, with one reflecting the initial conditions at the other a high-inclination, scattered component. Flybys blend the peaks in the profile and shift the median inclination to higher values, even for q > 300 AU. See Table 4 for median inclinations. -- 27 -- Fig. 5. -- Cumulative distributions for the mutual inclinations of all planets for all systems at the end of the simulations, weighted to account for the expected distribution of q. The initial number of planets or planet orbital density is the primary determinant for the number of highly inclined planets. Nonetheless, flybys still have an effect on low inclinations. Even if planets are born perfectly coplanar, the birth cluster of the system will result in an intrinsic inclination spread. To weight 2P1F1 and 3P1F1, each system's contribution to the histogram is scaled by qγ−1. The results for 3P1F1C are also shown (no weighting necessary), and are consistent with the unweighted 3P1F1 distribution in Figure 3. -- 28 -- Fig. 6. -- Top: The absolute maximum mutual inclination of all systems per q bin (left) and the median mutual inclination of all systems per bin. Flybys can alter inclinations for q at least out to 10 times the radial extent of a system. Bins are determined by demanding an equal number of systems per bin. The most distant q that do not form a full bin are not shown. Bottom: similar as in the top row, but for the maximum and median eccentricity distributions. The maximum values can reflect entirely internal processes, i.e., planet-planet excitation and scattering, while the median values do not. Fits to the data are shown by the red curves and are described in Section 5.1 -- 29 -- Fig. 7. -- Semi-major axes (SMA) versus inclination for each planet at the end of the simulation. Black crosses represent all planets that were initially interior to the 100 AU wide-orbit planet (blue). Scattering usually places inner planets on wide orbits, but can also place wide-orbit planets onto orbits of a few AU. -- 30 -- Fig. 8. -- Similar to Figure 7, but for planet orbital eccentricity and pericenters. -- 31 -- Fig. 9. -- Cumulative and specific distribution functions for the eccentricities of all planets weighted to account for a realistic distribution of q for a cluster of N = 300. Perturbations by passing stars have a small effect on the eccentricity distribution of the planets, with planet-planet excitation clearly dominating the distribution function. -- 32 -- Fig. 10. -- Pericenter and inclination relative to the x − y plane at the time when the host system has any one planet reach the maximum inclination that ever occurs during its evolution, as well as the pericenter and inclination at the time when the system has a planet reach the minimum pericenter that ever occurs. Black crosses represent planets with initial positions inside the 100 AU wide-orbit planet (blue circles). The star's spin is envisaged in these simulations to be normal to the x− y plane. There is a pileup of planets on retrograde orbits for the maximum inclination plots. Many of these planets are in the process of being ejected. -- 33 -- Fig. 11. -- Left: Cumulative distributions (unweighted) for all planets that have a pericenter q < 0.1 AU at any time during the simulation. Right: Initial planet semi-major axes and masses for all planets that orbit within q = 0.1 AU at some point during the simulations. Lower mass planets are preferentially scattered onto the more highly eccentric orbits. -- 34 -- Fig. 12. -- The fraction of systems per q bin for which all planets remain bound to their system at the end of the simulations. The bin width is allowed to vary to ensure that there are equal systems per bin (chosen here to be ∼ 150). The final bin that does not make the cutoff is ignored. For comparison with the fractions at large q, about 0.93 and 0.47 (lines) of all systems experience no ejections in 2P1F0 and 3P1F0, respectively. -- 35 -- Fig. 13. -- Radial separation (left), the relative radial proper motion (center) and the deviation of the relative tangential proper motion (right) distributions of planets on wide orbits for several eccentricity distributions. We assume a distance of 10 pc for conversion of the velocities to star-planet relative proper motions. Fig. 14. -- Top: The pericenter q, semi-major axis a, and apocenter Q for each planet in one of the 3P1F1TD systems. Bottom: The evolution of the inclinations relative to the x−y plane for the same planets shown in the top panel. The planets in the system become dynamically unstable, and the outermost planet becomes the innermost one. Its orbit eventually becomes eccentric enough for the planet to pass by the star at ∼ 0.02 AU, where dynamical tides quickly circularize the orbit. In this example, a hot Jupiter with an inclination relative to the x − y plane i ∼ 70◦, a semi-major axis a = 0.019, and e = 0.013 is made from a planet that was initially at 100 AU. -- 36 -- Fig. 15. -- A population of very wide orbit planets for 3P1F0, for which all planets with separations > 90 AU at the end of the simulation are shown. Circles represent planets that were not initially on wide orbits, while triangles represent planets that are initially at 100 AU. The colorbar represents the log of the initial semi-major axis in AU. Most of the triangles are clustered around a, r = 100, 100 AU. For the planets with r > 1000 AU, their orbits were interior to the 100 AU planet, but are preferentially at large a. Several points do extend beyond the limits of this plot. -- 37 -- Table 1: Definitions of symbols and abbreviations. Symbol Definition a, ai e i q, qf lyby N N0&N1 Q rc Γ(q, N ) γ(N ) ξm ξN η η(cid:48) ∆t η η(cid:48) Γ(cid:48) n σ v m Rsys r0 RH M(cid:12) K imedian emedian vK vr vphi vc δvr δvz qlarge fsurvive J G Mvir Npl(a) µ 2P 1F 0 2P 1F 1 3P 1F 0 3P 1F 1 3P 1F 1C Planetary Semi-major Axes Planetary Eccentricity Planetary Inclination Pericenter, Pericenter of Stellar Flyby Number of Stars in Cluster Min & Max Values of N considered Viral Parameter: = Total Kinetic Energy Total Potential Energy Core radius of stellar cluster Rate of encounters with pericenters < q in a cluster of size N Encounter rate exponent Cluster stellar-mass function Cluster stellar-number function Average # of encounters per star in ∆t Fraction of field stars experiencing at least 1 encounter in ∆t Typical interaction period within a cluster. Typically 10. # of encounters for systems with > 1 encounter Rate of Stellar Flybys Cluster stellar density Interaction cross section Average Speed of Star in Cluster Characteristic mass of Star in Cluster Outer Radius of Planetary System/Initial Disk Size Typical cluster size scale Mutual Hill Radius Solar Mass Planetary Separation in units of Mutual Hill Radii Median mutual planetary inclination Median mutual planetary eccentricity Keplerian Orbital Velocity Radial Velocity Azimuthal Velocity Circular Speed Radial Velocity Dispersion Azimuthal Velocity Dispersion Flyby q for which all stars are expected to experience one encounter Fraction of systems that retain all their planets Total Angular Momentum of a Cloud Core Velocity Gradient of Cloud Core Viral Cloud Core Mass Planetary Semi-major Axis Distribution G(mstar + mplanet) Simulations with 2 planets interior to 1 wide-orbit planet, with NO flyby Simulations with 2 planets interior to 1 wide-orbit planet, with a stellar flyby Simulations with 3 planets interior to 1 wide-orbit planet, with NO flyby Simulations with 3 planets interior to 1 wide-orbit planet, with a stellar flyby Simulations with 3 planets interior to 1 wide-orbit planet, with a stellar flyby q distribution of γ = 1.3 2P 1F 1T D Simulations with 2 planets interior to 1 wide-orbit planet, with a stellar flyby and tidal damping 3P 1F 1T D Simulations with 3 planets interior to 1 wide-orbit planet, with a stellar flyby and tidal damping Units AU AU AU AU - - - pc # star−1 Myr−1 - - - Myr AU pc AU km s−1 pc−1 - - - - - - - Definition Section 3 3 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2.1 2.1 2.1 2.1 2.1 2.1 2.1 3.1 3.1 3.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.1 5.2 5.2 5.2 5.3 5.3 3.1.2 3.1.2 3.1.2 3.1.2 3.1.2 5.4 5.4 -- 38 -- Table 2: The fraction of field stars that have had a close encounter < q for different q and range of cluster sizes N0 to N1. The column η = η(q, N0, N1, ∆t, dξN dN ) represents the average number of encounters per field star for flyby pericenter q. The definition of η (equation 3) includes stars that have had multiple encounters. The quantity η(cid:48) does not include multiple encounters, so it represents the fraction of field stars that have had at least one encounter. The average number of encounters among field stars that have had at least one encounter is given by the ratio of η to η(cid:48). We use ∆t = 10 Myr for these calculations. See Section for more details. q(AU) N0 N1 104 100 104 100 104 100 104 200 104 200 104 200 104 300 104 300 104 300 104 1000 104 1000 104 1000 10 30 100 10 30 100 10 30 100 10 30 100 η 0.18 0.098 0.044 0.38 0.22 0.011 0.59 0.35 0.19 2.3 1.5 1.0 η(cid:48) η/η(cid:48) 0.18 1 0.098 1 0.044 1 0.34 1.1 0.22 1 0.11 1 0.44 1.3 0.34 ∼ 1 0.19 0.82 0.79 0.74 1 2.8 1.9 1.4 Table 3: The expectation values for the maximum mutual inclination and eccentricity of planets in a given system for two different assumptions for the system size Rsys. The cluster size limits are taken to be between N0 and N1 for the integration of equation (5). N0 N1 (cid:104)i(cid:105) (deg,rad) 100 AU (cid:104)e(cid:105) (cid:104)i(cid:105) (deg,rad) 30 AU (cid:104)e(cid:105) 10 30 100 104 104 104 5.4, 0.094 3.5, 0.061 1.9, 0.033 0.083 0.056 0.042 1.8, 0.031 1.0, 0.017 0.49, 0.0086 0.043 0.035 0.029 -- 39 -- REFERENCES Adams, F. C. & Laughlin, G. 2001, Icarus, 150, 151 Adams, F. C., Proszkow, E. M., Fatuzzo, M., & Myers, P. C. 2006, ApJ, 641, 504 Andr´e, P. 2002, ApSS, 281, 51 Barranco, J. A. & Goodman, A. A. 1998, ApJ, 504, 207 Baruteau, C., Meru, F., & Paardekooper, S.-J. 2011, MNRAS, 416, 1971 Bastian, N., Gieles, M., Goodwin, S. P., Trancho, G., Smith, L. J., Konstantopoulos, I., & Efremov, Y. 2008, MNRAS, 389, 223 Bastian, N. & Goodwin, S. P. 2006, MNRAS, 369, L9 Bitsch, B. & Kley, W. 2010, A&A, 523, A30+ -- . 2011, A&A, 530, A41+ Boley, A. C. 2009, ApJ, 695, L53 Boley, A. C., Hayfield, T., Mayer, L., & Durisen, R. H. 2010, Icarus, 207, 509 Boley, A. C., Helled, R., & Payne, M. J. 2011, ApJ, 735, 30 Borucki, W. J., Koch, D. G., Basri, G., Batalha, N., Brown, T. M., Bryson, S. T., Cald- well, D., Christensen-Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Gautier, III, T. N., Geary, J. C., Gilliland, R., Gould, A., Howell, S. B., Jenkins, J. M., Latham, D. W., Lissauer, J. J., Marcy, G. W., Rowe, J., Sasselov, D., Boss, A., Charbonneau, D., Ciardi, D., Doyle, L., Dupree, A. K., Ford, E. B., Fortney, J., Holman, M. J., Seager, S., Steffen, J. H., Tarter, J., Welsh, W. F., Allen, C., Buchhave, L. A., Christiansen, J. L., Clarke, B. D., Das, S., D´esert, J.-M., Endl, M., Fabrycky, D., Fressin, F., Haas, M., Horch, E., Howard, A., Isaacson, H., Kjeldsen, H., Kolodziejczak, J., Kulesa, C., Li, J., Lucas, P. W., Machalek, P., McCarthy, D., MacQueen, P., Meibom, S., Miquel, T., Prsa, A., Quinn, S. N., Quintana, E. V., Ragozzine, D., Sherry, W., Shporer, A., Tenenbaum, P., Torres, G., Twicken, J. D., Van Cleve, J., Walkowicz, L., Witteborn, F. C., & Still, M. 2011, ApJ, 736, 19 Boss, A. P. 1997, Science, 276, 1836 Caselli, P., Benson, P. J., Myers, P. C., & Tafalla, M. 2002, ApJ, 572, 238 -- 40 -- U A 0 0 3 > q l l A U A 0 0 3 > q y t i c i r t n e c c E l a u t u M e n a l P ) g e d ( l a n F i y − x l a u t u M e n a l P ◦ 3 . 0 > ) g e d ( l a n F i y − x l a u t u M e n a l P ) g e d ( l a n F i y − x l a u t u M e n a l P ) g e d ( C I y − x t a d e t a r a p e s e r a s k a e p o w t e h T . ) 5 - 4 . g i F e e s ( g n i r e t t a c s o t e u d k a e p n o i t a n i - h g i h , d a o r b a d n a m e t s y s e h t t e n a l p - t e n a l P . s n o i t a l u m i s e h t n i s e t a t s l a n fi d n a l a i t i n i e h t r o f s t e n a l p d n u o b l l a f o s n o i t a n i l c n i n a i d e M : 4 e l b a T l a i t i n i e h t s t c e fl e r t a h t k a e p n o i t a n i l c n i - w o l a h t i w , s n o i t a n i l c n i n i n o i t u b i r t s i d k a e p - o w t a o t s d a e l e n o l a g n i r e t t a c s i g n b r u t r e p f o e c n e s e r p t a s t e n a l p l l a e s u a c e b ◦ e h T . s n o i t u b i r t s i d n o i t a n i l c n i l a u t u m e h t d n a e n a l p e h t o t e v i t a l e r e h t r o f s e e r g e d 5 0 . 0 n a h t s s e l e r a s n o i t a n i l c n i n a i d e m l a i t i n i e h T . n o i t u b i r t s i d k a e p - o w t s i h t t u o s r a e m s s r a t s . n e v i g o s l a e r a s e i t i c i r t n e c c e n a i d e m e h t , s n m u l o c o w t t s a l e h t n I . n o i t a n i l c n i l a i t i n i o r e z h t i w t e s e r a U A 0 0 1 i l c n y − x 3 . 0 ∼ f o e t a t s i - - 7 1 0 . 0 0 4 0 . 0 - 5 1 0 . 0 8 3 0 . 0 9 1 0 . 0 7 4 0 . 0 9 4 0 . 0 - - 9 1 . 0 5 4 . 0 - - - 3 1 . 0 3 3 . 0 - 5 . 3 1 1 - - - i 4 . 2 9 . 5 - - - 0 8 0 . 0 2 1 . 0 4 2 . 0 6 8 . 0 6 9 . 0 8 3 0 . 0 5 7 0 . 0 8 1 . 0 5 6 . 0 7 6 . 0 3 7 0 . 0 5 8 0 . 0 4 7 0 . 0 7 8 0 . 0 6 8 0 . 0 4 2 0 . 0 4 3 0 . 0 5 2 0 . 0 4 3 0 . 0 3 3 0 . 0 0 F 1 P 2 0 F 1 P 3 1 F 1 P 2 1 F 1 P 3 C 1 F 1 P 3 e m a N -- 41 -- Fig. 16. -- Cumulative histograms of the integrator energy conservation for the simulations. The median energy error for all simulations is smaller than dE/E ∼ 10−7, and almost all systems have energy conservation smaller than 10−5. There are seven systems in the simulations with flybys (2P1F1 and 3P1F1 combined) that have energy conservation worse than 10−4, but they do not change the general results. To make sure that systems with very poor energy conservation are not biased toward systems of interest, we list the median energy errors for all systems that have a planet with q < 0.1 AU at some time during the system's evolution, which are 2 × 10−7, 1 × 10−6, 7 × 10−7, and 1 × 10−6 for 2P1F0, 2P1F1, 3P1F0, 3P1F1, respectively. 10-1010-910-810-710-610-510-4dE/E0.00.20.40.60.81.0Cumulative FractionIntegrator Energy Error2P1F02P1F13P1F03P1F1 -- 42 -- Chambers, J. E. 1999, MNRAS, 304, 793 Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686, 580 Crida, A., Masset, F., & Morbidelli, A. 2009, ApJ, 705, L148 Dodson-Robinson, S. E., Veras, D., Ford, E. B., & Beichman, C. A. 2009, ApJ, 707, 79 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Goldreich, P. & Tremaine, S. 1980, ApJ, 241, 425 Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, ApJ, 406, 528 Heggie, D. C. & Rasio, F. A. 1996, MNRAS, 282, 1064 Holman, M., Touma, J., & Tremaine, S. 1997, Nature, 386, 254 Ivanov, P. B. & Papaloizou, J. C. B. 2004, MNRAS, 347, 437 Juri´c, M. & Tremaine, S. 2008, ApJ, 686, 603 Kalas, P., Graham, J. R., Chiang, E., Fitzgerald, M. P., Clampin, M., Kite, E. S., Stapelfeldt, K., Marois, C., & Krist, J. 2008, Science, 322, 1345 Kley, W., Bitsch, B., & Klahr, H. 2009, A&A, 506, 971 Kley, W. & Nelson, R. P. 2012, ArXiv e-prints Kratter, K. M., Murray-Clay, R. A., & Youdin, A. N. 2010, ApJ, 710, 1375 Lada, C. J. 2006, ApJ, 640, L63 Lada, C. J. & Lada, E. A. 2003, ARAA, 41, 57 Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk, M. H. 2010, ApJ, 719, 497 Lin, D. N. C. & Ida, S. 1997, ApJ, 477, 781 Lissauer, J. J., Fabrycky, D. C., Ford, E. B., Borucki, W. J., Fressin, F., Marcy, G. W., Orosz, J. A., Rowe, J. F., Torres, G., Welsh, W. F., Batalha, N. M., Bryson, S. T., Buchhave, L. A., Caldwell, D. A., Carter, J. A., Charbonneau, D., Christiansen, J. L., Cochran, W. D., Desert, J.-M., Dunham, E. W., Fanelli, M. N., Fortney, J. J., Gautier, III, T. N., Geary, J. C., Gilliland, R. L., Haas, M. R., Hall, J. R., Holman, M. J., Koch, D. G., Latham, D. W., Lopez, E., McCauliff, S., Miller, N., Morehead, R. C., Quintana, E. V., Ragozzine, D., Sasselov, D., Short, D. R., & Steffen, J. H. 2011a, Nature, 470, 53 -- 43 -- Lissauer, J. J., Ragozzine, D., Fabrycky, D. C., Steffen, J. H., Ford, E. B., Jenkins, J. M., Shporer, A., Holman, M. J., Rowe, J. F., Quintana, E. V., Batalha, N. M., Borucki, W. J., Bryson, S. T., Caldwell, D. A., Carter, J. A., Ciardi, D., Dunham, E. W., Fortney, J. J., Gautier, III, T. N., Howell, S., Koch, D. G., Latham, D. W., Marcy, G. W., Morehead, R. C., & Sasselov, D. 2011b, ArXiv e-prints Malmberg, D., Davies, M. B., & Heggie, D. C. 2011, MNRAS, 411, 859 Mamajek, E. E. & Meyer, M. R. 2007, ApJ, 668, L175 Marcy, G. W., Butler, R. P., Williams, E., Bildsten, L., Graham, J. R., Ghez, A. M., & Jernigan, J. G. 1997, ApJ, 481, 926 Mardling, R. A. 2007, MNRAS, 382, 1768 Marois, C., Macintosh, B., Barman, T., Zuckerman, B., Song, I., Patience, J., Lafreni`ere, D., & Doyon, R. 2008, Science, 322, 1348 Michael, S., Durisen, R. H., & Boley, A. C. 2011, ApJ, 737, L42+ Miller-Ricci, E., Meyer, M. R., Seager, S., & Elkins-Tanton, L. 2009, ApJ, 704, 770 Moorhead, A. V. & Adams, F. C. 2005, Icarus, 178, 517 -- . 2008, Icarus, 193, 475 Myers, P. C. & Benson, P. J. 1983, ApJ, 266, 309 Nagasawa, M., Ida, S., & Bessho, T. 2008, ApJ, 678, 498 Naoz, S., Farr, W. M., Lithwick, Y., Rasio, F. A., & Teyssandier, J. 2011, Nature, 473, 187 Nayakshin, S. 2010, MNRAS, 408, L36 Ogilvie, G. I. & Lubow, S. H. 2003, ApJ, 587, 398 Paardekooper, S.-J. & Mellema, G. 2006, A&A, 459, L17 Press, W. H. 2002, Numerical recipes in C++ : the art of scientific computing, ed. Press, W. H. Proszkow, E.-M. & Adams, F. C. 2009, ApJS, 185, 486 Raghavan, D., McAlister, H. A., Henry, T. J., Latham, D. W., Marcy, G. W., Mason, B. D., Gies, D. R., White, R. J., & ten Brummelaar, T. A. 2010, ApJS, 190, 1 -- 44 -- Rasio, F. A. & Ford, E. B. 1996, Science, 274, 954 Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARA&A, 25, 23 Siegler, N., Close, L. M., Cruz, K. L., Mart´ın, E. L., & Reid, I. N. 2005, ApJ, 621, 1023 Snellgrove, M. D., Papaloizou, J. C. B., & Nelson, R. P. 2001, A&A, 374, 1092 Spiegel, D. S. & Burrows, A. 2011, ArXiv e-prints Sumi, T., Kamiya, K., Bennett, D. P., Bond, I. A., Abe, F., Botzler, C. S., Fukui, A., Furusawa, K., Hearnshaw, J. B., Itow, Y., Kilmartin, P. M., Korpela, A., Lin, W., Ling, C. H., Masuda, K., Matsubara, Y., Miyake, N., Motomura, M., Muraki, Y., Nagaya, M., Nakamura, S., Ohnishi, K., Okumura, T., Perrott, Y. C., Rattenbury, N., Saito, T., Sako, T., Sullivan, D. J., Sweatman, W. L., Tristram, P. J., Udal- ski, A., Szyma´nski, M. K., Kubiak, M., Pietrzy´nski, G., Poleski, R., Soszy´nski, I., Wyrzykowski, (cid:32)L., Ulaczyk, K., & Microlensing Observations in Astrophysics (MOA) Collaboration. 2011, Nature, 473, 349 Triaud, A. H. M. J., Collier Cameron, A., Queloz, D., Anderson, D. R., Gillon, M., Hebb, L., Hellier, C., Loeillet, B., Maxted, P. F. L., Mayor, M., Pepe, F., Pollacco, D., S´egransan, D., Smalley, B., Udry, S., West, R. G., & Wheatley, P. J. 2010, A&A, 524, A25+ Veras, D., Crepp, J. R., & Ford, E. B. 2009, ApJ, 696, 1600 Veras, D. & Raymond, S. N. 2012, MNRAS, 421, L117 Walsh, K. J., Morbidelli, A., Raymond, S. N., O'Brien, D. P., & Mandell, A. M. 2011, Nature, 475, 206 Wright, J. T., Fakhouri, O., Marcy, G. W., Han, E., Feng, Y., Johnson, J. A., Howard, A. W., Fischer, D. A., Valenti, J. A., Anderson, J., & Piskunov, N. 2011, PASP, 123, 412 Wyatt, M. C. & Dent, W. R. F. 2002, MNRAS, 334, 589 Zakamska, N. L. & Tremaine, S. 2004, AJ, 128, 869 Zhu, Z., Hartmann, L., Nelson, R. P., & Gammie, C. F. 2012, ApJ, 746, 110 This preprint was prepared with the AAS LATEX macros v5.2. -- 45 -- Table 5: Histogram data for the distribution RJ (left) and J (right) in low-mass cloud cores, based on the Caselli et al. (2002) observations. The data are for the 20 sources (their Table 5) that have enough information to estimate RJ , the angular momentum barrier for a rotating, collapsing cloud. Note that bin values for the J distribution will not correspond to the RJ bin because conversion from J to RJ requires the mass of the cloud. Here F40AUbin and F0.4dex are the fraction of sources that fall within the given bin. The final row in the table shows the average of all the bins weighted by F40AUbin. No weighing for the angular momentum distribution is given. 0.6 0.1 140 100 0.2 20 60 RJ (AU) F40 AU bin (AU) N0 10 30 100 10 30 100 10 30 100 10 30 100 10 30 100 10 30 100 Weighted Average 0.05 180 0.05 log10(J(cm2s−2)) F0.4dex 19.9 20.3 20.7 21.1 21.5 0.05 0.2 0.35 0.35 0.05 (cid:104)imedian(cid:105) 1.2 0.68 0.33 3.5 2.1 1.0 5.4 3.5 1.9 7.1 4.8 2.8 8.5 6.1 3.7 2.9 1.9 1.0 (cid:104)emeidan(cid:105) 0.037 0.031 0.028 0.061 0.045 0.034 0.083 0.060 0.042 0.010 0.075 0.051 0.012 0.089 0.061 0.056 0.15 0.034 fsurvive 0.67 0.68 0.69 0.61 0.65 0.68 0.55 0.61 0.66 0.50 0.57 0.63 0.47 0.53 0.60 0.62 0.65 0.68
1303.1796
1
1303
2013-03-07T20:01:22
The orbit of the Chelyabinsk event impactor as reconstructed from amateur and public footage
[ "astro-ph.EP" ]
A ballistic reconstruction of a meteoroid orbit can be made if enough information is available about its trajectory inside the atmosphere. A few methods have been devised in the past and used in several cases to trace back the origin of small impactors. On February 15, 2013, a medium-sized meteoroid hit the atmosphere in the Chelyabinsk region of Russia, causing damage in several large cities. The incident, the largest registered since the Tunguska event, was witnessed by many thousands and recorded by hundreds of amateur and public video recording systems. The amount and quality of the information gathered by those systems is sufficient to attempt a reconstruction of the trajectory of the impactor body in the atmosphere, and from this the orbit of the body with respect to the Sun. Using amateur and public footage taken in four different places close to the event, we have determined precisely the properties of the entrance trajectory and the orbit of the Chelyabinsk event impactor. We found that the object entered the atmosphere at a velocity ranging from 16.0 to 17.4 km/s in a grazing trajectory, almost directly from the east, with an azimuth of velocity vector of 285$^o$, and with an elevation of 15.8$^o$ with respect to the local horizon. The orbit that best fits the observations has, at a 95% confidence level, a semi-major axis a = 1.26$\pm$0.05 AU, eccentricity e = 0.44$\pm$0.03, argument of perihelion $\omega$=95.5$^o\pm2^o$ and longitude of ascending node $\Omega$= 326.5$^o\pm0.3^o$. Using these properties the object can be classified as belonging to the Apollo family of asteroids. The absolute magnitude of the meteoroid was H= 25.8, well below the threshold for its detection and identification as a Potential Hazardous Asteroid (PHA). This result would imply that present efforts intended to detect and characterize PHAs are incomplete.
astro-ph.EP
astro-ph
The orbit of the Chelyabinsk event impactor as reconstructed from amateur and public footage Jorge I. Zuluagaa1, Ignacio Ferrína, Stefan Geensb aInstituto de Física – FCEN, Universidad de Antioquia, Clle. 67 No. 53-108, Medellín, Colombia bOgle Earth, c/o Alpen, Regeringsgatan 87, Stockholm, Sweden Abstract A ballistic reconstruction of a meteoroid orbit can be made if enough information is available about its trajectory inside the atmosphere. A few methods have been devised in the past and used in several cases to trace back the origin of small impactors. On February 15, 2013, a medium-sized meteoroid hit the atmosphere in the Chelyabinsk region of Russia, causing damage in several large cities. The incident, the largest registered since the Tunguska event, was witnessed by many thousands and recorded by hundreds of amateur and public video recording systems. The amount and quality of the information gathered by those systems is sufficient to attempt a reconstruction of the trajectory of the impactor body in the atmosphere, and from this the orbit of the body with respect to the Sun. Using amateur and public footage taken in four different places close to the event, we have determined precisely the properties of the entrance trajectory and the orbit of the Chelyabinsk event impactor. We found that the object entered the atmosphere at a velocity ranging from 16.0 to 17.4 km/s in a grazing trajectory, almost directly from the east, with an azimuth of velocity vector of 285º, and with an elevation of 15.8º with respect to the local horizon. The orbit that best fits the observations has, at a 95% confidence level, a semi-major axis a = 1.26±0.05 AU, eccentricity e = 0.44±0.03, argument of perihelion ω = 95.5º±2º and longitude of ascending node Ω = 326.5º±0.3º. Using these properties the object can be classified as belonging to the Apollo family of Near Earth Asteroids. The absolute magnitude of the meteoroid was H= 25.8, well below the threshold for its detection and identification as a Potential Hazardous Asteroid (PHA). This result would imply that present efforts intended to detect and characterize PHAs are incomplete and are missing approximately half the objects able to impact our planet and cause local damage. Keywords: Meteors; Asteroids, Orbit; Asteroids, Impact 1. Introduction There is an increasing level of awareness about the potential risks posed by the impact of medium- sized to large Near Earth Asteroids (NEAs). Global efforts have been made to detect, characterize and predict the trajectories of the several thousand asteroids that have a non-negligible probability of                                                                                                                 impacting our planet and causing damage to highly populated areas. And yet there remain unpredictable events such as the one witnessed in the Chelyabinsk region of Russia, where a small Abreviations: HA, Harmful Asteroids; ChEA, Chelyabinsk Event Asteroid   Corresponding author at: Instituto de Física – FCEN, Universidad de Antioquia, Clle. 67 No. 53-108, Medellín, Colombia. Email addresses: [email protected] (Jorge I. Zuluaga), [email protected] (Ignacio Ferrin), [email protected] (Stefan Geens) asteroid with a diameter estimated at 17-20 m hit the atmosphere, producing a super bolide and an explosion with an energy of around 440 kton (Yeomans & Chodas, 2013). Figure 1. Footage from two of the vantages points used in this work: the Revolution Square in Chelyabinsk (left panel and Privokzalnaya Square in the same city. Lines show the direction and length of the shadow at different times in the video. The numbers indicate an arbitrary label for the snapshots where the shadows where measured. The Chelyabinsk event happened at 09:20 local YEKT time (UTC+6 hours), just after sunrise and at a time when over a million people were awake and potential witnesses to the display. Hundreds of amateur cameras, mobile phones, security and public cameras and even professional recording devices registered the development of the atmospheric event from its first appearance up to the powerful explosion and fragmentation of the parent body at a relatively low altitude. This event, the largest recorded since the Tunguska impact (Andreev, 1990; Sekanina, 1998; Gasperini, et al., 2007) has been historic in several aspects. First, it happened over a region enclosing populated areas (there are four cities near the impact region with a combined population of 1.5 million). Second, it happened close to a peak commuting time, when many people were outdoors. Third, it happened in an era when devices capable of recording video are cheap and ubiquitous, allowing pervasive observations from many different local vantage points. And finally, it happened in the information age, when anybody is able to share almost instantaneously their experiences through social networks. All these conditions ensured that this unexpected but historic natural event would become one of the best documented ever. The reconstruction of the impactor trajectory in the case of super bolides has become customary (see e.g. Balabh et al. 1978, Trigo-Rodríguez et al. 2010). Several observing networks have been installed around the world to detect and facilitate the ballistic reconstruction of the parent body (see e.g. Caplecha, 1987; Trigo-Rodríguez et al. 2006). Citizen has been also involved in these efforts in the past (Huziak & Sarty, 1994). Despite these efforts, only several orbits have been successfully determined, mainly due to the lack of enough triangulation information (see e.g. Trigo-Rodríguez et al. 2007). Although no known meteor network registered the Chelyabinsk super bolide, a large number of amateur and public video recordings acted as an ad-hoc sensor network, able to potentially provide information for a successful reconstruction of the trajectory through the atmosphere. Challenges in using this method include the relative scarcity of this ad-hoc network, the variable quality of the footage, and the lack of synchronization between the time recording systems. However, the pervasive presence of GPS devices and free public access to accurately positioned satellite imagery within the context of virtual globes such as Google Earth allow the pinpointing of the precise geographic locations of these recording devices. 2. Methods After checking almost a hundred amateur videos and recordings taken by security and public cameras, we selected four videos for their high quality and for the availability of precise location information, both elements essential to a successful triangulation of the impactor trajectory. The videos were recorded at (1) Revolution Square in downtown Chelyabinsk (55.16080°N, 61.40249°E), (2) Privokzalnaya square just 2 km to the south (55.14392°N, 61.41421°E), (3) the marketplace of Korkino, a city 30 km to the south of Chelyabinsk (54.89093°N, 61.39956°E) and (4) the Central Square of Kamensk-Uralsky (56.41500°N, 61.91858°E), a city 150 km to the north of the impact. For additional information about the vantage points, see Appenix A. The base of the triangulation covers a longitudinal distance of around 170 km, which is of the same order of magnitude as the length of the impactor trajectory. The highest quality videos were analyzed frame by frame during the 5-second duration of the brightest part of the event. Using a method originally devised by one of us (S. Geens) we measured the shadows cast by objects in the scene and estimated from them the elevation and azimuth relative to reference directions of the fireball as a function of time (see figure 1). This method was used in three of the four recordings (Revolution Square, Privokzalnaya Square and Korkino), where cameras were pointing to the floor and no images of the fireball in the sky were available (in the Korkino market place a brief appearance of the bolide is available but at very low quality). In the last recording (Kamensk-Uralsky central square) direct images of the bolide were analyzed, providing additional information about the trajectory. The azimuths of reference directions, distances and sizes of the familiar objects in the footage were obtained from information provided by Google Earth and some archived images of the places used in the triangulation. In order to accurately measure lengths from a two dimensional image, we have corrected for perspective (see perspective grid used in figure 1). Since no information is presently available about the physical properties of the recording devices, no corrections for optical deformations were applied to the measurements. Figure 1 shows snapshots of the measurements performed on the footage at two of the vantage points used in this reconstruction. We assign errors to azimuths and elevations according to the uncertainty of the determination of, for example, the tip of the shadow or the center of the bolide image in the Kamensk-Uralsky footage. Accordingly, the assumption that the errors for these quantities are normal is reasonable. One of the trickiest parts of a triangulation procedure is the precise synchronization of the observations performed from different vantage points (see e.g. Borovicka, 1990; Green, 2010). Only two of the videos (those at Revolution Square and Privokzalnaya Square) had proper synchronized time stamps. Those sites were however the closest together and a synchronized triangulation of the trajectory using these observations is affected by uncertainties. In order to avoid the requirement of synchronization, we devised an asynchronous method we called the altazimuth-footprint method. In the method, as many observed elevations and azimuths as possible are measured at a given vantage point and compared with the altazimuth footprint of a test trajectory calculated with one of our orbits (see figure 2). For each observation point and at each vantage location we compute the elevation expected for a given test trajectory at the measured azimuth Ao of that point. The expected elevations ht were then compared with the observed ones ho. For that purpose we computed the least- [ℎ!"#−ℎ! 𝐴!" ; 𝕋, 𝕍! ]! !! !! 𝜒 ! (𝕋) =   square statistics: Δℎ!"# !!! !!! Here the index j runs over the nv = 4 vantage points (represented by 𝕍! ) and the index i over the nj jth vantage point. 𝕋 represents the test trajectory. observations performed at that j-th vantage point. Δhoji is the error of the ith observed elevation at the Figure 2. The altazimuth-footprint triangulation method. Every observer’s location has an elevation vs. altitude signature (dots with error bars). Three signatures pinpoint the orbit very precisely. Notice that time is not involved in this plot. The solid lines are theoretical determinations using the best-fit orbit obtained in this work (ZFG2013). We have also included the theoretical prediction by the orbit solution provided by the team of the Astronomical Institute of the Czech Academy of Sciences (BSS2013, Borovicka et al. 2013) and NASA (YC2013, Yeomans & Chodas, 20013). Our reconstruction is very satisfactory. Each test trajectory was characterized by 5 parameters: longitude, latitude, altitude, azimuth and elevation as measured in the projected impact point. The best-fit trajectory was calculated by minimizing the least-square statistics with respect to the previous 5 parameters. Given the fact that the measured elevations are derived from a measuring procedure affected by normal errors, and since the number of observations was statistically significant (38 points), we assume that the statistics are distributed according to a Chi-square distribution with 33 degrees of freedom. Using this fact, we compute for each trajectory the one-tailed p-value and use it to compute minimum and maximum limits for the trajectory parameters. The altazimuth-footprint method presented here is trivially applicable to an arbitrary number of locations. Best-fit 59.8703E 55.0958N 219.0 m 105.0° 15.8° 324.3° 4.73° 16.7 km/s 68.3 km Maximum -0.043E -0.19N -0.4 m -1.7° -0.32° -1.51° -1.12° -0.68 km/s -3.30 km Minimum +0.051E +0.15N +0.4 m +2.2° +0.27° +1.66° +1.18° +0.65 km/s +3.62 km Projected Impact Site Longitude Latitude Altitude Azimuth* Elevation* Right Ascension Declination Velocity Height * Bolide radiant ** Difference of maximum and minimum with respect to best-fit trajectory in a sample with n=153 trajectories with p-value 0.05<p<0.95. Table 1. Properties of the best-fit trajectory for this work. The trajectory is assumed rectilinear and having a projected point on the ground located at the given latitude and longitude. Velocity was estimated from the Kamensk Uralsky vantage point using the highest observed point in the trajectory. It should be stressed that elevations and azimuths were computed properly with respect to the average geoid, with equatorial radius RE = 6378.2064 and flattening f = 1/294.9787 (Acton, 1996). All the calculations involving reference systems and ephemerides were performed using the NOVAS Package, developed by the U.S. Naval Observatory (Bangert et al. 2011) and the SPICE Toolkit (Acton, 1996). 3. Results After applying the altazimuth-footprint method to our set of vantage points and observations, we obtained the properties of the best-fit trajectory that we summarize in Table 1. According to our triangulation the projected impact site was not even close to Lake Chebarkul, as is assumed in previous reconstruction attempts (Zuluaga & Ferrin, 2013). This implies that if a piece of the parent body is found in the lake it is a fragment that separated during entry. The projected impact site misses by several kilometers the city of Miass, 83 km to the west of Chelyabinsk. In order to reconstruct the heliocentric orbit of the impactor we took as the orbit reference point the highest observed point in the trajectory as registered from the farthest vantage point, i.e. the central square of Kamensk-Uralsky. This particular point in the trajectory is suitable because: 1) It is high enough in the atmosphere (71 km in our best fit trajectory) to ensure that the estimated speed is the least affected by atmospheric drag and 2) the speed can be reliably estimated by direct observations of the fireball performed at this vantage point. Using the available information we estimate that the velocity of the impactor at that height was 16.0-17.4 km/s. This is almost 1-2 km/s below NASA estimate (Yeomans & Choda et al. 2013) and coincides in the upper end with the estimation by the Czech group (Borovicka et al. 2013). We transform the geodetic state vector at the orbit reference point (position expressed in latitude, longitude and altitude and velocity expressed in terms of speed, azimuth and elevation), first to cartesian planetocentric coordinates in the rotating International Terrestrial Reference System (ITRS), and from there to the Geocentric Celestial Reference System (GCRS). Using the JPL ephemeris for the Earth at the exact time of the impact, we finally obtain the precise state vector of the impactor as referred to the barycenter of the solar system. Although a direct conversion of the state vector to classical elements as referred to the Sun can be performed at this point, we preferred to numerically integrate the trajectory of the body backwards up to 4 years before the impact. For that purpose we used the Mercury integrator (Chambers, 2008), setting precisely the position of the major Solar System planets. We did this to know if small planetary or lunar perturbations could have altered the trajectory previous to the body impact. No significant perturbations were found in this time frame. The orbital elements reported in Property Best-fit orbit Min. Max. 1.34 0.47 0.72 1.97 3.46 96.8 326.8 - Standard Deviation 0.033 0.018 0.005 0.007 0.19 0.76 0.08 - 1.266 AU A 0.435 e 0.716 AU q 1.816 AU Q 2.984° i 95.08° ω 326.54° Ω M* 106.85° All elements are referred to J2000.0 * Measured at the epoch of impact, jd = 2456338.639279. Table 2 Elements of the best-fit heliocentric orbit. Standard deviation, minimum and maximum were calculated for a sample of n=50 orbits having p-values of the least-square statistics 0.05<p<0.95. 1.21 0.40 0.71 1.71 2.76 93.4 326.5 - In table 2 we present the osculant elements measured at 0.01 years before impact (Earth-body distance, Δ > 0.05 AU). In figure 3 we show the orbital elements a and e as compared with that of three NEA families, the Amors, the Apollos and the Athens. We see that the parent asteroid of the Chelyabinsk event can be classified unambiguously as an Apollo asteroid, confirming preliminary attempts (Zuluaga & Ferrin, 2013). It is also interesting to notice that Tunguska impactor may have been an Apollo (Andreev, 1990). The two impactors have suspisciously similar atmospheric trajectories (impact time, azimuth, elevation and geographic location) (Sekanina, 1998) and hence they may be related. However any detailed analysis of this particular issue is beyond the scope of this paper. Using the best-fit orbit, we then compute the observational parameters of the parent body a few days before impact (see figure 4). To estimate the brightness of the we perform a backward integration of the orbit with a small step size (6 hours). We assumed a geometric albedo pV = 0.28± 0.13 which is the mean value of 9 Apollo asteroids (Veeder, 1989), and a diameter of 18 m (Yeomans & Chodas, 2013). Three phase laws were selected (see e.g. Delahod et al. 2001). These values imply that the object had an absolute magnitude H = 25.8. Only a few hours before impact does the apparent magnitude rise to levels detectable by current surveys. However, its angular distance to the sun is lower than the search limit of these same surveys. Having an absolute magnitude H>22, the object has not been classified as a Potentially Hazardous Asteroid, PHA. The object should actually be classified as an HA or a Harmful Asteroid, of which it is the second member after the Tunguska event impactor. Since the current number of officially known PHAs, i.e. Δmin<0.05 AU and H<22, is 1381, an extrapolation of the absolute magnitude cumulative distribution from H=22 to H=25.8 gives an estimate of 1300±50 additional potentially hazardous asteroids. These objects are not capable of global damage but are still capable of local damage. With this result the number of PHAs larger than the Chelyabinsk event impactor is raised from the current 1381 to 2680±70 bodies, almost double the previous value. Figure 3. Orbital classification of the Chelyabinsk event impactor. Big red point indicates the position of the best-fit orbit elements. Black diamonds show n=50 orbits with p-value of the altazimuth-footprint least square 0.05<p<0.95. The object lies in the region of the Apollo asteroids and has a Tisserand invariant of 2.8. Figure 4. Calculation of the asteroid brightness in the days before impact. Just few hours before the impact the parent body could be potentially observed by NEA surveys. However since its angular distance to the sun was probably less than 15o it was out of the scope of the present visual surveys. 5. Conclussions There are several lessons we can learn from this work. (1) To confirm the fundamental role that active enthusiasts, a.k.a. citizen astronomers, can play in scientific research, especially in cases when unexpected events occur. (2) We are missing half of the PHAs by our own definition of what a PHA is. (3) Although objects smaller than 100 m cannot produce global damage they can still produce significant local damage. (4) A simple calculation shows that if the impactor had been delayed by about 3.5-4 minutes, the impact would have taken place over central Europe, were the damage could have been much greater. (5) The object approached Earth from the Sun side, which is not covered by current optical surveys. This side is totally unshielded. Acknowledgements The data contributing to the successful reconstruction of the orbit were taken by people in Chelyabinsk, Korkino and Kamensk-Uralsky. We thank these contributors for submitting their footage to public access. We also want to thank those citizen scientists who alerted us to useful videos or pinpointed their location information in the comments to the original blog post on Ogle Earth (http://ogleearth.com) in the days after the event. Special thanks to SebastienP, liilliil, Sean Mac, Robin Whittle, Kuuuurija, Serge, Latuha Valeryi, ssvilponis, Dmitry DD, comeT, Sirius, g1smd, Steve and Gary (Online handles are self-reported, information was independently verified). J.Z. and I.F. acknowledge support from the CODI/UdeA. Appendix A. The videos All the footage used in this work is publicly available from videos on YouTube (verified March 4, 2013). The location of the vantage points where the videos were recorded and the YouTube URL for each of them are listed in the table below: Vantage point Revolution Square (Chelyabinsk) Altitude URL https://www.youtube.com/v/bXifSi2K278 230 m 238 m https://www.youtube.com/v/Qin41lP9r2U https://www.youtube.com/v/odKjwrjIM-k Privokzalnaya Square (Chelyabinsk) Market Place (Korkino) Central Square (Kamensk Uralsky) Table A1. Location of the vantages points and link to the videos used in the reconstruction. The authors have preserved backup copies of the videos, in case they are removed from YouTube. No public access to these backups is available. In the event the videos are no longer available on YouTube, researchers interested in the material should contact [email protected] for a copy. A complete Google Earth KMZ file showing the trajectory and referencing the vantage points used in the reconstruction. It can be download from this URL: https://www.youtube.com/v/iCawTYPtehk Longitud, Latitude 61.40249°E, 55.16080°N 61.41421°E, 55.14392°N 61.39956°E, 54.89093°N 61.91858°E, 56.41500°N 241 m 169 m http://urania.udea.edu.co/sitios/facom/pages/chelyabinsk-meteoroid.rs/files/chelyabinsk- meteoroidfkpdk/ZFG2013-TrajectoryVantagePointsVideos.kmz The positions of other locations with links to videos recorded at those locations are referenced in the same file. Other supplementary information and updates are available here: http://astronomia.udea.edu.co/chelyabinsk-meteoroid. References Acton Jr, C. H. Ancillary data services of NASA’s navigation and ancillary information facility. Planetary and Space Science 44(1), 65–70 (1996). Andreev, G. V. (1990). Was the Tunguska 1908 Event Caused by an Apollo Asteroid? In Asteroids, Comets, Meteors III (Vol. 1, p. 489). Ballabh, G. M., Bhatnagar, A., & Bhandari, N. (1978). The orbit of the Dhajala meteorite. Icarus, 33(2), 361-367. Borovicka, J. (1990). The comparison of two methods of determining meteor trajectories from photographs. Bulletin of the Astronomical Institutes of Czechoslovakia, 41, 391-396. Ceplecha, Z. (1987). Geometric, dynamic, orbital and photometric data on meteoroids from photographic fireball networks. Bulletin of the Astronomical Institutes of Czechoslovakia, 38, 222-234. Delahodde, C. E., Meech, K. J., Hainaut, O. R., Dotto E. Detailed phase function of comet 28P/Neujmin 1. A&A 376, 672-685 (2001) Chambers, J. E. (2008). A hybrid symplectic integrator that permits close encounters between massive bodies. Monthly Notices of the Royal Astronomical Society, 304(4), 793-799. Gasperini, L., Alvisi, F., Biasini, G., Bonatti, E., Longo, G., Pipan, M., & Serra, R. (2007). A possible impact crater for the 1908 Tunguska Event. Terra Nova, 19(4), 245-251. Green, W. (2010, May). Asynchronous Meteor Observations, Discovery, and Characterization of Shower Statistics and Meteorite Recovery. In Symposium on Telescope Science (p. 123). Huziak, R., & Sarty, G. (1994). The Value of Fireball Reports from the General Public-the 1993OCT30 Western Canada Fireball as a Case Study. Journal of the Royal Astronomical Society of Canada, 88, 332. Sekanina, Z. (1998). Evidence for asteroidal origin of the Tunguska object. Planetary and space science, 46(2), 191-204. Trigo-Rodriguez, J. M., Madiedo, J. M., Castro-Tirado, A. J., Ortiz, J. L., Llorca, J., Fabregat, J., ... & Gálvez, F. (2007). Spanish Meteor Network: 2006 continuous monitoring results. WGN, Journal of the International Meteor Organization, 35, 13-22. Trigo-Rodriguez, J. M., Madiedo, J. M., Castro-Tirado, A. J., Ortiz, J. L., Llorca, J., Fabregat, J., ... & Gálvez, F. (2007). Spanish Meteor Network: 2006 continuous monitoring results. WGN, Journal of the International Meteor Organization, 35, 13-22. Trigo‐Rodríguez, J. M., Borovička, J., Spurný, P., Ortiz, J. L., Docobo, J. A., Castro‐Tirado, A. J., & Llorca, J. (2010). The Villalbeto De La Peña Meteorite Fall: Ii. Determination Of Atmospheric Trajectory And Orbit. Meteoritics & Planetary Science, 41(4), 505-517. Veeder, G.J., Hanner, M.S., Matson, D.L., Tedesco, E.F., (1989). Radiometry of Near-Earth Asteroids. An. J., 97, 1211-1219. Unpublished and Web Material Borovicka, J., Spurny, P., and Shrbeny, L., trajectory and orbit of the chelyabinsk superbolide, Electronic Telegram Central Bureau for Astronomical Telegrams, IAU (2013), No. 3423. Bangert, J., W. Puatua, G. Kaplan, J. Bartlett, W. Harris, A. Fredericks, and A. Monet (2011). User’s guide to NOVAS, version c3.1. URL: http://aa.usno.navy.mil/software/novas/novas_info.php (2013) Yeomans, D & Chodas, P. Additional details on the large fireball event over Russia on Feb. 15, 2013. (Unpublished result). URL: http://neo.jpl.nasa.gov/news/fireball_130301.html (2013) Zuluaga, J.I., Ferrin, I. A preliminary reconstruction of the orbit of the Chelyabinsk Meteoroid (Unpublished Material), preprint at: http://arxiv.org/abs/1302.5377 (2013).
0903.1720
1
0903
2009-03-10T09:55:32
On the Stability of Elliptical Vortices in Accretion Discs
[ "astro-ph.EP" ]
(Abriged) The existence of large-scale and long-lived 2D vortices in accretion discs has been debated for more than a decade. They appear spontaneously in several 2D disc simulations and they are known to accelerate planetesimal formation through a dust trapping process. However, the issue of the stability of these structures to the imposition of 3D disturbances is still not fully understood, and it casts doubts on their long term survival. Aim: We present new results on the 3D stability of elliptical vortices embedded in accretion discs, based on a linear analysis and several non-linear simulations. Methods: We derive the linearised equations governing the 3D perturbations in the core of an elliptical vortex, and we show that they can be reduced to a Floquet problem. We solve this problem numerically in the astrophysical regime and we present several analytical limits for which the mechanism responsible for the instability can be explained. Finally, we compare the results of the linear analysis to some high resolution simulations. Results: We show that most anticyclonic vortices are unstable due to a resonance between the turnover time and the local epicyclic oscillation period. In addition, we demonstrate that a strong vertical stratification does not create any additional stable domain of aspect ratio, but it significantly reduces growth rates for relatively weak (and therefore elongated) vortices. Conclusions: Elliptical vortices are always unstable, whatever the horizontal or vertical aspect-ratio is. The instability can however be weak and is often found at small scales, making it difficult to detect in low-order finite-difference simulations.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. 1577src June 2, 2018 c(cid:13) ESO 2018 9 0 0 2 r a M 0 1 . ] P E h p - o r t s a [ 1 v 0 2 7 1 . 3 0 9 0 : v i X r a On the Stability of Elliptical Vortices in Accretion Discs Geoffroy Lesur and John C. B. Papaloizou Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Centre for Mathematical Sciences, Wilberforce Road, Cambridge CB3 0WA, UK e-mail: [email protected] Received date / Accepted date ABSTRACT Context. The existence of large-scale and long-lived 2D vortices in accretion discs has been debated for more than a decade. They appear spontaneously in several 2D disc simulations and they are known to accelerate planetesimal formation through a dust trapping process. In some cases, these vortices may even lead to an efficient way to transport angular momentum in protoplanetary discs when MHD instabilities are inoperative. However, the issue of the stability of these structures to the imposition of 3D disturbances is still not fully understood, and it casts doubts on their long term survival Aims. We present new results on the 3D stability of elliptical vortices embedded in accretion discs, based on a linear analysis and several non-linear simulations. Methods. We introduce a simple steady 2D vortex model which is a non-linear solution of the equations of motion, and we show that its core is made of elliptical streamlines. We then derive the linearised equations governing the 3D perturbations in the core of this vortex, and we show that they can be reduced to a Floquet problem. We solve this problem numerically in the astrophysical regime, including a simplified model to take into account vertical stratification effects. We present several analytical limits for which the mechanism responsible for instability can be explained. Finally, we compare the results of the linear analysis to some high resolution numerical simulations obtained with spectral and finite difference methods. A discussion is provided, emphasising the astrophysical consequences of our findings for the dynamics of vortices. Results. We show that most anticyclonic vortices are unstable due to a resonance between the turnover time and the local epicyclic oscillation period. A small linearly stable domain is found for vortex cores with an aspect-ratio of around 5. However, our simulations show that it is only the vortex core that is stable, with the instability still appearing on the vortex boundary. In addition, we find numerically that results obtained under the assumption of incompressibility are not affected by the introduction of a moderate compressibility. Finally, we show that a strong vertical stratification does not create any additional stable domain of aspect ratio, but it significantly reduces growth rates for relatively weak (and therefore elongated) vortices. Conclusions. Elliptical vortices are always unstable, whatever the horizontal or vertical aspect-ratio is. The instability can however be weak and is often found at small scales, making it difficult to detect in low-order finite-difference simulations. Key words. accretion, accretion disks -- instabilities -- hydrodynamics 1. Introduction The existence of 2D long-lived vortices in accretion discs was first proposed by von Weizsacker (1944) in an out- moded model of planet formation. This idea was re- vived by Barge & Sommeria (1995) to accelerate planetes- imals formation by a dust trapping process. This kind of vortex is often observed in 2D simulations of discs (see e.g. Godon & Livio 1999; Umurhan & Regev 2004; Johnson & Gammie 2005b; Bodo et al. 2007), since 2D tur- bulence is known to generate an inverse cascade of en- ergy leading to large 2D vortices (Onsager 1949). Vortices may also be generated by 2D instabilities such as Rossby wave instabilities (Lovelace et al. 1999) or baroclinic insta- bilities (Klahr & Bodenheimer 2003; Petersen et al. 2007), although the latter is still a matter of ongoing debate (Johnson & Gammie 2005a, 2006). These vortices may play at least two important roles regarding accretion disc dy- namics. First, they could lead to an efficient angular mo- mentum transport process in regions in which the magneto- rotational instability (Balbus & Hawley 1998) doesn't op- erate, such as in dead zones (Gammie 1996). Second, they are a very efficient way to accelerate the planetesimal formation process in protoplanetary discs (Johansen et al. 2004). However, the stability of these vortices when small 3D disturbances are imposed is largely unknown. In the astrophysics community, this issue has been in- vestigated mainly numerically. Shen et al. (2006) examined the formation of 2D vortices starting from 2D turbulence in fully compressible simulations. According to their results, a small 3D noise added to their initialy 2D configuration de- stroys the coherent vortices in a few orbits, relaxing the flow to its laminar state. Barranco & Marcus (2005) also com- puted the evolution of 3D vortices using an anelastic code incorporating vertical stratification. As Shen et al. (2006), they found that midplane vortices were destroyed by 3D perturbations. However, they also showed that off-midplane vortices could survive for several hundreds of orbits, leading to the possibility of a stabilizing effect due to the stratifi- cation 2 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices It is often assumed that these vortices are unstable because of the elliptical instability. The elliptical insta- bility is a parametric instability appearing when a mul- tiple of the vortex turnover frequency matches an iner- tial wave frequency, leading to a positive resonance. It is observed when the backgound flow follows closed stream- lines, and being localized on individual streamlines is a instability (in particular it doesn't need to involve local the vortex boundaries). This instability was first found numerically by Pierrehumbert (1986) and described using Craik & Criminale (1986) solutions by Bayly (1986) for pure elliptical flows. The rotating case was studied by Craik (1989), who showed that anticyclonic elliptical flows can be stable for some rotation rates. Interested readers may con- sult Kerswell (2002) for a more extensive discussion of the elliptical instability and its development in fluid mechanics. In the present paper, we investigate the elliptical in- stability in the context of accretion disc vortices. We first present a steady 2D vortex model, which is a non-linear solution of the local disc equations. We then present the linearised equations governing 3D perturbations inside the vortex. A criterion for the instability is derived from these equations and a physical understanding of the mechanism responsible for the instability is provided. We briefly ex- tend these results to a simplified stratified case, and we compare our findings to fully non-linear simulations of ac- cretion discs vortices. Finally, we provide a discussion and a comparison with previous work. 2. Local model of an elliptical vortex. 2.1. Shearing-sheet model In the following, we will assume a local model for the accre- tion disc, following the shearing-sheet approximation. The reader may consult Hawley et al. (1995), Balbus (2003) and Regev & Umurhan (2008) for an extensive discussion of the properties and limitations of this model. As a simplification, we will assume the flow is incompressible, consistently with the small shearing box model (Regev & Umurhan 2008). The shearing box equations are found by considering a Cartesian box centred at r = R0, rotating with the disc at angular velocity Ω = Ω(R0). We define R0φ → x and r − R0 → −y for consistency with the standard notation for plane Couette flows (e.g. Drazin & Reid 1981). Note that this definition differs from the standard notation used in shearing boxes (Hawley et al. 1995) with x → −ySB, y → xSB and z → zSB. In this rotating frame, one obtains the following set of governing equations subsection. Since we are looking for 2D solutions in the (x, y) plane, we can omit the Coriolis force as it will only change the pressure distribution. We define this vortex by an elliptical patch of constant vorticity ωt = −S + ωv (the "core") where −S is the background flow vorticity and ωv is the vorticity of the vortex itself. Outside of this core, the vorticity is assumed to be ωt = −S, extending to infinity. According to Kida (1981) (Eq. 2.9), such a vortex is steady if the semi-major axis is aligned with x and if its vorticity satisfies ωv S = − (3) 1 χ(cid:16) χ + 1 χ − 1(cid:17), where we have defined the vortex aspect-ratio χ = a/b, a and b being respectively the vortex semi-major and semi- minor axis. One deduces from this result that only anticy- clonic vortices (Ω and ωv having opposite sign) are stable in Keplerian flows. Since this solution is steady, no stream- line goes through the core boundaries, and the streamlines inside the core have to be elliptical, with the same aspect- ratio as the vortex core. Thanks to this property, we can write the velocity field in the vortex core, assuming it's centered on x = y = 0, as u0 x = S χy, 1 χ − 1 1 u0 y = −S This solution can be written in the simpler form (χ − 1) x. 1 χ u0 i = SAij xj, defining A = 1 χ − 1  0 0 0 χ 0 −χ−1 0 0  . 0 (4) (5) (6) (7) 2.3. Explicit solution A complete solution for the velocity field can be found defining the streamfunction ψ(x, y) so that ux = −∂yψ and uy = ∂xψ. This streamfunction satisfies: ∆ψ =(cid:26) −S + ων inside the core, outside. −S (8) One solves these equations in elliptical coordinates, defining (µ, ν) by ∂tu + ∇ · (u ⊗ u) = −∇Π − 2Ω × u + 2ΩSye ∇ · u = 0. y, (1) (2) x = f cosh(µ) cos(ν), y = f sinh(µ) sin(ν), (9) (10) In these equations, we have defined the mean shear S = −r∂rΩ, which is set to S = (3/2)Ω for a Keplerian disc. The generalised pressure Π = P/ρ0 is calculated solving a Poisson equation with the incompressibility condition. One can check easely that the velocity field u = Sye x is a steady solution of these equations. 2.2. The Kida (1981) solution We want to study the stability of a steady elliptical vortex embedded in the sheared flow described in the previous with f = ap(χ2 − 1)/χ2. In these coordinates, the core boundary is found at µ = µ0 with tanh(µ0) = χ−1. Requiring that ψ and ∂µψ are continuous at µ0, one finds the following expressions for ψ: ψi = − Sf 2 2(χ − 1)(cid:16)χ−1 cosh2(µ) cos2(ν) +χ sinh2(µ) sin2(ν)(cid:17), (11) G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 3 which leads to: k(t) = k0(cid:16) sin(θ) cos(cid:2)φ(t)(cid:3)e −χ sin(θ) sin(cid:2)φ(t)(cid:3)e + cos(θ)e y x z(cid:17) (16) where k0, θ and t0 are integration constants and φ(t) = S/(χ − 1)(t − t0) is the turnover angle. Therefore, the per- turbations will have a rotating wavevector with a turnover time equal to the vortex turnover time T = 2π(χ − 1)/S. Because of the incompressibility condition, θ = 0 will corre- spond to horizontal motion perturbations whereas θ = π/2 will imply vertical ones. We also note that the rotating wavevector will involve smaller structures in the y direc- tion. According to this result, the horizontal aspect ratio of the perturbation wavevector is equal to that of the vortex (χ). We can then simplify (13) eliminating the pressure, which leads to the final system (17) =h(cid:16) 2kikj k2 − δij(cid:17) ¯Rjmivm, k2 − δij(cid:17) ¯Ajm + 2(cid:16) kikj dvi dφ where ¯A = (χ− 1)A and ¯Rjm = (χ− 1)ǫjlmΩl/S. Plugging solution (16) in this equation leads to a Floquet problem for v, as already pointed out by Bayly (1986). Interestingly, the Floquet problem doesn't depend on the norm of the wavevector k ≡ k0, but just on its direction. Therefore, this problem is scale-independant, at least in the inviscid limit. The solution to this problem is known to be a superposition of Floquet modes written v = exp(γt)f [φ(t)], (18) where f is periodic with period 2π. To determine the Floquet exponents γ, we compute the eigenvalues exp(γT ) of the matrix M (2π) where M (φ) satisfies the generalised equation dMin dφ =h(cid:16) 2kikj k2 −δij(cid:17) ¯Ajm +2(cid:16) kikj k2 −δij(cid:17) ¯RjmiMmn, (19) with the initial condition Mij(0) = δij. (20) A numerical approach is required to solve this problem in most cases. However, some limits can be understood using analytical approaches, as we will see in the next section. 3. The elliptical instability 3.1. Horizontal instability It is possible to derive an analytical criterion for the insta- bility in the limit k = kz e z. In this limit, vz = 0 and (17) is reduced to dvx dφ dvy dφ (χ − 1)ivy, (χ − 1)ivx. = h − χ + = h 1 χ − This system describes horizontal epicyclic oscillations with frequency 2Ω S 2Ω S (21) (22) Fig. 1. Vortex solution streamlines in a sheared flow with χ = 3. As expected, inside the vortex (delimited by the blue bold line) the streamlines are elliptical, with the same aspect ratio as the vortex itself. ψo = − Sf 2 4(χ − 1)2h1 + 2(µ − µ0) +2(χ − 1)2 sinh2(µ) sin2(ν) χ − 1 + χ + 1 exp[−2(µ − µ0)] cos(2ν)i, (12) where the subscripts i and o stand for inside and outside the core. As expected, the inner solution reproduces the ellipti- cal flow described in the previous subsection (Eqns. 4 -- 5). In the outer solution, one recognises the background shear (3rd term), which dominates for large µ. Moreover, this solution exhibits a linear term in µ corresponding to the long-range perturbation of the vortex. In cartesian coordi- nates, this linear dependance translates into a logarithmic tail for ψ(x, y), showing that the vortex presence can be felt far from the core. As we will see in section 5 below, this property leads to numerical artefacts when one tries to fit this solution in a finite-size box. For completeness, we show in Fig. 1 the streamlines obtained from solution (11) -- (12). As expected, the streamlines inside the core are ellipses of constant aspect ratio χ. Outside the core, one still finds closed streamlines but with a much more elongated struc- ture. 2.4. Perturbation equations In the following, we concentrate on the evolution of pertur- bations in the vortex core, assuming the latter is infinite. This corresponds to a limit in which the perturbations are small compared to the typical horizontal size of the core. To study the evolution of 3D perturbations in a 2D ellip- tical flow, we write the total velocity field as u = u0 + v where v is supposed to have an infinitely small amplitude. Following Kelvin (1880) and Craik & Criminale (1986), we use a time explicit Fourier decomposition for the perturba- tion v = v(t) exp(ik(t) · x). Plugging this solution in (1) leads at first order to: vi + ixkvi( kk + SkjAjk) = −ikiΠ − SvjAij −2ǫijkΩjvk kjvj = 0 (13) (14) where we have included the tidal term of (1) in Π. To satisfy (13) for any xk, one has to solve: kk + SkjAjk = 0, (15) κ2 = S 2(cid:16)R − χ χ − 1(cid:17)(cid:16)R − 1 χ(χ − 1)(cid:17), (23) 4 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices having defined the rotation number as R = 2Ω/S. In the limit χ → ∞ where the vortex is infinitely weak, we find the classical epicyclic frequency κ2 = 2Ω(2Ω − S), which is equal to the Keplerian frequency in discs. This epicyclic frequency is imaginary when 1 2 + p1 + 4/R 2 < χ < R R − 1 . (24) If a columnar vortex is in this regime, its horizontal layers will tend to drift exponentially in the (x, y) plane, leading to the destruction of the vertical coherence of the structure. In accretion disc vortices, this regime is found for 3/2 < χ < 4. 3.2. A generalization We note that the above discussion may be generalized to ap- ply to a wider class of steady flows (u0 y). The linearized equations of motion are x, u0 Dvi + vj ∂u0 i ∂xj − 2Ωǫijzvj = − ∂Π′ ∂xi , 0 which direcrly leads to (24). To show this we first note that if α =(cid:18) ∂u0 ∂y − 2Ω(cid:19) and β = ∂u0 y ∂x x + 2Ω! (30) then both α and β must be either positive or negative. Without loss of generality we assume both to be positive (the case when they are both negative may be recovered by setting vx → −vx). Then from equations (27)-(28) we may obtain d(vxvy) dt = −αv2 y − βv2 x. or setting qy = −vy d(vxqy) y + βv2 x. = αq2 dt (31) (32) Thus if αmin and βmin are the minimum values of α and β respectively, we have (25) d(vxqy) dt ≤ αminq2 y + βminv2 x, (33) where Π′ is the pressure perturbation. The operator or equivalently D ≡ ∂ ∂t + u0 j ∂ ∂xj (26) denotes the convective derivative. We now assume pertur- bations are local in z so that we may assume that the z dependence is through a factor exp(ikzz), where kz is very large in the sense that the length scale k−1 is smaller than any other in the problem. From the incompressibility con- dition ∇ · v = 0, and the z component of (25) we conclude that Π′ = O(k−2 z ). Thus it may be dropped from the x and y components of (25) which then give the pair z Dvx + vx Dvy + vy ∂u0 x ∂x ∂u0 y ∂y x = −vy(cid:18) ∂u0 = −vx ∂u0 ∂y − 2Ω(cid:19) , + 2Ω! . y ∂x (27) (28) Replacing D by d dt , with the understanding that the time derivative is to be taken on a fixed streamline, we see that in general the system (27)-(28) leads to a second order or- dinary differential equation with periodic coefficients, the period being the time to circulate round the chosen stream- line. But note that for the problem on hand the coefficients are constant because the unperturbed velocity is a linear function of the coordinates. In more general cases one has a Floquet problem of the type described above for every streamline indicating that unstable modes are indeed local- ized on streamlines. In order to connect with (24) we note that one can prove that if everywhere on a chosen stream- line x ∂y − 2Ω(cid:19) ∂u0 (cid:18) ∂u0 y ∂x + 2Ω! > 0, (29) there will be an instability. In fact for the problem on hand this is precisely equivalent to the above condition that κ2 < d(vxqy) dt −2pαminβminvxqy ≤(cid:16)√αminqy −pβminvx(cid:17)2 Thus it follows that vxqy grows faster than exponentially with growth rate 2√αminβmin which means that there is a Floquet exponent corresponding to exponential growth. .(34) 3.3. Numerical results In the general case (θ > 0 and χ > 1), the Floquet prob- lem (17) can't be solved analytically. We therefore use the following numerical approach. We first evolve the M ma- trix following Eq. 19 for a given χ and θ with a 4th order Runge-Kutta-Fehlberg algorithm, keeping the error down to 10−10. The eigenvalues are then computed from M (2π) using the GNU scientific library routines. This procedure is repeated for 1000 values of χ and 1000 values of θ to get a 2D representation of the regions for which the instability is found. To test our numerical results, we have first reproduced the results of Bayly (1986) for a vortex in a non rotat- ing sheared flow. One find in this case that vortices are unconditionally unstable to fully 3D disturbances (Fig. 2). Moreover, in the limit χ → 1, only modes with θ = 60◦ are found to be unstable, consistently with Bayly's findings. Finally, we note that the growth rate for the instability weakens as we elongate the vortex. This finding, although apparently opposite to the classical result for the ellipti- cal instability, is due to the fact that our growth rate is measured in inverse shear time and not in inverse vortex turnover time. We have used our numerical method to solve the Floquet problem in the Keplerian case (Fig. 3). In this case, we observe essentially a two bands structure. The first band (low-χ band) is located between χ = 1 and χ = 4. It is associated with a short growth time (typically of the order of one shear time), and it includes the horizontal in- stability described in section 3.1 for θ = 0. In the limit χ → 1 (infinitely strong vortex), we find the same result G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 5 Fig. 2. Growth rate (in shear time) of the 3D instability deduced numerically from the Floquet problem (17) in the non rotating case. This result is identical to the original elliptical instability described by Bayly (1986) and allow us to test our numerical approach. The colour scale represent the logarithm of the growth rate. as in the non rotating case (instability for θ = 60◦), which was to be expected since the vortex turnover time is much smaller than the rotation time in this limit. The second band (high-χ band) is found for χ > 6 and is much weaker since it involves growth rate of the order of 10−2S. This band gets wider and weaker as we go to large aspect ratios (and weak vortices), with a growth rate maximum found for χ ≃ 11. We have tried to get a larger resolution in the re- gion χ ∼ 5.5 -- 6 and it seems that this band reaches θ = 0◦ for χ ∼ 5.9, but the growth rates involved are very small. Note that another narrower band is observed in the regime of high χ and smaller θ in Fig. 3. As we will see in the following, this band is due to higher order resonance and is therefore weaker than the bands described previously. To get a simpler representation of the elliptical insta- bility in the Keplerian case, we have plotted the maximum growth rate as a function of χ in Fig. 4. We observe the same basic features as on the colormaps with 2 bands. We also observe that the high-χ band decays as χ gets larger. This property is to be expected since the limit χ → ∞ corresponds to a pure shear flow, which is known to be lin- early stable. However, for a finite (but large) χ, vortices are always unstable, contrary to claims in the literature (Lithwick 2007). Therefore, it seems that no elliptical instability is found for 4 < χ < 5.9. However, we recall that up to now we have only considered perturbations interior to the vortex core. But as we will see later, instabilities may also appear outside the vortex core. 3.4. Physical interpretation It may be noted that the system (17) can be written as a single Hill's equation (see for example Waleffe 1990). In this case, one finds that the periodic excitation frequency Fig. 3. Growth rate of the 3D instability deduced numeri- cally from the Floquet problem (17) in the Keplerian case for large χ. One observes the vertical modes for 3/2 < χ < 4 and full 3D modes for χ < 3/2 and χ > 6 up to at least χ = 100. A weak secondary band is found under the pri- mary band for large χ. 100 10−1 γ 10−2 101 χ Fig. 4. Maximum growth rate in the Keplerian case (in shear time). We remark the low-χ band with large growth rates (χ < 4), the high-χ band with smaller growth rates (χ > 6) and a stable region in between. The growth rate decays for large χ when the vortex gets weaker, as expected. is twice the vortex frequency S/(χ − 1), since the time- dependent wave numbers kx and ky always appear quadrat- ically. In the limit θ → 0 for which the excitation is small, one would expect a series of instability bands. By analogy with Mathieu's equation, one should find such a band when the local epicyclic frequency κ is an harmonic of the vortex frequency: κ = n S χ − 1 , (35) with n ǫ N . In the Keplerian case, one would therefore expect instability bands arising from θ = 0 for χ1 ≃ 4.65, χ2 ≃ 5.89, χ3 ≃ 7.28. . . However, according to Fig. 3, the first unstable band χ = 4.65 (n = 1) doesn't appear. This 6 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices peculiar property can be understood writing the epicyclic frequency as (36) κ2 = S 2hRΓ + 1 (χ − 1)2i, where Γ = ωt/S + R is the absolute vorticity of the back- ground vortex. According to this expression, the n = 1 resonance condition is equivalent to the constraint Γ = 0. Interestingly, the absolute vorticity also appears explicitly in the linearized vertical vorticity equation: ∂tωz + u0 · ∇ωz = SΓ∂zvz (37) where ωz = ∂xvy − ∂yvx is the vertical vorticity of the per- turbation. According to this equation, ωz is a conserved quantity if the background absolute vorticity Γ is zero. We conclude from this argument that the n = 1 resonance con- dition cannot lead to an instability, since ωz could not be conserved in that case. Therefore, the first unstable band appears for n = 2 (χ ≃ 5.89), which is consistent with our numerical result. As stated before, higher order bands (n > 2) are also observed in the computation (Fig. 3), but they are much weaker, and one can't check the origin of these bands on the θ = 0 axis with enough precision to compare with eq. (35). 4. Stratified case 4.1. Model and Equations Barranco & Marcus (2005) showed using anelastic numeri- cal simulations that off-midplane vortices were able to sur- vive for several hundred orbits whereas midplane vortices were destroyed rapidly. This suggests that stratification might be a way to stabilise vortices and suppress the el- liptical instability. To study this problem in a simple con- figuration, we have considered the case of a 3D vortex cen- tered at z = z0, i.e. above the midplane. We then follow the evolution of perturbations with k ≫ z−1 inside the vortex. In first approximation, the evolution of these perturbations can be described using the Boussinesq approximation which can be written ∂tu + ∇ · (u ⊗ u) = −∇Π − 2Ω × u + 2ΩSye (38) y 0 +N ζe z, ∂tζ + ∇ · (uζ) = −N uz, ∇ · u = 0, where ζ is the potential temperature and N is the Brunt- Vaisala frequency. In the following, we will assume a stable vertical stratification, i.e. a real Brunt-Vaisala frequency. The vortex solution found previously is still valid in the Boussinesq approximation. Therefore, we can follow the same procedure as the one used in the non stratified case to get dvi dφ dζ dφ k2 − δij(cid:17) ¯Ajm + 2(cid:16) kikj = h(cid:16) 2kikj (cid:16) kikj N (χ − 1) k2 − δij(cid:17)ζδjz , − N (χ − 1) = − vz. S S k2 − δij(cid:17) ¯Rjmivm (41) These equations associated with the solution (16) lead to a stratified Floquet problem. As previously, this problem (39) (40) (42) Fig. 5. Growth rate of the 3D instability deduced numeri- cally from the Floquet problem (41) in the Keplerian case with N/S = 1.0. One observe once again the vertical modes for 3/2 < χ < 4 and many weak bands involving fully 3D perturbations. The colour scale represent the logarithm of the growth rate. can be solved by diagonalizing a 4× 4 evolution matrix M , following the same procedure as in the non stratified case. The numerical method used to solve these equations is the same as in the previous section, except we have added an extra dimension in the solver to handle the evolution of the potential temperature. 4.2. Results We plot the growth rate of the instability as a function of χ and θ for a moderately stratified vortex N = S in Fig. 5. We note that the influence of the stratification is strong in the domain in which the vortex is weak (χ > 5), and signif- icant differences are observed when comparing with Fig. 3. First, we remark that the stable region observed in the non stratified case (4 < χ < 5.9) is not present in the N = S stratified case. This result is due to the presence of high frequency buoyancy modes which can match the turnover frequency when the inertial modes cannot. Moreover, the high-χ band has been replaced by a series of weaker bands. These bands can be understood as reminiscences of the res- onance condition (35), since in the limit θ → 0 the effect of stratification disappears. When we compute the maximum growth rate as a func- tion of χ for various stratification frequencies (Fig. 6), we note that the growth rate at large χ decreases significantly with stronger stratification. In this sense, the stratification tends to weaken the elliptical instability, but does not sup- press it. 5. Non-linear simulations 5.1. Resolving the elliptical instability As pointed out in the previous section, the elliptical in- stability is always present for highly elongated vortices (χ > 6), although it is quite weak. We have also found that no elliptical instability was observed for 4 < χ < 5.9 when stratification was omitted. To check these results, one has G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 7 N/S=0 N/S=0.1 N/S=1 γ 10−1 10−2 10−3 2 4 6 8 χ 10 12 14 Fig. 6. Maximum growth rate in the Keplerian case (in shear time) for various stratification frequencies. As the stratification gets stronger, the growth rate decreases for weak vortices (χ > 6). We also note that for a strong enough stratification, the region without instability (4 < χ < 5.9) disappears. Finally, the horizontal instability described in the previous section (3/2 < χ < 4, θ = 0) is not affected by the stratification, as expected. y x ∼ kz and kmax to perform non-linear simulations in which this instability is resolved. A condition to get this instability in a numeri- cal simulation may be found assuming we have a 2D vortex embedded in a disc with a height H, the vortex core being larger than H. In the limit χ ≫ 6, we find the elliptical instability for θ ∼ π/3 (Fig. 3). According to (16), unstable modes will have in this case kmax ∼ χkz. Moreover, kz > 2π/H since 3D perturbations have to fit vertically inside the disc. If we assume we need n points to resolve one wavelength (n > 2), then the longest wave- length modes unstable to the elliptical instability will be resolved if the x resolution is n points per scale height and the y resolution is nχ points per scale height. This last con- dition is not often satisfied in 3D global simulations, and may explain why long-lived and stable vortices are some- times observed. Note that the value n may be quite high when using low order finite difference methods (one might expect n ∼ 10) since one wants the (numerical) dissipation rate for these wavelengths to be smaller than the growth rate, which is itself small compared to the shear frequency. Therefore, very accurate (e.g. spectral) numerical methods are preferable to study this kind of weak instability. To check for resolution artefacts, the non stratified in- compressible simulations and the compressible simulations carried out with NIRVANA below have been computed with two different resolutions (the "high" resolution being twice the "low" resolution in each direction). No significant differ- ence was found between high resolution and low resolution runs. In particular, the same growth rates were obtained, with the same localisation properties for the instabilities. However, in low resolution runs carried out with NIRVANA that were seeded with truncation errors, the unstable scales were initially a factor ∼ 2 larger, but with the growth rates and later nonlinear behaviour being very similar. This can Fig. 7. Vortex streamlines in a sheared flow with χ = 3. This solution is computed with an FFT method, assuming periodic boundary conditions in x and y. This solution and the infinite solution (Fig. 1) are extremely similar, except around (x = ±5, y = 0) where an X point is observed. be understood as a consequence of the scale free property of the linear instability that holds once scales are sufficiently small but still larger than the dissipation scale, here ex- pected to be the grid scale. Noting that conclusions would be unaltered if the low resolution runs were adopted, only high resolution results will be presented in this section. 5.2. Numerical solution for an isolated vortex To check the existence of the elliptical instability in ac- cretion disc vortices, one needs a solution for an isolated vortex. In section 2, we have derived an exact solution for a vortex in an infinite box. Numerically however, one has to work in a finite size domain, using essentially the shearing- sheet boundary conditions (Hawley et al. 1995). As pointed out previously, the infinite solution has a slowly decaying tail which can lead to artefacts when one starts with this so- lution in a finite-size shearing box. When one tries this pro- cedure, the 2D solution rapidly develops instabilities near the boundaries, especially along the x axis, leading to large amplitude oscillations. To prevent this, we have chosen to solve (8) using fast Fourier transforms (FFT), which are a natural choice for our boundary conditions. A solution obtained by this method in a (10 × 4) box is shown in Fig. 7. The solution found by this procedure is very sim- ilar to the infinite solution, except near (x = ±5, y = 0) where an X point is observed. This difference probably ex- plains the unsteadiness observed when one uses the infinite solution numerically. We find that periodic solutions de- rived using the FFT procedure are better suited for numer- ical simulations. Although these solutions are not exactly steady (mainly because of time varying boundary condi- tions), they keep the same shape and the same amplitude for several turnover times when one uses high precision nu- merical methods, which is enough for our purpose. As pointed out previously, this vortex model is not a proper infinite elliptical flow, and the approximation used in our linear analysis may not be valid. Nevertheless, the core of these elliptical vortices is made of elliptical stream- lines. Since our linear analysis is essentially a local analysis of stability on individual streamlines, one expects our re- sults to apply to the core of such vortices. Note however that other parametric instabilities may appear outside of the vortex core because closed streamlines still exist in this region (see Fig. 7). We have computed the evolution of several isolated el- liptical vortices in a shearing box with dimensions Lx × 8 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 2 / 1 v < > z2 χ=3 χ=5.5 χ=11 10−2 10−3 10−4 10−5 10−6 10−7 0 50 100 150 200 250 300 350 400 450 t y t i c o e v l l a c i t r e v e r a u q s n a e m t o o R 1 0.1 0.01 0.001 0.0001 1e-05 0 20 40 60 80 100 shearing time Fig. 8. Time evolution of phv2 zi for several isolated vor- tices in shearing box simulations with no stratification. We find an exponential growth in agreement with the linear prediction for χ = 3 and χ = 11. We also find an instability for χ = 5.5 although the elliptical streamlines are stable in our linear analysis. Fig. 11. The root mean square vertical velocity obtained from the compressible simulations is plotted as a function of time for χ = 2. The cases shown are vs/cs = 1.3 (dotted line), vs/cs = 0.65 (dashed line), and vs/cs = 0.13 (full line). Initial values increase with vs/c because the initially imposed state variables are further from being in a station- ary state. Ly × Lz using an incompressible spectral method (see Lesur & Longaretti 2005, for a complete description of the numerical method). We have used (Nx × Ny × Nz) = (256 × 512 × 64) points and an eighth order hyperviscosity instead of the classical viscosity to prevent the dissipation of the weaker unstable modes. The vortices are initialised in the center of the box setting a vorticity patch with dimen- sion (λx, λy). We always set λy = 1, Ly = 4 and Lz = 2, λx and Lx being adjusted according to the vortex aspect-ratio χ. For χ = 3, we set Lx = 10 whereas for χ = 5.5 and χ = 11 we set Lx = 20. For each run, we initialise the ve- locity field solving (8) in Fourier space, and we add a small (10−7) 3D white noise to the 2D solution. 5.3. Non stratified case phv2 To follow the evolution of 3d motions, we have computed zi, where hi denotes a volume average procedure. Time histories of this quantity are given in Fig. 8 for χ = 3, χ = 5.5 and χ = 11. As expected from our linear analysis, we find an exponential growth for both χ = 3 and χ = 11. Since the growth rate for χ = 3 is very fast, the instability saturates at t ≃ 40 and the vortex structure is destroyed, as shown in Fig. 9. In the exponential regime, we find a growth rate γ = 0.37 for χ = 3 and γ = 0.016 for χ = 11, in agreement with theoretical values (γ3 = 0.44 ; γ11 = 0.017). We also find an instability for χ = 5.5 with γ5.5 = 0.055 which was not expected in our linear analysis (χ = 5.5 elliptical streamlines are stable according to Fig. 4). To check the localisation of the instability, we present (x, y) slices of the vertical velocity for χ = 3, χ = 5.5 and χ = 11 in Fig. 10. We find that unstable modes are localised inside the vortex for χ = 3 and χ = 11, as expected from our linear analysis. However, for χ = 5.5, the instability appears outside of the vortex core, in a region in which one find closed non-elliptical streamlines (see Fig. 1). In this particular case, no vertical motions are found in the vortex core, which is consistent with our analysis. The existence of this outer instability can be understood quite simply following the physical description used for the elliptical instability (section 3.4). We know that these insta- bilities arise when the epicyclic frequency is an harmonic of a closed streamline frequency multiplied by 2π (the stream- line frequency being the inverse of the time needed by a fluid particle to cover the whole streamline and get back to its starting point). Moreover, as one moves away from the vortex core, the streamline frequency tends to 0 since the flow tends towards a pure shear flow, whereas the epicyclic frequency tends to the Keplerian frequency. Therefore, one will always be able to find a resonance outside of the vortex core, and consequently a parametric instability. 6. Effects of compressibility We have also investigated the effects of compressibil- ity. To do this we performed three dimensional hydrody- namic simulations using NIRVANA (Ziegler & Yorke 1997). NIRVANA has been used frequently in the past to study various problems involving MHD turbulence in the shearing box (Fromang & Papaloizou 2006; Papaloizou et al. 2004). Because of its diffusive character, it is best suited to vortices with small χ that are not too elongated. For this reason we limit consideration to χ = 2. We consider isothermal shear- ing boxes with vertical gravity. In units such that the full radial width of the elliptical vorticity patch is unity, the box dimensions were (Lx = 3.46, Ly = 1.73, Lz = 2.31). The numbers of grid points used were (Nx = 196, Ny = 98, Nz = 128). A vertical gravitational acceleration −Ω2z was applied over 80% of the vertical computational domain, being set to zero in two domains extending from the verti- cal boundaries, each being of an extent equal to 10% of the whole vertical domain. The boundary conditions of period- icity in shearing coordinates could then be applied. We com- ment that the use of periodic boundary conditions in the vertical direction, with small buffer zones near the bound- aries that allow these conditions to be applied consistently, has been found to be effective in vertically stratified shear- ing box simulations of MHD turbulence (eg. Stone et al. 1996). This approach allows for regular treatment of the G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 9 Fig. 9. 3D snapshots showing the evolution of a χ = 3 vortex at t = 26 (top left), t = 36 (top right), t = 41 (bottom left) and t = 47 (bottom right). We have represented in transparent grey an isocontour of vorticity delimiting the vortex core, and in colour the volume rendered vertical velocity normalised. We find that the instability is localized inside the vortex and is dominated by small scale vertical wave-numbers (θ ∼ 0), as expected. Moreover, the vortex structure is destroyed when the instability saturates, i.e. for t ≃ 40. boundary without significant artefacts being produced in the mid-plane regions of the flow, as would be expected for local instabilities of the type considered here. To test this as well as the effects of expanding the computational do- main in general, we have performed a test simulation with the compuational domain doubled in size in each direction for the case with the largest value of vs/cs considered be- low. The dimensions of the elliptical vorticity patch were not changed. The growth rate and nonlinear development of the instability were indeed essentially the same in the mid-plane regions, which because of higher inertia carried most of the energy, as in the smaller domain simulations presented below. The instability in the larger domain case readily spreads to the more extensive regions outside the original vorticity patch. We considered three values of the isothermal sound speed cs. Adopting the shearing time as the unit of time, these were such that vs/cs were 1.3, 0.65, and 0.13. Here the velocity vs, being unity in our dimensionless units is the maximum velocity difference across the vorticity patch due to the background shear. The incompressible two dimensional flow field calcu- lated in section 5.2 was applied on horizontal planes. As above this was calculated by solving the vorticity equa- tion with the imposition of periodic boundary conditions. To make this possible, one must subtract out the mean vorticity in the patch so that the net surface integral over the (x, y) plane is zero, which results in some un- steadiness as described above. However, in the compress- ible case, additional unsteadiness occurs because an asso- ciated unbalanced density change is generated. This was calculated and included in the set up by noting that for a strictly steady state horizontal flow of the type consid- s ln ρ + (1/2)u2− (3/2)Ω2y2 + ΦG, with ered here, Q(ψ) = c2 ΦG = Ω2z 2/2 being the gravitational potential, should be a function only of the stream function ψ. This is readily de- termined from the imposed steady flow from the condition dQ dψ = 1 2 Ω + ωv, (43) where ωv is taken to be the same function of ψ as for the infinite vortex patch. The latter is of course an approxima- tion as we modified the infinite patch by imposing periodic boundary conditions. Nonetheless its use to calculate the initial density on horizontal plane resulted in initial states that became increasingly more steady as cs increased. In each of the simulations the two dimensional vortex structure was found to remain for about 40 shearing times before undergoing a linear instability which began with a very short wavelength in the vertical direction. The root mean square (weighted with density) vertical velocity is plotted in Fig. 11. Initial values increase with vs/c because the initially imposed flow is further from being stationary. 10 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices Fig. 10. 2D snapshots of the vertical velocity field in the (x, y) plane for χ = 3 (top), χ = 5.5 (middle) and χ = 11 (bottom). The vortex core is delimited by a thick grey line. We find that 3D perturbations are localized inside the vortex core for χ = 3 and χ = 11 whereas they are found outside of the vortex core for χ = 5.5. Streamlines in the horizontal midplane are shown for vs/cs = 1.3 at t = 37, and vs/cs = 0.13 at t = 47 in Fig. 12. Although linear instability has begun, the initially imposed flow is found to be well preserved for the case with vs/cs = 0.13 which most closely resembles the incompressible case. In this case the maximum growth rate in the linear phase of 0.75S is in good agreement with the linear prediction obtained from equation (23). However, when vs/cs = 1.3 the flow shows significantly greater time variability. The vortex is squeezed in the ra- dial direction and trailing features begin to form outside the initial vortex core which turn into density waves which, however, are not well represented in our small computa- tional domain. Global vortical structures survive until they are eventually destroyed by the instability Snapshots of the vertical velocity field in the horizontal midplane are shown after the onset of linear instability in Fig. 13 for the three simulations. The 3D perturbations are found to be local- ized inside the original elliptical boundary of the vorticity patch just as in the incompressible case. tions. We followed the same procedure as in the non strat- ified case to set up the vortex structure. We have run the simulations with a moderate stratification (N = S), for which one expects a departure from the non stratified case zi1/2, for various when χ > 4 (Fig. 5). The time history of hv2 vortex aspect-ratios, is presented in Fig. 14. For χ = 3 we find γ = 0.35 and γ = 0.01 for χ = 5.5. No instability is observed in our simulations for χ = 11. These measured growth rates are always below the ex- pected growth rate from our linear analysis (γ3 = 0.43, γ5.5 = 0.024 and γ11 = 0.004) which might be due to exces- sive numerical dissipation in the stratified case, despite the high resolution and the hyper viscosity prescription used. For both γ = 3 and γ = 5.5, the instability appears to be localised inside the vortex core, as expected. We can con- clude from our results that the elliptical instability can be detected in stratified simulations, but resolution and nu- merical dissipation must be carefully controlled to get ac- curate results. 6.1. Stratified case To study the vertically stratified case, we have imple- mented the Boussinesq equations (38) in our spectral code. Compared to (38), we have added an eighth order hyper diffusivity to the velocity and potential temperature equa- We have described a 3D instability appearing in elliptical vortices (Kida vortices) embedded in accretion discs, known as the elliptical instability. We have shown analytically, for the case of no stratification that this elliptical instability is always found in vortex cores, except for a narrow range 7. Discussion G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 11 Fig. 12. Streamlines in the (x, y) plane for χ = 2 obtained from the compressible simulations shown after the onset of linear instability. These are with vs/cs = 1.3 at t = 37 (left), and vs/cs = 0.13 at t = 47 (right). The initially imposed flow is well preserved when vs/cs = 0.13. When vs/cs = 1.3 the flow shows greater time variability, with the vortex being sqeezed in the radial direction. Some trailing features are established outside the initial vortex core. χ=3 χ=5.5 χ=11 10−2 10−3 10−4 10−5 10−6 10−7 2 / 1 v < > z2 0 50 100 150 200 t 250 300 350 400 Fig. 14. Time evolution of phv2 zi for several isolated vor- tices in stratified shearing box simulations with N = S. We find an exponential growth in agreement with the linear prediction for χ = 3 and χ = 5.5. The expected instability for χ = 11 is not observed, probably because of a too large numerical dissipation. of vortex aspect-ratio (4 < χ < 5.9). When the vortex is weak (elongated vortex), the instability involves small ra- dial wavelengths compared to azimuthal and vertical ones, the ratio being of the order of χ. Moreover, the inclusion of a stable stratification tends to weaken the instability for large χ but it does not suppress it. Numerical simulations have essentially confirmed our linear analysis. We have found the predicted growth rate in the case where the instability was expected inside the core. Furthermore, instabilities outside of the vortex core were observed when the core was linearly stable, in agreement with our physical interpretation. The inclusion of compress- ibility in the simulations didn't change the behaviour of the instability, and analytical results were recovered. We also studied numerically the effect of vertical Boussinesq stratifi- cation. Again, analytical results were recovered, except for weak vortices (χ = 11) for which stability was observed, probably because of numerical diffusion. These results tend to indicate that elliptical instabili- ties, or more generally parametric instabilities, are always present in accretion disc vortices. However, they often in- volve small radial wavelengths (compared to the disc thick- ness) and small growth rates (compared to the orbital time scale). Put together, these properties make elliptical insta- bilities hard to capture numerically because of numerical diffusion at the grid scale. For example, highly accurate spectral methods with high-order hyper-diffusivity were re- quired to marginally resolve the instability in a χ = 11 vortex. In this work, we have not addressed the saturation of this instability. This is however rather intentional, as we think that the non-linear outcome should be studied to- gether with the production mechanism for these vortices. In fact, the evolution of these structures will be dictated ul- timately by the competition between the production mech- anism and the saturated 3D instability. In this context, studying the saturation of this instability on its own will be of little interest, since it will be modified by the vortex production process. Note too that the time-scale needed by the saturated instability to destroy the vortex structure is unknown. In particular, there is no reason to think that this time-scale is related to the growth rate in the linear stage. For example, a weakly unstable vortex could be de- stroyed in one orbit once saturation is reached. Therefore, no firm conclusion can be drawn from this work concern- ing the evolution of the vortex structure itself. However, in the absence of a sustaining production process, eventually one would expect the destruction of the vortex structure when the perturbation amplitude reaches the background flow amplitude as was indeed observed for our χ = 3 simu- lation. This work has also several limitations. In particular, we have assumed an isolated 2D vortex as a starting point. However, one might want to study 3D vortices includ- ing a complicated vertical structure, with e.g. a mean vertical flow in the core (similar to cyclones on Earth for example). This kind of structure was suggested by 12 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices Fig. 13. 2D snapshots of the vertical velocity field in the (x, y) plane for χ = 2 obtained from the compressible simulations shown after the onset of linear instability. These are with vs/cs = 1.3 at t = 37 (top), vs/cs = 0.65 at t = 42 (middle) and vs/cs = 0.13 at t = 47. The elliptical boundary of the vorticity patch is overplotted. We find that the 3D perturbations are localized inside this boundary as in the incompressible case. Barranco & Marcus (2005) for off-midplane vortices, and there is no doubt that stability properties will be modi- fied in this case. Moreover, we have mostly restricted our study to elliptical streamlines, since these are tractable an- alytically. Although some extensions to more general cases were shown, we have not provided any formal proof of in- stability in the most general case. However, the instability mechanism exhibited in the elliptical case appears to be quite general, and we think it's unlikely that a steady vor- tex could be stable everywhere to 3D perturbations. parametric instabilities will certainly modify the global out- come of these processes, with probably new physical effects. Acknowledgements. The authors thank Jeremy Goodman, Gordon Ogilvie, Orkan Umurhan and Fran¸cois Rincon for useful discus- sions. The simulations presented in this paper were performed us- ing the Darwin Supercomputer of the University of Cambridge High Performance Computing Service (http://www.hpc.cam.ac.uk/), pro- vided by Dell Inc. using Strategic Research Infrastructure Funding from the Higher Education Funding Council for England. GL acknowl- edges support by STFC . Finally, these results point out the necessity of car- rying high-resolution 3D simulations of the mechanisms suggested in the literature for vortex formation. This includes for example Rossby wave instabilities (Li et al. 2001), baroclinic instabilities (Klahr & Bodenheimer 2003; Petersen et al. 2007) or gaps due to giant planets (de Val-Borro et al. 2007). In any case, the presence of 3D References Balbus, S. A. 2003, ARA&A, 41, 555 Balbus, S. A. & Hawley, J. F. 1998, Reviews of Modern Physics, 70, 1 Barge, P. & Sommeria, J. 1995, A&A, 295, L1 Barranco, J. A. & Marcus, P. S. 2005, ApJ, 623, 1157 Bayly, B. J. 1986, Phys. Rev. Lett., 57, 2160 Bodo, G., Tevzadze, A., Chagelishvili, G., et al. 2007, A&A, 475, 51 G. Lesur and J. C. B. Papaloizou: Stability of Elliptical Vortices 13 Craik, A. D. D. 1989, J. Fluid Mech., 198, 275 Craik, A. D. D. & Criminale, W. O. 1986, Proc. R. Soc. Lond. A, 406, 13 de Val-Borro, M., Artymowicz, P., D'Angelo, G., & Peplinski, A. 2007, A&A, 471, 1043 Drazin, P. & Reid, W. 1981, Hydrodynamic stability (Cambridge Univ. Press) Fromang, S. & Papaloizou, J. 2006, A&A, 452, 751 Gammie, C. F. 1996, ApJ, 457, 355 Godon, P. & Livio, M. 1999, ApJ, 523, 350 Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1995, ApJ, 440, 742 Johansen, A., Andersen, A. C., & Brandenburg, A. 2004, A&A, 417, 361 Johnson, B. M. & Gammie, C. F. 2005a, ApJ, 626, 978 Johnson, B. M. & Gammie, C. F. 2005b, ApJ, 635, 149 Johnson, B. M. & Gammie, C. F. 2006, ApJ, 636, 63 Kelvin, L. 1880, Phil. Mag., 10, 155 Kerswell, R. R. 2002, Ann. Rev. Fluid Mech., 34, 83 Kida, S. 1981, Physical Society of Japan, Journal, vol. 50, Oct. 1981, p. 3517-3520., 50, 3517 Klahr, H. H. & Bodenheimer, P. 2003, ApJ, 582, 869 Lesur, G. & Longaretti, P.-Y. 2005, A&A, 444, 25 Li, H., Colgate, S. A., Wendroff, B., & Liska, R. 2001, ApJ, 551, 874 Lithwick, Y. 2007, ArXiv e-prints 0710.3868, 710 Lovelace, R. V. E., Li, H., Colgate, S. A., & Nelson, A. F. 1999, ApJ, 513, 805 Onsager, L. 1949, Nuov. Cim. Supp., 6, 279 Papaloizou, J. C. B., Nelson, R. P., & Snellgrove, M. D. 2004, MNRAS, 350, 829 Petersen, M. R., Julien, K., & Stewart, G. R. 2007, ApJ, 658, 1236 Pierrehumbert, R. T. 1986, Phys. Rev. Lett., 57, 2157 Regev, O. & Umurhan, O. M. 2008, A&A, 481, 21 Shen, Y., Stone, J. M., & Gardiner, T. A. 2006, ApJ, 653, 513 Stone, J. M., Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1996, ApJ, 463, 656 Umurhan, O. M. & Regev, O. 2004, A&A, 427, 855 von Weizsacker, C. F. 1944, Zeit. fur Astrophys., 22, 319 Waleffe, F. 1990, Phys. Fluids A, 2, 76 Ziegler, U. & Yorke, H. W. 1997, Computer Physics Communications, 101, 54
1309.2849
1
1309
2013-09-11T15:03:12
The Production of Small Primary Craters on Mars and the Moon
[ "astro-ph.EP" ]
We model the primary crater production of small (D < 100 m) primary craters on Mars and the Moon using the observed annual flux of terrestrial fireballs. From the size-frequency distribution (SFD) of meteor diameters, with appropriate velocity distributions for Mars and the Moon, we are able to reproduce martian and lunar crater-count chronometry systems (isochrons) in both slope and magnitude. We include an atmospheric model for Mars that accounts for the deceleration, ablation, and fragmentation of meteors. We find that the details of the atmosphere or the fragmentation of the meteors do not strongly influence our results. The downturn in the crater SFD from atmospheric filtering is predicted to occur at D ~ 10-20 cm, well below the downturn observed in the distribution of fresh craters detected by the Mars Global Surveyor (MGS) Mars Orbiter Camera (MOC) or the Mars Reconnaissance Orbiter (MRO) Context Camera (CTX). Crater counts conducted on the ejecta blanket of Zunil crater on Mars and North Ray crater on the Moon yielded crater SFDs with similar slopes and ages (~1 Ma, and ~58 Ma, respectively) to our model, indicating that the average cratering rate has been constant on these bodies over these time periods. Since our Monte Carlo simulations demonstrate that the existing crater chronology systems can be applied to date young surfaces using small craters on the Moon and Mars, we conclude that the signal from secondary craters in the isochrons must be relatively small, as our Monte Carlo model only generates primary craters.
astro-ph.EP
astro-ph
1  2  3  4  5  6  7  8  9  10  11  12  13  14  15  16  17  18  19  20  21  22  23  24  25  26  27  28  29  30  31  32  33  34  35  36  37  38  The Production of Small Primary Craters on Mars and the Moon J.-P. Williams*, A. V. Pathare, O. Aharonson Jean-Pierre Williams* Dept. Earth and Space Sciences University of California Los Angeles, CA 90095, USA [email protected] 310-825-4414 (office) 310-825-2279 (fax) Asmin V. Pathare Planetary Science Institute Tucson, AZ 85719, USA [email protected] Oded Aharonson Center for Planetary Science Weizmann Inst. Of Science Rehovot, 76100 Israel [email protected] *Corresponding Author Keywords: Cratering; Mars; Mars surface   Page 1 of 45  39  40  41  42  43  44  45  46  47  48  49  50  51  52  53  54  55  56  57  58  59  60  61  Abstract We model the primary crater production of small (D < 100 m) primary craters on Mars and the Moon using the observed annual flux of terrestrial fireballs. From the size-frequency distribution (SFD) of meteor diameters, with appropriate velocity distributions for Mars and the Moon, we are able to reproduce martian and lunar crater-count chronometry systems (isochrons) in both slope and magnitude. We include an atmospheric model for Mars that accounts for the deceleration, ablation, and fragmentation of meteors. We find that the details of the atmosphere or the fragmentation of the meteors do not strongly influence our results. The downturn in the crater SFD from atmospheric filtering is predicted to occur at D ~ 10-20 cm, well below the downturn observed in the distribution of fresh craters detected by the Mars Global Surveyor (MGS) Mars Orbiter Camera (MOC) or the Mars Reconnaissance Orbiter (MRO) Context Camera (CTX). Crater counts conducted on the ejecta blanket of Zunil crater on Mars and North Ray crater on the Moon yielded crater SFDs with similar slopes and ages (~1 Ma, and ~58 Ma, respectively) to our model, indicating that the average cratering rate has been constant on these bodies over these time periods. Since our Monte Carlo simulations demonstrate that the existing crater chronology systems can be applied to date young surfaces using small craters on the Moon and Mars, we conclude that the signal from secondary craters in the isochrons must be relatively small, as our Monte Carlo model only generates primary craters.   Page 2 of 45  62  63  64  65  66  67  68  69  70  71  72  73  74  75  76  77  78  79  80  81  82  83  1. Introduction The accumulation of craters on a planetary surface can be used to determine relative ages of areas of geologic interest. Assigning absolute ages is done with modeled impact crater isochrons, a technique that has been developed over several decades (e.g. Hartmann, 1966; Neukum and Wise, 1976; Hartmann, 1999; Neukum and Ivanov, 1994; Hartmann and Neukum, 2001; Hartmann, 2005). For crater ages on Mars, isochrons are derived from the size-frequency distribution (SFD) of craters observed on the lunar maria for which we have dated Apollo samples (Wilhelms, 1987), scaled to account for the ratio of meteoroids at the top the martian atmosphere relative to the Moon’s and the differences in gravity and average impact velocity of intersecting orbits. The resulting isochrons yield an expected crater SFD for a given age surface and provide a means of understanding the absolute timescale of major geological and geophysical processes. The SFD is typically described as a power-law with slope n. Deviations from the power- law occur through various processes which, in general, preferentially alter the smaller diameter crater population, making small craters more challenging to use for age-dating surfaces. Because of the frequency at which small craters form however, they provide the ability to discriminate surface ages of geologically young regions and features at a higher spatial resolution where only small craters are available for dating. This level of resolution is required to establish the temporal relation of recent geologic activity on Mars such as gully and landslide formation, volcanic resurfacing, sedimentation, exhumation, dune activity, glaciation and other periglacial landforms, and the possible relation of such features to obliquity cycles of ~107 y timescale (e.g. Basilevsky et al., 2009; Burr et al., 2002; Hartmann and Berman, 2000; Kadish et al., 2008; Lanagan et al.,   Page 3 of 45  84  85  86  87  88  89  90  91  92  93  94  95  96  97  98  99  100  101  102  103  104  105  106  2001; Malin and Edgett, 2000a, 2000b, 2001; Mangold, 2003; Marquez et al., 2004; Quantin et al., 2007; Reiss et al., 2004; Shean et al., 2006; Schon et al., 2009). The Mars Obiter Camera (MOC) aboard the Mars Global Surveyor (MGS), with ~1.5 m resolution/pixel (Malin et al., 1992), identified 19 fresh craters over a ~6.8 year period (Malin et al., 2006). Data from the High Resolution Imaging Science Experiment (HiRISE) aboard the Mars Reconnaissance Orbiter (MRO) is currently providing image data with up to 25 cm pix-1 resolution (McEwen et al., 2007a), and has imaged and confirmed >200 small (< 50 m diameter) fresh impact craters having formed within the last few decades following their discovery by the Context (CTX) camera (Malin et al., 2007) on the same spacecraft (Byrne et al., 2009; Daubar et al., 2010, 2011, 2012; Daubar and McEwen, 2009; Dundas and Byrne, 2010; Ivanov et al., 2008, 2009, 2010; Kennedy and Malin, 2009; McEwen et al., 2007b, 2007c). This offers an opportunity to study the production of small meter-scale craters on the surface of Mars in greater detail and refine isochron models for dating young surfaces on Mars. In this paper, we model crater populations using a Monte Carlo simulation to explore the primary crater production function at small diameters and the potential influence of present-day atmospheric filtering. We explore ablation, deceleration, and fragmentation of projectiles as they traverse the martian atmosphere and compare our results with the 44 fresh craters reported by Daubar et al. (2013) that have been well constrained by CTX before- and after-images. We then assess the current impact-crater isochron model of Hartmann (2005) and the estimated ratio of meteoroids at the top of the martian atmosphere relative to the top of the terrestrial atmosphere, which is the primary source of uncertainty in isochron models (Hartmann, 2005). 2.Model   Page 4 of 45  107  108  109  110  111  112  113  114  115  116  117  118  119  120  121  122  123  124  125  2.1 Meteoroids traversing the atmosphere Crater populations are modeled assuming a power law distribution of projectiles at the top of the atmosphere and account for possible fragmentation and the dependence of mass and velocity on deceleration and ablation in the atmosphere. The size distribution of projectiles at the top of Mars’ atmosphere is adapted from the power-law fit to satellite observations of the annual flux of near-Earth objects colliding with the Earth for objects with diameters < 200 m (Brown et al., 2002): log (N) = co – do log (Dp), where N is the cumulative number of bolides colliding with the Earth per year, Dp is the meteoroid diameter in meters, co = 1.568, and do = 2.70. This SFD has been scaled by a factor 2.6, the nominal ratio of meteoroids at the top of Mars’ atmosphere relative to the Moon from the latest isochron model iteration of Hartmann (2005), The kinetic energy of an object entering the atmosphere is lost to deceleration and ablation. The decelerating force due to aerodynamic drag is (e.g., Baldwin and Sheaffer, 1971; (cid:1856)(cid:1874)(cid:1856)(cid:1872) (cid:3404) (cid:1829)(cid:3005) (cid:2025)(cid:3028)(cid:1827)(cid:1874) (cid:2870) 2(cid:1865) (cid:3397) (cid:1859)(cid:4666)(cid:1878)(cid:4667) sin (cid:2016) Chyba et al., 1993; Melosh, 1989) where ρa is the local density of the atmosphere. The parameters A, v, and m, are the cross- sectional area, the velocity, and the mass of the object respectively, g is the local gravitational (cid:1856)(cid:2016)(cid:1856)(cid:1872) (cid:3404) (cid:1859)(cid:4666)(cid:1878)(cid:4667) cos (cid:2016) (cid:1874) acceleration at altitude z, and θ is the angle of the trajectory measured from the local horizontal Eq. 2 assumes a flat surface geometry and precludes the possibility of projectiles skipping out of the atmosphere. The drag coefficient, CD, is ~1 in the continuum flow regime (Podolak et al., (2) (1) 1988). Heating of the projectile’s surface during entry is efficiently shed by ablation   Page 5 of 45  (3) (cid:1856)(cid:1865)(cid:1856)(cid:1872) (cid:3404) (cid:1829)(cid:3009) (cid:2025)(cid:3028)(cid:1827)(cid:1874) (cid:2871) 2(cid:2014) where CH is the heat transfer coefficient and ζ is the heat of ablation (Bronshten, 1983). To assess the conditions in which deceleration and ablation become substantial, we employ a heuristic model. Taking the gravity term to be negligible in Eq. (1) for the moment, and the approximation of an exponential density scale height for the atmosphere, ρa = ρo exp(−z/H), where z is the altitude, H is the scale height, and ρo is the atmospheric density at the surface, the (cid:1865)(cid:3033) (cid:3404) (cid:1865)(cid:3036) exp(cid:3427)(cid:3398)(cid:2026)(cid:3435)(cid:1874)(cid:3036)(cid:2870) (cid:3398) (cid:1874)(cid:3033)(cid:2870)(cid:3439)(cid:3431) final mass of the meteor, mf, can be related to the initial velocity, vi, and final velocity, vf, by where mi is the initial mass of the object at entry and σ = CH/(2ζCD) is the ablation coefficient. (4) Combining the above equations and solving (e.g., Davis, 1993), the dependence of mass and velocity on deceleration and ablation can be seen (Figure 1). Smaller, faster objects are more readily filtered out by the atmosphere. The condition for significant deceleration can be estimated to occur when the meteoroid mass is equivalent to the column of atmospheric mass it encounters, mi ~ ρoHA. Similarly, we can estimate the condition for substantial ablation. Ablation is more efficient at higher velocities as the ablative energy is proportional to the product of the drag force and the traversed distance, or, dm·σ-1  ρav2A·vdt. A transition to a high ablation regime will occur when the energy to ablate the entire meteoroid mass, m·σ-1, is equivalent to the energy required to traverse the atmosphere to the surface at a given velocity, ρoHA·v2, neglecting deceleration. Using these criteria, we can classify the projectiles based on velocity and mass (Figure 2). For large, slow projectiles, mf ~ mi and vf ~ vi, and the projectile will reach the surface relatively unchanged. For smaller, slow objects where m < ρoHA, and dm/dt~0, the final velocity decreases   Page 6 of 45  126  127  128  129  130  131  132  133  134  135  136  137  138  139  140  141  142  143  144  145  146  147  148  149  150  151  152  153  154  155  156  157  158  159  160  161  162  163  164  165  166  exponentially (cid:1874)(cid:3033)~(cid:1874)(cid:3036) exp (cid:4674)(cid:3398) (cid:3096)(cid:3290)(cid:3009)(cid:3002)(cid:3040) (cid:2929)(cid:2919)(cid:2924) (cid:3087)(cid:4675). Fast meteors, defined as having high initial velocities where 2 > σ-1, will experience significant ablation, mf < e-1· mi. Large, fast meteors will survive vi complete ablation if mσ-1 < ρoHA·v2, where the ablation will be limited by deceleration. These intermediate projectiles define the wedge shaped region in Figure 2a. Our taxonomic boundaries based on initial projectile masses and velocities are derived from our numeric modeling for a range of projectile masses and entry velocities (Figure 2). 2.2 Monte Carlo simulation We generated model crater populations using a Monte Carlo simulation employing the power- law distribution of meteoroids entering the top of the atmosphere from the previous section and the normalized distribution of entry velocities at Mars (Bland and Smith, 2000; Davis, 1993; (cid:1832)(cid:4666)(cid:1874)(cid:3036) (cid:4667) (cid:3404) 0.0231(cid:1874)(cid:3036) exp (cid:4680)(cid:3398) (cid:3436)(cid:1874)(cid:3036) (cid:3398) 1.806 8.874 (cid:3440)(cid:2870) (cid:4681) Flynn and McKay, 1990; Popova et al., 2003): where vi ≥ 5 km s-1 (the escape velocity of Mars) with a mean velocity of 10.2 km s-1. Following (4) Love and Brownlee (1991), the probability distribution of entry angles is taken to be sin(2θ), which has a maximum at 45° and drops to zero at 0° and 90°. A distribution of 5 groups of material types derived from the relative observed number of objects entering the terrestrial atmosphere is taken from Ceplecha et al. (1998). The material groups differ in their ability to penetrate the atmosphere, with the average ablation coefficient, , and bulk density, ρm, for each group listed in Table 1. Model results for ordinary chondrites are shown in Figure 3 for a single initial entry angle of 45°, demonstrating a good match with the results shown in Figure 1. The resulting crater volumes from the projectiles impacting the surface can be described by a scaling law that relates impact velocity and projectile and target characteristics (see Melosh   Page 7 of 45  167  168  169  170  171  172  173  174  175  176  177  178  179  180  181  182  183  184  185  186  187  188  (1989) for a review). In converting the crater SFD derived from the lunar surface to Mars, Hartmann (2005) assumes crater diameters, D, scale with impact energy as E0.43 and with gravity as g−0.17 to account for differences in mean impact velocity and gravitational acceleration between the two bodies. For small projectiles, the depth of excavation by an impact is small enough that lithostatic stresses are small relative to the characteristic yield stress of the regolith. In such cases the resulting transient crater diameter is determined by the yield strength, Ȳ, of the target material (“strength scaling”), and no longer scales with gravity. The transition from strength scaling to gravity scaling occurs when Ȳ ~ ρtgRp where Rp, the projectile radius, is taken to be the characteristic depth and ρt is the target density (Holsapple, 1993). For a regolith of density 2000 kg m-3 and yield strength of 0.1 – 1 MPa, the transition between strength and gravity dominated regimes occurs at D ~ 27 – 270 m. This implies that the 1 – 10 m scale fresh craters discovered by spacecraft over the last decade (Malin et al, 2006; Daubar et al., 2013) are predominately in the strength scaling regime. Strength scaling results in a reduction in crater volume relative to gravity scaled craters. As a consequence, the distribution of crater diameters would be expected to be shallower than that predicted by the isochrons of Hartmann (2005) at the smallest sizes, which assume all crater diameters scale with gravity. The cratering efficiency, πv = ρtV/mf, where ρt is the target density and V is the transient (cid:4678)(cid:2024)(cid:2870) (cid:3397) (cid:2024)(cid:2871)(cid:2870)(cid:2878)(cid:3091)(cid:2870) (cid:4679)(cid:2879) (cid:2871)(cid:3091)(cid:2870)(cid:2878)(cid:3091) crater volume, is proportional to (Holsapple, 1993): assuming impactor and target densities are the same, where μ is an empirical constant (ranging (4) from 1/3 – 2/3). The parameters π2 and π3 are dimensionless numbers describing the cratering efficiency for gravity scaling and strength scaling, respectively, where π2 = gRp /vf 2, the ratio of the lithostatic pressure at a depth equivalent to the projectile radius and the initial dynamic   Page 8 of 45  189  190  191  192  193  194  195  196  197  198  199  200  201  202  203  204  205  206  207  pressure generated by the impact, and π3 = Ȳ/ρtvf 2, the ratio of effective target yield strength to the initial dynamic pressure (Holsapple, 1993). For smaller projectiles, the π2 term becomes negligible and πv becomes constant with impactor size, depending only on velocity and the (5) material strength of the target. At large impactor sizes, π2 dominates and the cratering efficiency is then dependent not only on velocity, but impactor size. The expression for transient crater (cid:1848) (cid:3404) (cid:1837)(cid:2869) (cid:3436)(cid:1865)(cid:3033)(cid:2025)(cid:3047) (cid:3440) (cid:3437)(cid:2024)(cid:2870) (cid:3436)(cid:2025)(cid:3040)(cid:2025)(cid:3047) (cid:3440)(cid:2869)(cid:2871) (cid:3397) (cid:2024)(cid:2871)(cid:2870)(cid:2878)(cid:3091)(cid:2870) (cid:3441)(cid:2879) (cid:2871)(cid:3091)(cid:2870)(cid:2878)(cid:3091) volume is (Holsapple, 1993; Richardson et al., 2007; Dundas et al., 2010): where K1, like μ and Ȳ, are experimentally derived properties of the target material. Taking the 3, assuming the transient crater depth is roughly 1/3 its transient crater volume to be (1/24)πDt diameter, Dt (Melosh 1989; Schmidt and Housen, 1987), this equation can be expressed in terms (cid:1830) (cid:3404) 1.3(cid:1837) (cid:4593)(cid:1830) (cid:3436)(cid:2025)(cid:3040)(cid:2025)(cid:3047) (cid:3440)(cid:2869)(cid:2871) (cid:3437)(cid:2024)(cid:2870) (cid:3436)(cid:2025)(cid:3040)(cid:2025)(cid:3047) (cid:3440)(cid:2869)(cid:2871) (cid:3397) (cid:2024)(cid:2871)(cid:2870)(cid:2878)(cid:3091)(cid:2870) (cid:3441)(cid:2879) (cid:3091)(cid:2870)(cid:2878)(cid:3091) of final crater diameter where K' = (4K1)1/3 for a spherical meteoroid and the factor 1.3 is the ratio of the final rim-to-rim diameter and the transient crater diameter (Holsapple, 1993). We adopt values for Mars (6) consistent with dry desert alluvium of Ȳ = 65 kPa, μ = 0.41, K1 = 0.24, and ρt = 2000 kg m-3 (Holsapple, 1993; Holsapple and Housen, 2007) A minimum impact velocity threshold of 0.5 km s-1 is selected for the formation of explosive impact crater formation as projectiles with lower speeds are unlikely to generate a shock wave of sufficient magnitude to crush the target material and excavate an explosive crater (Popova et al. 2003). The definition of this velocity is not sharply defined and will depend on target material properties. Pressure wave velocities in competent basalt are typically ~4.5 – 6.5   Page 9 of 45  208  209  210  211  212  213  214  215  216  217  218  219  220  221  222  223  224  225  226  227  228  229  230  km s-1, with values an order-of-magnitude lower for loose, unconsolidated soils. Apollo seismic investigations found velocities to be ~0.1 – 0.3 km s-1 in the upper 100 m of the lunar regolith (Kovach and Watkins, 1973; Watkins and Kovach, 1973). The initial altitude of the modeled projectiles is 100 km. The gas density at this altitude is reduced by four orders-of-magnitude relative to the atmospheric density at the planet’s surface, and thus deceleration (Eq. 1) is reduced by four orders-of-magnitude relative to when the meteoroid is near the surface and is therefore initially negligible. This altitude also represents the approximate transition between the free molecular flow regime and the continuum regime (i.e. Knudsen number ~ 0.1, depending on the object size, for a mean free path of ~ 1 cm). 2.3 Aerodynamic breakup The fragmentation of meteoroids during their flight through the atmosphere likely influence the resulting SFD of smaller diameter craters, as more than half (56%) of the >200 fresh craters observed by HiRISE are comprised of clusters of individual craters (Daubar et al., 2013). Numerous additional examples of crater clusters have been observed on the surface of Mars (Popova et al., 2007) implying that the fragmentation of meteoroids that penetrate deep into the atmosphere is a common phenomenon on Mars. The aerodynamic breakup of a projectile is expected when the dynamic pressure experienced during entry, ρav2, exceeds the bulk strength, m, of the projectile. Bolides entering the terrestrial atmosphere display a wide range of strengths inferred from the broad range of observed breakup event altitudes with bulk strengths shown to be ~0.1 – 10 MPa (Ceplecha et al., 1998; Popova et al., 2011). The estimated bulk strengths are low compared to the tensile strength of recovered samples (typically 1-10 %), implying fractures and other zones of weakness within the bodies determine bulk strength rather than material composition (Popova et al., 2011). Dynamic pressures experienced by projectiles entering the   Page 10 of 45  231  232  233  234  235  236  237  238  239  240  241  242  243  244  245  246  247  248  249  250  251  252  253  Martian atmosphere will commonly exceed 1 MPa and therefore fragmentation should be expected, albeit at a lower altitude than observed in the terrestrial atmosphere. Larger, faster objects experience greater dynamic pressure during flight (Figure 4). Though there is no explicit size dependence on dynamic pressure, smaller objects begin to ablate and decelerate higher in the atmosphere, and therefore experience smaller peak loading at higher altitudes than equivalent larger objects. This implies a size dependence on fragmentation from entry dynamics alone. Further, it is anticipated that larger objects are inherently weaker as they likely contain a greater number of defects, and a power law relation based on statistical strength theory (Weibull, 1951) between bulk strength and sample mass is typically assumed in models (e.g. Artmieva and Shuvalov, 2001; Popova et al., 2003; Svetsov et al., 1995). However, recent analysis of meteorite fall observations in the terrestrial atmosphere suggests that, at least for stony objects, the relation between bulk strength and sample mass may be weaker (Popova et al., 2011). A possible size dependence on fragmentation was reported by Ivanov et al. (2010) based on a preliminary analysis of ~70 fresh craters, and is also exhibited by the 44 fresh craters observed by Daubar et al. (2013), 25 of which consist of two or more craters (57%). This percentage increases for craters with D > 5 m to 73% and decreases to 41% for craters with D < 5 m. 3. Results 3.1 Fragmentation The effects of fragmentation on the crater SFD is explored with the model. We include aerodynamic break up of meteoroids in the simulation allowing fragmentation to occur when the dynamic pressure exceeds a threshold bulk strength, m = ρav2 (Figure 5). Fragmentation is assumed to occur as a single event; however, multiple fragmentation events at different points in   Page 11 of 45  254  255  256  257  258  259  260  261  262  263  264  265  266  267  268  269  a meteoroid’s trajectory are sometimes observed during the flight of terrestrial fireballs. Catastrophic fragmentation (a point source type explosion with thousands of fragments) also occurs occasionally. We constrain the number of fragments to 2-100 and each fragment is followed individually to the termination of its flight in order to assess how fragmentation alters the resulting SFD. The number of fragments generated is assumed to follow a power-law probability: we adopt a power-law slope of -1.5 for our nominal model and adjust this value to explore the sensitivity of our results on this exponent. Decreasing the power-law slope results in a more uniform distribution of fragment numbers, while increasing the slope produces a distribution favoring fewer, larger, fragments. A cascade of fragment sizes is then generated by iteratively subtracting random mass fractions from the initial mass which typically results in a small number of larger fragments and many smaller fragments. We additionally explore selecting mass fractions with a uniform random distribution of fragment sizes. The model is run with 106 initial projectiles with a minimum allowable projectile diameter of 2 cm. A significant number of these events do not survive, but this minimum value was found to resolve the atmospheric downturn in the resulting crater SFD allowing the smallest permissible craters to form. The same bulk strength is selected for all the projectiles, and adjusted until the ratio of crater clusters to total impacts for D > 5 m is similar to the 73% 270  observed for the 44 craters reported by Duabar et al. (2013). We find that a bulk of strength of m 271  272  √2 bins of the resulting crater diameters. Both the observed and model craters are dominated by = 0.65 MPa best reproduces the observations, as shown in Figure 6, which plots histograms in 273  crater clusters at D > 5 m of approximately the same fraction. Effective diameters are estimated 274  275  276  for the clusters by Deff = (i Di 3)1/3, which represents the diameter of an equivalent crater due to a non-fragmented meteoroid (Malin et al., 2006; Ivanov et al., 2008).   Page 12 of 45  277  278  279  280  281  282  283  284  285  286  287  288  289  290  291  292  293  294  295  296  297  298  299  There is a downturn in the SFD of the craters observed by Daubar et al. (2013) at D ~ 4-5 m. However, the model craters continue to increase in number at smaller diameter bins down to the D = 0.24 m bin in Figure 6. The majority of craters have small diameters < 1m, and the fraction of clusters in each bin becomes negligible, with the total fraction of all clusters comprising only 4% of the total model crater population. This implies that the atmospheric downturn has not been observed in the fresh craters, and that a significant number of craters with D < 4-5 m have not been identified. The discovery of fresh craters relies on the detection of dark spots in CTX images in dusty regions. A swarm of fragments will disturb a larger surface area creating, in general, larger dark spots. The predominance of single craters at D < 5 m indicates that detection will be more challenging as dark spots become inherently smaller relative to D at these sizes, partly explaining the downturn in the observed SFD at 4-5 m. One might conclude that fragmentation, being only 4% of the events, may not be an important process in shaping the SFD. However, crater clusters begin to dominate the SFD at the very diameters that can be reliably detected by CTX, and therefore the effects on absolute ages derived from crater production functions are considered. Comparing the SFD of the model craters with craters produced by the same projectiles with fragmentation suppressed (Figure 7), a difference in age of 7% is obtained when fitting the Hartmann (2005) production function for D > 4 m. The influence of fragmentation on the crater SFD can be increased by taking the probability of the number of fragments generated when m is exceeded as random and uniform, instead of following the -1.5 slope power law (Figure 5b), and a uniform distribution of fragment masses. In this case we obtain a factor ~0.69 difference in age relative to the age derived if no fragmentation is allowed to occur (Figure 7b). This represents the case we have considered to have the largest likely influence on the SFD.   Page 13 of 45  300  301  302  303  304  305  306  307  308  309  310  311  312  313  314  315  316  317  318  319  320  321  322  The change in the crater SFD with fragmentation can be attributed to the mass dependence on ablation and deceleration and corresponding changes in crater scaling, which is not accounted for when calculating effective crater diameters, Deff. Individual fragments experience greater deceleration and ablation per unit volume than the larger parent meteoroid. If enough of the total mass is shifted into a high deceleration/ablation regime when the parent object is broken into individual fragments, Deff may no longer be representative of the crater generated by an equivalent unfragmented meteoroid. If the ratio of Deff to the actual D resulting unfragmented crater. In general, Deff /D < 1 when Deff ≲ 10 m for the nominal fragmentation from the unfragmented projectile is << 1, then Deff is a poor representation of the equivalent case, with the potential reliability of Deff decreasing with size (Figure 8). Scatter in Deff /D values events result in Deff /D values slightly greater than 1, predominately at Deff ≳ 10 m, due to the is largely due to variations in meteoroid velocities with lower Deff /D for faster objects. Some dependence of impactor size on gravity scaling where crater efficiency decreases with increasing impactor size. Consequently, the crater efficiency can be larger for the fragments than for the parent projectile is some cases. 3.2 Impact-crater isochrons The technique of dating the Martian surface using predicted crater SFDs for well-preserved surfaces using impact-crater isochrons has been developed over several decades going back to the early lunar exploration of the 1960s (e.g. Hartmann, 1966; Neukum and Wise, 1976; Hartmann, 1999; Neukum and Ivanov, 1994; Hartmann and Neukum, 2001; Hartmann, 2005). Radiometric and exposure ages from Apollo and Luna samples, correlated with crater populations, have anchored the lunar cratering chronology. Scaling lunar isochrons for Mars to account for the ratio of meteoroids at the top of the martian atmosphere relative to the Moon and   Page 14 of 45  323  324  325  326  327  328  329  330  331  332  333  334  335  336  337  338  339  340  341  342  343  344  345  the difference in gravity and average impact velocity provides a means of understanding the absolute timescale of major geological and geophysical processes on Mars with order of magnitude accuracy. Degradation of craters over time will result in a departure in the crater population from the isochrons. An inflection at D~1.5 km separates the isochrons into two branches with different power-law exponents. This inflection, initially observed in the lunar craters, divides the crater distribution into a shallow branch with a slope of ~2 and a steep branch of ~3 on a log-differential plot. In his latest iteration of Mars isochrons, Hartmann (2005) included an atmospheric model (Popova et al., 2003) to encapsulate the effects of atmospheric entry of meteors resulting in curvature in the isochrons at sub-km crater diameters. During impact, ejecta traveling large distances from a crater are capable of generating isolated secondary craters that could potentially be mistaken for primary craters. What fraction of observed crater populations are primary versus secondary, and the consequence of unidentified secondary craters in dating surfaces, is a matter of debate. McEwen et al. (2005), and McEwen and Bierhaus (2006), argue that small crater populations are dominated by secondary craters, rendering smaller craters unreliable for dating. Hartmann (2005, 2007) however points out that in formulating isochrons, no attempt is made to exclude all secondaries and contends that the accumulation of secondary craters are not “contamination” but rather part of the signal. A population of primary plus secondary craters will have a steeper SFD power law slope than a population devoid of secondary craters, and McEwen et al. (2005) has proposed that the increase in slope of the steep branch of crater SFDs is the result of secondaries. Our approach of using the observed projectile production function at the top of the terrestrial atmosphere provides an opportunity to independently test the veracity of lunar and martian isochrons. We can also constrain the potential influence of secondary craters, as our   Page 15 of 45  346  347  348  349  350  351  352  353  354  355  356  357  358  359  360  361  362  363  364  365  366  367  368  Monte Carlo model only produces primary crater populations. If secondary craters are a substantial fraction of crater populations, then our model crater SFD should deviate in both slope and age from the isochrons. Additionally, since we use the present-day flux of projectiles, any significant deviation in age estimates can be attributed to a variation in impact rate over time. We test our model against crater counts conducted on the proximal ejecta of the crater Zunil. The crater Zunil is selected because it is likely the last D = 10 km scale crater to form on Mars (McEwen et al., 2005; Hartmann et al., 2010) with an estimated age of ~1 Ma based on the isochrons of Hartmann (2005), so the distribution of small craters on its ejecta are likely to be predominately primary. Crater counts were conducted on a ~5 km2 area north of the crater rim (Figure 9). The resulting crater SFD is consistent with an age ~1 Ma, similar to counts conducted by Hartmann et al. (2010). We then modeled a population of craters over a 1 Myr period of time for the same surface area as our crater counts. The resulting SFD from both the counts and the Monte Carlo model overlap for bins D > 2 m and fall near the 1 Myr isochron of Hartmann (2005) with a similar slope on a log-differential plot (Figure 9b). Cumulative crater frequencies are also shown (Figure 9c), and yield effective crater retention ages of 0.903  0.077 Ma and 0.820  0.071 Ma for the counts and the model respectively. Effective diameters from the crater counts were determined by assuming any craters within 40 m of each other were part of a cluster. upward revision of ages to 1.03(cid:2879)(cid:2868).(cid:2868)(cid:2875)(cid:2875) (cid:2878)(cid:2868).(cid:2868)(cid:2875)(cid:2876) Ma and 0.966  0.074 Ma respectively (Figure 9d). As For comparison, the SFD of the individual observed craters and model craters result in a modest crater densities increase on a surface, the identification of individual clusters becomes more difficult, making age determinations more challenging for smaller craters on older surfaces. We also conducted a similar Monte Carlo comparison with the 44 fresh CTX-CTX detected craters. A crater population is modeled for a corresponding area and time represented by   Page 16 of 45  369  370  371  372  373  374  375  376  377  378  379  380  381  382  383  384  385  386  387  388  389  390  391  the CTX craters. The model results in a SFD overlapping with the 1 yr isochron of Hartmann (2005). However, the fresh craters fall below the isochron as noted by Daubar et al. (2013). The largest crater, D = 33.8 m, is consistent with a single D ~ 30 m crater predicted to form annually for the scaled observation area. However, smaller diameter bins are inexplicably under-populated (Figures 9,10) Note that the atmospheric downturn is expected to occur at much smaller diameters (sub-meter), and fragmentation will not substantially deplete the SFD at these smaller sizes (Figure 6b). A lack of a contribution from secondaries to the isochrons also cannot explain the discrepancy, as our model predicts a primary SFD with a similar slope and age as the isochron. Fresh craters discovered using MOC by Malin et al. (2006) also have an apparent deficit of craters at all bin diameters when scaled to the 1-yr isochron of Hartmann (2005). The cumulative distribution of fresh craters, including both CTX-CTX and MOC detections, is shown in Figure 10. The MOC SFD has been scaled to an equivalent time/area for comparison. The MOC camera, with a lower resolution, did not detect any craters below D ~ 10 m and detected fewer craters in all bin sizes that overlap with the CTX craters. All bin sizes (except for the largest CTX crater) fall below the isochron, and this discrepancy is greater in the lower resolution dataset. This indicates that observational bias may be the best explanation for the difference between the predicted SFD and the observed SFDs. If this is the case, then over time, as additional, larger diameter craters are identified by CTX/HiRISE, the SFD will start to overlap with the isochron at larger diameter bins since the larger craters are more likely to create detectable dark spots on the surface. This would then suggest that fresh craters were undercounted by MOC at all diameters observed. 3.3 Lunar crater counts   Page 17 of 45  392  393  394  395  396  397  398  399  400  401  402  403  404  405  406  407  408  409  410  411  412  413  414  As an additional test, we compare results from the model with crater counts of small craters on the moon (Figure 11). The ejecta of North Ray crater was selected for its relatively young age, which has been constrained by cosmic ray exposure ages of Apollo 16 samples to be ~50 Ma (Arvidson et al., 1975). Our small (0.1 km2) study area only contains craters D  22 m. Applying the cratering chronology of Neukum et al. (2001), an age of 58.9  11 Ma is derived from the resulting cumulative SFD, consistent with the age derived by Hiesinger et al., (2012) using a larger overlapping area. A population of projectiles is then generated, again using the size distribution of small near-Earth objects colliding with the terrestrial atmosphere (Brown et al., 2002) and the velocity distribution of Marchi et al. (2009), for a period of 58.9 Ma. This results in a total of 42895 craters for the 0.1 km2 surface area using projectiles >0.02 m diameter. Crater scaling parameters appropriate for the lunar regolith are employed assuming a regolith strength of 10 kPa, K1 = 0.132, and ρt = 1500 kg m-3 (Holsapple, 1983; Holsapple and Housen, 2007; Vasavada et al., 2012). The resulting SFD of the Monte Carol model craters results in a similar age of 57.2  11 Ma (Figure 11). This demonstrates the usefulness of our approach, as we are able to reproduce a ~50 – 60 Ma surface age on the Moon using a limited area containing only small craters. If craters in this area were predominately secondary craters at these diameters, then our Monte Carlo model should have predicted a much younger surface age. A similar approach was taken by Ivanov (2006) to demonstrate that small craters on the Moon must be predominately primaries on young surfaces ( 100 Ma), consistent with our model. Our ability to reproduce isochron ages using a small area with a limited range of diameters not only validates the applicability of small craters for dating young < 60 Ma surfaces, but also indicates that the cratering rate has on average been constant over this period (Ivanov,   Page 18 of 45  415  416  417  418  419  420  421  422  423  424  425  426  427  428  429  430  431  432  433  434  435  436  437  2006). How secondary craters alter the SFD and whether such small craters can be reliably used for age dating surfaces in all cases is beyond the scope of this work, but our Monte Carlo approach does independently confirm that crater-count chronometry systems can accurately date the youngest surfaces on both the Moon and Mars. 3.4 Sensitivity analysis The sensitivity of our results to changes in model parameters is evaluated (Table 2). Our nominal case results in a good fit to the Hartmann (2005) annual isochron for D ≥ 20 m where the model produces a SFD that results in an age of 1.03  0.017 yr (Figure 12). The Monte Carlo model does not produce as close a match at smaller sizes where the SFD is best fit with an age of 0.734  0.0018 yr for D = 4 - 20 m. This may result from a greater atmospheric influence on the meteoroids at smaller sizes than predicted by the atmospheric model of Popova et al. (2003) that is incorporated into the isochron, or could be due to Hartmann (2005) assuming gravity scaling only which would also result in less downturn in the isochron as crater scaling transitions into the strength regime (Williams et al. 2010). The mass of the atmospheric column that a meteoroid must traverse will depend on the location of the event as the elevation of the surface can vary substantially. The highest elevation, the summit of Olympus Mons, is ~21 km above datum, while at the other extreme, the interior of Hellas Basin, is ~8 km below. This represents a range of over 2.5 atmospheric scale heights. One of the most striking characteristics of the topography is that its distribution is bimodal due to the north-south dichotomy in crust thickness (Smith et al., 1999). The more heavily cratered highlands in the southern hemisphere have an average elevation of ~1.5 km while the distribution of lowland elevations narrowly peak 5.5 km lower than the highlands (Aharonson et al., 2001). This translates to an increase in average atmospheric density at the surface by a   Page 19 of 45  438  439  440  441  442  443  444  445  446  447  448  449  450  451  452  453  454  455  456  457  458  459  460  factor 1.6 between the southern highlands, ρo = 0.0135 kg m-3, and the northern lowlands, ρo = 0.0222 kg m-3. Taking the same model projectiles and generating craters for the two atmosphere densities representative of the average elevations of the highlands and lowlands, two crater SFDs are generated for comparison (Figure 12b). Craters D ≥ 20 m result in ages 1.14 yr and 0.993 yr respectively, bracketing the nominal model results. This is a change of +10.7% and -3.6%. Fitting the smaller diameters, D = 4 – 20 m, results in ages 0.92 yr and 0.69 yr, a change of 25.3% and -6.0%. The greater sensitivity at the small crater diameters is to be expected as smaller projectiles are more sensitive to atmospheric conditions. Increasing and decreasing the ablation coefficients, , has a similar effect as changing the atmospheric surface density, as an increase in the rate of mass loss results in smaller objects which in turn results in greater deceleration. Models for the same projectiles with ablation coefficients half the nominal values and double the nominal values were run (Figure 12c). The derived ages for craters D ≥ 20 m increased by 13.5% and decreased by 18.2% for  ½ and 2 respectively, and similarly for craters D = 4 – 20 m increased by 28.9% and decreased by 30.4%. The model assumes that the different taxonomic types of fireballs distinguished from the larger Earth-entering meteoroids (Ceplecha et al., 1998; Popova et al., 2003) have the same fractional distribution at Mars. These types of fireballs differ in their ability to penetrate the atmosphere and therefore if the proposed fractions of projectiles are substantially different on average from that observed in the terrestrial atmosphere (Ceplecha et al., 1998), the results will change. For example, a greater percentage of higher density objects will experience less deceleration and result in larger craters for a given population of projectiles. The smallest craters in the simulation, cm-scale, are formed exclusively by iron objects due to their high density   Page 20 of 45  461  462  463  464  465  466  467  468  469  470  471  472  473  474  475  476  477  478  479  480  481  482  483  (7800 kg m-3). Iron meteors however are a small fraction of the total population (3%). The majority of the objects are Ordinary Chondrites (29%), Carbonaceous Chondrites (33%), and Cometary Material (26%). None of the Soft Cometary Material objects, with a density of 270 kg m-3, survive passage through the martian atmosphere. To demonstrate the sensitivity of the model results to different projectile compositions, we run the simulation assuming pure compositions of the three predominant types: Ordinary Chondrites, Carbonaceous Chondrites, and Cometary Material (Figure 12d). The Ordinary Chondrites, with a density greater than the average combined density, results in a factor >2 increase in age, while the Carbonaceous Chondrites reduce the age by 37.3% for D = 4 – 20 m and 15.3% for D ≥ 20 m. A meteor population comprised entirely of Cometary Material results in substantially younger age estimates, 0.0466 and 0.15 yr for the small and large diameters respectively, younger than the ages derived from either the MOC or CTX-CTX craters, with a slope that deviates from the isochron increasingly at smaller D. The crater counts from Zunil in Figure 9c are shown in Figure 12d for comparison, scaled to the 1yr isochron, showing a slope consistent with a population of objects with average density similar to the nominal values assumed and inconsistent with a substantial fraction of icy, low-density impactors. The crater scaling becomes increasingly sensitive to the target material properties as projectile diameters decrease with weaker material producing a larger crater for a given projectile. Details of the target properties, particularly the effective strength, may not be well constrained and can vary by location, feature, or geologic unit. Dundas et al. (2010) has demonstrated how this can complicate geologic interpretation as surfaces of the same age, but differing material properties, may yield different ages. We explore the sensitivity of the model to a range of target properties. Figure 12e shows the results for the same population of projectiles   Page 21 of 45  484  485  486  487  488  489  490  491  492  493  494  495  496  497  498  499  500  501  502  503  504  505  506  for four different targets. We take the lunar regolith of section 3.3 as a lower bound on target strength representing a low density, dry, moderately porous soil, and soft rock and hard rock from Holsapple (1993) for two targets with greater strength and density than our nominal Mars regolith. For the rock materials, μ = 0.55, the higher value reflecting lower porosity and therefore less energy dissipation (Holsapple and Housen, 2007), and K1 = 0.20. The densities are assumed to be 2250 and 2500 kg m-3 and the material strengths 7.6 and 18.0 MPa for the soft rock and hard rock respectively (Holsapple, 1993; Richardson et al., 2007). The hard rock gives an upper bound representative of competent, young basalt that lacks a substantial regolith cover. increases with decreasing D. The populations begin to converge (cid:1830) ≳ 100 m. For D > 20 m, the The discrepancy in ages between resulting crater SFDs of the different target materials lunar regolith model yields a similar age to the nominal model however for smaller craters, the weaker target material results in larger craters and a corresponding increase in estimated age, 0.903 versus 0.734 yr. The stronger rock materials yield smaller craters and the decrease in estimated ages with the rock materials results in SFDs that overlap with the observed CTX-CTX and MOC fresh craters (Figure 12e). For our model to be consistent with the SFD of the fresh craters, the majority of the craters would have to have formed in relatively strong, competent rock. Many of the craters are on Amazonian units in the Tharsis region which is dominated by Amazonian lava flows, however, the largest CTX crater, D = 33.8 m, formed on the southeast flank of Pavonis Mons on a relatively young Amazonian volcanic unit (Scott et al., 1998). This crater is consistent with Hartmann 2005 which predicts an annual D ~ 30 m diameter crater and its presence in the fresh crater population cannot be attributed to the crater having formed in as a statistical outlier. However, it is incorrect to interpret the √(cid:1866) error bars in comparison to weaker target material relative to the other fresh craters. It is tempting to dismiss the single crater Page 22 of 45    507  508  509  510  511  512  513  514  515  516  517  518  519  520  521  522  523  524  525  526  527  528  529  model isochrons as these error bars indicate the 1- range of likely observations not models given the actual number seen. If we wait an additional year, we may observe no additional events of this size. However the real flux of events of this size clearly cannot be 0 given that a single event has been observed and so models with exceedingly low formation rates must be formally rejected (Aharonson, 2007). If continued observations of fresh CTX-CTX craters verify that the current observed SFD is indeed an accurate reflection of the present-day impact crater population, then our model demonstrates that it is possible to constrain the fraction of impactor material types and average target material properties for Mars. For example, a population of impactors with average densities similar or higher than that observed for terrestrial fireballs could be ruled out as they are not consistent with the current CTX-CTX observations. 4. Discussion The fact that the slope of the Hartmann (2005) isochrons are the same as that derived in our study using the present-day annual flux of terrestrial fireballs implies that the isochron system does not incorporate many unrecognized secondary craters. This is quite surprising considering McEwen et al. (2005) identify ~107 secondary craters generated by the Zunil crater forming impact event at distances up to 3500 km away from the primary crater. This implies that small secondaries should be abundant on the Martian surface. We specifically selected young surfaces to conduct counts that likely reflect a primary crater population so it is not surprising that our counts have similar slopes to our model. The isochron system however, is derived from counts conducted on lunar maria with average ages of 3.5 Ga, surfaces that have had ample time to accumulate secondaries. Obvious secondaries were not included in those counts and presumably only unrecognizable distant secondaries were included. It remains unclear how   Page 23 of 45  530  531  532  533  534  535  536  537  538  539  540  541  542  543  544  545  546  547  548  549  550  551  552  secondaries are influencing the crater statistics as we do not identify a secondary signal. We have not attempted a systematic survey to identify a range of crater diameters and surface ages where secondaries are influencing the crater SFD, which is beyond the scope of this study, and caution in using small craters remains warranted as this remains an unresolved issue. However, we do demonstrate that for young surfaces that are not likely contaminated by secondaries, useful ages can be derived from crater counts using small craters. 5. Conclusions Taking the size distribution of the observed annual flux of terrestrial fireballs, scaling the number by 2.6 for Mars, the nominal value of Hartmann (2005), and appropriate velocity distributions for Mars and the Moon, we are able to generate crater populations with a remarkably similar SFD to that predicted by the isochrons of Hartmann (2005) for Mars and Neukum et al. (2010) for the Moon. Our Monte Carlo model generates a population of primary craters, implying that the isochrons do not contain a significant secondary crater signal as large numbers of secondaries would result in a steeper isochron slope than predicted by our model. This further implies that the inflection from the shallow branch to the steep branch is not attributable to the addition of unidentified secondary craters. Crater counts conducted on the ejecta blanket of Zunil and North Ray crater yield SFDs with similar slopes and ages as our model— ~1 Ma and ~58 Ma, respectively, for Mars and the Moon—demonstrating that the cratering rate on average has been fairly constant on these bodies over these time periods. We find that the details of the atmospheric model for Mars do not substantially alter the results. The Hartmann (2005) isochrons include an atmospheric model (Popova et al., 2003), and differences between the lowlands and highlands do not result in significant deviations from the Hartmann isochron, nor do variations in assumed ablation coefficients. Fragmentation did not   Page 24 of 45  553  554  555  556  557  558  559  560  561  562  563  564  565  566  567  568  569  570  571  572  573  574  significantly alter the model results and smaller projectiles (D < 5 m) typically were decelerated enough to avoid fragmentation. The atmospheric downturn predicted by our model occurs at D ~ 10-20 cm, corresponding to projectile masses roughly equivalent to the mass of the atmospheric column the objects are traversing. This is well below the downturn observed in the fresh craters detected by CTX at D = 4-5 m (Daubar et al., 2013). The downturn in the CTX SFD is likely due to detection limits of CTX. The largest fresh crater, D = 33.8 m, observed, is consistent with our model and the prediction of the Hartmann isochron of one annual D ~ 30 m crater; however, fewer craters were observed than predicted by the isochron at smaller diameters. It is possible that D < 30 m craters are underrepresented in the observed CTX craters and continued observation, allowing for larger craters to form over a longer observation period, may potentially resolve this question. Acknowledgements J.-P. W. and A. P. were supported by a NASA Mars Data Analysis Program Grant Number NNX11AQ64G. References Aharonson, O. (2007) The modern impact cratering flux at the surface of Mars, Lunar Planet. Sci. Conf., 38th, #2288. Aharonson, O., M. T. Zuber, and D. H. Rothman (2001) Statistics of Mars' topography from the Mars Orbiter Laser Altimeter' Slopes, correlations, and physical Models, J. Geophys. Res., 106, 23,723–23,735. Artmieva, N. A., and V. V. Shuvalov (2001) Motion of a fragmented meteoroid through the planetary atmosphere, J. Geophys. Res., 106, 3297–3309.   Page 25 of 45  575  576  577  578  579  580  581  582  583  584  585  586  587  588  589  590  591  592  593  594  595  596  597  Arvidson, R., G. Crozaz, R. J. Drozd, C. M. Hohenberg, C. J. Morgan (1975) Cosmic ray exposure ages of features and events at the Apollo landing sites, Moon, 13, 259–276. Baldwin, B., and Y. Sheaffer (1971) Ablation and breakup of large meteoroids during atmospheric entry, J. Geophys. Res., 76, 4653–4668. Basilevsky, A. T., G. Neukum, S. C. Werner, A. Dumke, S. van Gasselt, T. Kneissl, W. Zuschneid, D. Rommel, L. Wendt, M. Chapman, J. W. Head, and R. Greeley (2009) Episodes of floods in MangalaValles, Mars, from the analysis of HRSC, MOC and THEMIS images, Planet. Space Sci., 57, 917–943. Bland, P. A. and T. B. Smith (2000) Meteorite accumulations on Mars, Icarus, 144, 21–26. Bronshten, V. A. (1983) Physics of meteoric phenomena, Dordrecht, Reidel. 356 p. Brown, P., R. E. Spalding, D. O. ReVelle, E. Tagliaferri, and S. P. Worden (2002) The flux of small near-Earth objects colliding with the Earth, Nature, 420, 294–296. Burr, D. M., J. A. Grier, L. P. Keszthelyi, A. S. McEwen (2002) Repeated aqueous flooding from the Cerberus Fossae: Evidence for very recently extant, deep groundwater on Mars, Icarus, 159, 53–73. Byrne S., C. M. Dundas, M. R. Kennedy, M. T. Mellon, A. S. McEwen, S. C. Cull, I. J. Daubar, D. E. Shean, K. D. Seelos, S. L. Murchie, B. A. Cantor, R. E. Arvidson, K. S. Edgett, A. Reufer, N. Thomas, T. N. Harrison, L. V. Posiolova, F. P. Seelos (2009) Distribution of Mid-Latitude Ground Ice on Mars from New Impact Craters, Science, 25, 1674–1676. Ceplecha, Z., J. Borovička, W. G. Elford, D. O. Revelle, R. L. Hawkes, V. Porubčan, M. Šimek (1998) Meteor Phenomena and Bodies, Space Sci. Rev., 84, 327–471. Chyba, C. F., P. J. Thomas, and K. J. Zahnle (1993), The 1908 Tunguska explosion - Atmospheric disruption of a stony asteroid, Nature, 361, 40–44.   Page 26 of 45  598  599  600  601  602  603  604  605  606  607  608  609  610  611  612  613  614  615  616  617  618  619  Daubar, I. J., and A. S. McEwen (2009) Depth to diameter ratios of recent primary impact craters on Mars, Lunar Planet. Sci. Conf., 40th, #2419. Daubar, I. J., S. Byrne, A. S. McEwen, and M. Kennedy (2010) New Martian Impact Events: Effects on Atmospheric Breakup on Statistics, 1st Planetary Cratering Consortium. Daubar, I. J., A. S. McEwen, S. Byrne, C. M. Dundas, A. L. Keske, G. L. Amaya, M. Kennedy, and M. S. Robinson (2011) New craters on Mars and the Moon, Lunar Planet. Sci. Conf., 42nd, #2232. Daubar I. J., A. S. McEwen, S. Byrne, M. R. Kennedy, and B. Ivanov (2013) The current martian cratering rate, Icarus, in press Davis, P. M. (1993) Meteoroid impacts as seismic sources on Mars, Icarus, 105, 469–478. Dundas, C. M., and S. Byrne (2010) Modeling Sublimation of Ice Exposed by Recent Impacts in the Martian Mid-Latitudes, Icarus, 206, 716–728. properties in the cratering record of young platy‐ridged lava on Mars, Geophys. Res. Lett., Dundas C. M., L. P. Keszthelyi, V. J. Bray, and A. S. McEwen (2010) Role of material 37, L12203, doi:10.1029/2010GL042869. Flynn, G. J., and D. S. McKay (1990) An assessment of the meteoritic contibution to the martian soil, J. Geophys. Res., 95, 14,497–14,509. Hartmann, W. K. (1966), Martian cratering, Icarus, 5, 565-576. Hartmann, W. K. (1999), Martian cratering VI: crater count isochrons and evidence for recent volcanism from Mars Global Surveyor, Meteor, Planet. Sci., 34, 167-177. Hartmann, W. K. (2005), Martian cratering 8: Isochron refinement and the chronology of Mars, Icarus, 174, 294–320.   Page 27 of 45  620  621  622  623  624  625  626  627  628  629  630  631  632  633  634  635  636  637  638  639  640  641  642  643  Hartmann, W. K. (2007), Martian cratering 9: toward resolution of the controversy about small craters, Icarus, 186, 274–278. Hartmann, W. K. and D. C. Berman (2000), Elysium Planitia lava flows: crater count chronology and geological implications, J.Geophys. Res., 105, 15,011–15,025. Hartmann,W. K., G. Neukum, (2001) Cratering chronology and evolution of Mars. In: Altwegg, K., Ehrenfreund, P., Geiss, J., Huebner, W.F. (Eds.), Composition and Origin of Cometary Materials, Kluwer Academic, The Netherlands, pp. 165–194. Hartmann, W. K., C. Quantin, S. C. Werner, and O. Popova (2010) Do young martian ray craters have ages consistent with the crater count system?, Icarus, 208, 621–635. Hiesinger H., C. H. van der Bogert, J. H. Pasckert, L. Funcke, L. Giacomini, L. R. Ostrach, and M. S. Robinson (2012) How old are young lunar craters?, J. Geophys. Res., 117, E00H10, doi:10.1029/2011JE003935. Holsapple, K. A. (1993), The scaling of impact processes in planetary science, Ann. Rev. Earth Planet. Sci., 21, 333–373. Holsapple , K. A. and K. R. Housen (2007) A crater and its ejecta: An interpretation of Deep Impact, Icarus, 345–356. Ivanov, B. A. (2006) Earth/Moon impact rate comparison: Searching constraints for lunar secondary/primary cratering proportion, Icarus, 183, 504–507. Ivanov, B., H. J. Melosh, A. S. McEwen, and the HiRISE team (2008) Small impact crater clusters in high resolution HiRISE images, Lunar Planet. Sci. Conf., 39th, #1221. Ivanov, B. A., H. J. Melosh, A. S. McEwen, and the HiRISE team (2009) Small impact crater clusters in high resolution HiRISE images – II, Lunar Planet. Sci. Conf.,40th, #1410. Ivanov, B. A., H. J. Melosh, A. S. McEwen, and the HiRISE team (2010) Small impact crater clusters in high resolution HiRISE images – III, Lunar Planet. Sci. Conf.,41th, #2020.   Page 28 of 45  644  645  646  647  648  649  650  651  652  653  654  655  656  657  658  659  660  661  662  663  664  665  666  Kadish, S. J., J. W. Head, R. L. Parsons, D. R. Marchant, (2008) The Ascraeus Mons fan-shaped deposit: Volcano–ice interactions and the climatic implications of cold-based tropical mountain glaciation, Icarus, 197, 84–109. Kennedy, M. R., and M. C. Malin (2009) 100 new impact crater sites found on Mars, AGU Fall Meeting, #P43D-1455 Kneissel, T., S. van Gasselt, and G. Neukum (2011) Map-projection-independent crater size- frequency determination in GIS environments–New software tool for ArcGIS, Planet. Space Sci., 59, 1243–1254. Kovach, R. L., and J. S. Watkins (1973) The velocity structure of the Lunar crust, The Moon, 7, 63–75. Lanagan, P. D., A. S. McEwen, L. P. Keszthelyi,, T. Thordarson (2001), Rootless cones on Mars indicating the presence of shallow equatorial ground ice in recent times, Geophys. Res. Lett., 28, 2365–2367. Love, S. G., and D. E. Brownlee (1991) Heating and thermal transformation of micrometeoroids entering the Earth’s atmosphere, Icarus, 89, 26–43. Malin, M. C., and K. S. Edgett (2000a) Evidence for recent groundwater seepage and surface runoff on Mars, Science, 288, 2330–2335. Malin, M. C., and K. S. Edgett (2000b) Sedimentary rocks of early Mars, Science, 290, 1927– 1937. Malin, M. C., and K. S. Edgett (2001) Mars Global Surveyor Mars Orbiter Camera: Interplanetary cruise through primary mission, J. Geophys. Res., 106, 23,429–23,570. Malin, M. C., G. E. Danielson, A. P. Ingersoll, H. Masursky, J. Veverka, M. A. Ravine, and T. A. Soulanille (1992) Mars orbiter camera, J. Geophys. Res., 97, 7699–7718.   Page 29 of 45  667  668  669  670  671  672  673  674  675  676  677  678  679  680  681  682  683  684  685  686  687  688  Malin, M. C., K. S. Edgett, L. Posiolova, S. McColley, and E. Noe Dobrea (2006) Present impact cratering rate and the contemporary gully activity on Mars: Results of the mars global surveyor extended mission, Science, 314, 1573–1557. Malin, M. C., J. F. Bell III, B. A. Cantor, M. A. Caplinger, W. M. Calvin, R. T. Clancy, K. S. Edgett, L. Edwards, R. M. Haberle, P. B. James, S. W. Lee, M. A. Ravine, P. C. Thomas, and M. J. Wolff (2007) Context Camera Investigation on board the Mars Reconnaissance Orbiter, J. Geophys. Res., 112, E05S04, doi:10.1029/2006JE002808. Mangold, N. (2003) Geomorphic analysis of lobate debris aprons on Mars at Mars Orbiter Camera scale: Evidence for ice sublimation initiated by fractures, J. Geophys. Res., 108, 10.1029/2002JE001885, pp. GDS 2-1. Marquez, A., Fernandez, C., Anguita, F., Farelo, A., Anguita, J., de la Casa, M.-A. (2004) New evidence for a volcanically, tectonically, and climatically active Mars, Icarus, 172, 573–581. Marchi , S., S. Mottola, G. Cremonese, M. Massironi, and E. Martellato (2009) A new chronology for the Moon and Mercury, Astron. J., 137, 4936–4948. McEwen, A. S., and E. B. Bierhaus (2006) The importance of secondary cratering to age constraints on planetary surfaces, Annu. Rev. Earth Planet. Sci., 34, 535–567. McEwen, A. S., B. S. Preblich, E. P. Turtle, N. A. Artemieva, M. P. Golombek, M. Hurst, R. L. Kirk, D. M. Burr, and P. R. Christensen (2005), The rayed crater Zunil and interpretations of small impact craters on Mars, Icarus, 176, 351–381, doi:10.1016/j.icarus.2005.02.009. McEwen, A. S., E. M. Eliason, J. W. Bergstrom, N. T. Bridges, C. J. Hansen, W. A. Delamere, J. A. Grant, V. C. Gulick, K. E. Herkenhoff, L. Keszthelyi, R. L. Kirk, M. T. Mellon, S. W. Squyres, N. Thomas, C. M. Weitz (2007a) Mars Reconnaissance Orbiter's High Resolution   Page 30 of 45  689  690  691  692  693  694  695  696  697  698  699  700  701  702  703  704  705  706  707  708  709  710  Imaging Science Experiment (HiRISE), J. Geophys. Res., 112, E05S02, doi:10.1029/2005JE002605. McEwen, A. S., J. A. Grant, L. L. Tornabene, S. Byrne, and K. E. Herkenhoff (2007b) HiRISE observations of small impact craters on Mars, Lunar Planet. Sci. Conf., 38th, #2009. McEwen, A. S., L. L. Tornabene, and the HiRISE Team (2007c) Modern Mars: HiRISE observations of small, recent impact craters on Mars, 7th Int. Conf. Mars, LPI Contribution No. 1353, #3086. Melosh, H. J. (1989), Impact Cratering: A Geologic Process, 245 pp., Oxford Univ. Press, New York. Michael, G. G., and G. Neukum (2009) Planetary surface dating from crater size-frequency distribution measurements: Partial resurfacing events and statistical age uncertainty, Earth Planet. Sci. Lett., 294, 223–229. Neukum, G. and D. Wise, (1976) Mars: a standard crater curve and possible new time scale, Science, 194, 1381–1387. Neukum, G., and B. A. Ivanov (1994), Crater Size Distributions and Impact Probabilities on Earth from Lunar, Terrestrial-planet, and Asteroid Cratering Data, in Hazards Due to Comets and Asteroids, edited by T. Gehrels, M. S. Matthews, and A. M. Schumann, pp. 359–416. Neukum, G., B. A. Ivanov, and W. K. Hartmann (2001) Cratering records in the inner solar system in relation to the Lunar reference system, Space Sci. Rev., 96, 55–86. Podolak, M., J. B. Pollack, and R. T. Reynolds (1988) Interactions of planetesimals with protoplanetary atmospheres, Icarus, 73, 163–179. Popova, O., I. Nemtchinov, and W. K. Hartmann (2003) Bolides in the present and past martian atmosphere and effects on cratering processes, Meteorit. Planet. Sci., 38, 905–925.   Page 31 of 45  711  712  713  714  715  716  717  718  719  720  721  722  723  724  725  726  727  728  729  730  731  Popova, O., W. K. Hartmann, I. Nemtchinov, D. C. Richardson, and D. C. Berman (2007), Crater clusters on Mars: shedding light on martian ejecta launch conditions, Icarus, 190, 50– 73. Popova, O., J. Borovička, W. K. Hartmann, P. Spurný, E. Gnos, I. Nemtchinov, and J. M. Trigo- Rodríguez (2011) Very low strengths of interplanetary meteoroids and small asteroids, Meteorit. Planet. Sci., 46, 1525–1550. Quantin, C., N. Mangold, W. K. Hartmann, and P. Allemand (2007), Possible long-term decline in impact rates 1. Martian geological data, Icarus, 186, 1–10. Reiss, D., S. van Gasselt, G. Neukum, and R. Jaumann (2004) Absolute dune ages and implications for the time of formation of gullies in Nirgal Vallis, Mars, J. Geophys. Res. 109, E06007, doi:10.1029/2004JE002251. Richardson, J. E., H. J. Melosh, C. M. Lisse, and B. Carcich (2007) A ballistic analysis of the Deep Impact ejecta plume: Determining comet Tempel 1’s gravity, mass, and density, Icarus, 190, 357–390. Schmidt, R. M., and K. R. Housen (1987), Some recent advances in the scaling of impact and explosion cratering , Int. J. Impact Eng., 5, 543–560. Schon, S. C., J. W. Head, and C. I. Fassett (2009) Unique chronostratigraphic marker in depositional fan stratigraphy on Mars: Evidence for ca. 1.25 Ma gully activity and surficial meltwater origin, Geology, 37, 207–210. Scott, D. H., J. M. Dohm, and J. R. Zimbelman (1998) Geologic map of Pavonis Mons volcano, Mars, US Geol. Surv. Misc. Invest. Ser. Map I-2561.   Page 32 of 45  732  733  734  735  736  737  738  739  740  741  742  743  744  745  746  747  748  749  750  751  752  753  754  Shean, D. E., J. W. Head, M. Kreslavsky, G. Neukum and HRSC Co-I Team (2006) When were glaciers present in Tharsis? Constraining age estimates for the Tharis Montes fan-shaped deposites, Lunar Planet. Sci. Conf., 37th, #2092. Smith, D. E., M. T. Zuber, S. C. Solomon, R. J. Phillips, J. W. Head, J. B. Garvin, W. B. Banerdt, D. O. Muhleman, G. H. Pettengill, G. A. Neumann, F. G. Lemoine, J. B. Abshire, O. Aharonson, C. D. Brown, S. A. Hauck, A. B. Ivanov, P. J. McGovern, H. J. Zwally, T. C. Duxbury (1999) The global tpography of Mars and implications for surface evolution, Science, 284, 1495–1503. Svetsov, V. V., I. V. Nemtchinov, and A. V. Teterev (1995) Disintegration of large meteoroids in Earth’s atmosphere: Theoretical models, Icarus, 116, 131–153. Vasavada, A. R., J. L. Bandfield, B. T. Greenhagen, P. O. Hayne, M. A. Siegler, J.-P. Williams, and D. A. Paige (2012) Lunar equatorial surface temperatures and regolith properties from the Diviner Lunar Radiometer Experiment, J. Geophys. Res., 117, E00H18, doi:10.1029/2011JE003987. Watkins, J. S. and R. L. Kovach (1973) Seismic investigation of the lunar regolith, Proc. Fourth Lunar Sci. Conf., 3, 2561–2574. Weibull, W. A. (1951) A statistical distribution function of wide applicability, J. Appl. Mech., 10, 140–147. Wilhelms, D. E. (1987) The geologic history of the Moon, U.S. Geol. Surv. Prof. Pap., 1348, 302 pp. Williams, J.-P., O. Aharonson, and A. V. Pathare (2010) The production of small primary craters on Mars, Lunar Planet. Sci. Conf., 41st, #2574.   Page 33 of 45  755  756  757  758  759  760  761  762  763  764  765  766  Tables Table 1 Distribution of material types based on fireball network observations (Ceplecha et al., 1998). Density, ρm (kg m-3) Ablation Coef.,  % obs. Group 7.010-8 Irons 3 7800 1.410-8 3700 29 Ordinary Chondrites 4.210-8 2000 33 Carbonaceous Chondrites 10.010-8 Cometary Material 26 750 21.010-8 Soft Cometary Material 270 9 Table 2 Sensitivity analysis Best-fit1 isochron from Hartmann (2005) Model (1 yr) D = 20 – 300 m D = 4 – 20 m Nominal 1.030  0.017 0.734  0.002 Highlands (ρo = 0.0135 kg m-3) 0.920  0.002 1.140  0.018 Lowlands (ρo = 0.0222 kg m-3) 0.687  0.002 0.993  0.017 1.170  0.018 0.946  0.002   ½ 0.511  0.002 0.843  0.016   2 Ordinary Chondrite only 2.13  0.025 1.75  0.003 Carbonaceous Chondrite only 0.46  0.002 0.872  0.017 Cometary Marterial only 0.15  0.007 0.047  0.001 Lunar Regolith 0.903  0.002 0.959  0.016 Soft Rock 0.561  0.013 0.327  0.001 Hard Rock 0.193  0.001 0.295  0.009 1 Isochrons fit to cumulative SFD using Craterstats2 program (Michael and Neukum, 2009) Figures   Page 34 of 45  767  768  769  770  771  772  773  774  775  776  777  778  779  Figure 1. (a) Initial velocity versus final velocity and (b) the ratio of initial and final projectile mass versus initial velocity for a range of projectile diameters (10 cm – 1 m) where objects have properties of ordinary chondrites (Table 1). Smaller, faster objects are more effectively decelerated and ablated. Figure 2. (a) General classes of projectiles based on initial mass and velocity. Large, slow projectiles: remain relatively unchanged as they are not energetic enough to ablate and too massive for the atmosphere to decelerate. Small, slow projectiles: the mass of the atmospheric column becomes comparable to the mass of the object and deceleration occurs. Fast projectiles: deceleration and ablation are significant as the energy of traversing a column of atmosphere at a   Page 35 of 45  780  781  782  783  784  785  786  787  788  789  790  791  792  793  794  given velocity exceeds that required to ablate the entirety of its mass. The triangular wedge defines intermediate objects that would significantly ablate if deceleration does not occur, however deceleration limits the ablation. Model results of the ratios of initial and final (b)velocities and (c) masses are shown for projectile masses non-dimensionalized by the mass of the atmospheric column encountered on the x-axis and the initial projectile velocity squared by the energy per unit mass, σ−1, on the y-axis. For σ = 1 × 10−8 kg J−1, slow projectiles have vi < 10 km s-1. Figure 3. Scatter plots of results from the Monte Carlo simulation with a bivariate distribution of velocities and projectile diameters for 2 × 105 events with initial entry angles 45° and ordinary chondritic compositions. (a) Impact velocity versus initial entry velocity, (b) initial and final mass ratio versus entry velocity, and (c) final crater diameter versus crater diameter without an atmosphere, i.e. crater diameter resulting from initial mass and velocity. The color scale of (a-c) shows the initial projectile diameter. (d-f) are the same as (a-c) but events are tagged with 5   Page 36 of 45  795  796  797  798  799  800  801  802  803  804  805  806  807  808  809  colors representing the general class of projectiles as defined in figure 2 (Slow/Fast, Small/Large). This trend is show with arrows in (f). Deviations from the grey dashed lines in (a),(c),(d), and (f), and from the horizontal in (b) and (e), shows the magnitude of ablation and/or deceleration experienced by the objects. Crater diameters are determined assuming parameters of dry soil with effective strength, Ȳ = 65 kPa and ρt = 2000 kg m-3 (Holsapple, 1993). Figure 4. Profiles of (a) the percent of initial velocity, (b) the percent of initial diameter, and (c) the dynamic pressure for six ordinary chondrites in the martian atmosphere as a function of altitude with entry angles 45°. (d) Assumed atmospheric density with altitude. Note that while the faster projectiles decelerate to a greater relative extent, they experience larger dynamic pressure upon entry. Ablation is more sensitive to velocity, however, deceleration is more sensitive to projectile size, and therefore smaller objects tend to experience lower dynamic pressures.   Page 37 of 45  810  811  812  813  814  Figure 5. Histogram of fragmentation altitudes using m = 0.65 MPa. Most fragmentation occurs ~2-3 scale heights above the surface with few fragmentation events occurring at lower altitudes.   Page 38 of 45  815  816  817  818  819  820  821  822  823  824  825  826  827  Figure 6. Histograms of crater diameters for (a) the 44 fresh craters reported by Daubar et al. (2013) and (b) the model using m = 0.65 MPa. Diameters of crater clusters are effective diameters (see text). The inset is the model craters for D > 5 m. Figure 7. (a) The modeled crater SFDs where no fragmentation occurs (i.e. m >> ρav2 for all meteoroids) and two different fragmentation characteristics: one favoring fewer fragments with few large fragments and many small fragments (nominal model), and one with a uniform random distribution of fragment numbers and sizes to increase the influence of fragmentation on the crater SFD. (b) The ratio of crater SFDs with and without fragmentation. The influence of fragmentation on the SFD is greatest at D  2 – 20 m.     Page 39 of 45  828  829  830  831  832  833  834  835  836  837  838  839  840  Figure 8. Model Deff /D for: (a) individual crater clusters (b) a 2D histogram and (c) the mean in log Deff bins. Error bars represent the standard deviation in each bin. Deff is increasingly unreliable at smaller sizes below ~10 m.   Figure 9. Crater counts conducted on a ~5 km2 area north of the Zunil crater rim using HiRISE image PSP_001764_1880 which was calibrated and map-projected using ISIS (Integrated Software for Imagers and Spectrometers) and imported into Arcmap, with CraterTools (Kneissl et al., 2011) used to identify and measure crater diameters. (a) Location of Zunil crater count area in HiRISE image PSP_001764_1880. (b) Log-differential plot of Deff for crater counts for Zunil and the 44 new impact sites constrained by CTX (Daubar et al., 2013) (blue) and model   Page 40 of 45  841  842  843  844  845  results (red) for the corresponding surface area and time. Cumulative crater frequency of (c) Deff and (d) for individual craters for crater counts (blue) and model results (red) using the Craterstats2 software tool (Michael and Neukum, 2009).   Page 41 of 45  846  847  848  849  Figure 10. The cumulative SFD for the fresh craters from CTX-CTX detections (Daubar et al., 2013), MOC (Malin et al., 2006), and model results, scaled to the same time/area, and the annual Hartmann (2005) isochron (gray line).   Page 42 of 45  850  851  852  853  854  855  856  Figure 11. (a) Location of North Ray crater count area (350 m  300 m blue box) in LRO NAC image M129187331 (NASA/GSFC/Arizona State Univ.). (b) The cumulative crater frequency of the crater counts (blue) and the model (red) using the Craterstats2 software tool (Michael and Neukum, 2009). The gray line indicates crater saturation.   Page 43 of 45  857  858  859  860  861  862  Figure 12. Sensitivity of cumulative SFD to model parameters. (a) Nominal model, fresh craters identified with MOC (Malin et al., 2006), and CTX (Daubar et al., 2013), and counts conducted on Zunil ejecta scaled to an annual isochron of Hartmann (2005). Estimated ages are for D = 4 – 20 m and D = 20 – 300 m. Nominal model assumes: ρo = 0.02 kg m-3 consistent with the elevation of Zunil crater (-2.8 km), the distribution of meteoroid types with corresponding   Page 44 of 45  863  864  865  866  867  868  ablation coefficients as given by Ceplecha et al. (1998), and target material properties consistent with dry desert alluvium, K1,= 0.24, μ = 0.41 and Ȳ = 65 kPa. (b) Results for atmospheric surface densities for the Highlands and Lowlands. (c) Results scaling the ablation coefficients, , by 0.5 and 2. (d) Results assuming all projectiles made of ordinary chondrites, carbonaceous chondrites, or cometary material as defined in Table 1. (e) Results assuming target material properties of lunar regolith, soft rock, and hard rock.   Page 45 of 45 
1603.05677
3
1603
2016-03-22T18:39:33
Discovery of a Gas giant Planet in Microlensing Event OGLE-2014-BLG-1760
[ "astro-ph.EP" ]
We present the analysis of the planetary microlensing event OGLE-2014-BLG-1760, which shows a strong light curve signal due to the presence of a Jupiter mass-ratio planet. One unusual feature of this event is that the source star is quite blue, with $V-I = 1.48\pm 0.08$. This is marginally consistent with source star in the Galactic bulge, but it could possibly indicate a young source star in the far side of the disk. Assuming a bulge source, we perform a Bayesian analysis assuming a standard Galactic model, and this indicates that the planetary system resides in or near the Galactic bulge at $D_L = 6.9 \pm 1.1 $ kpc. It also indicates a host star mass of $M_* = 0.51 \pm 0.44 M_\odot$, a planet mass of $m_p = 180 \pm 110 M_\oplus$, and a projected star-planet separation of $a_\perp = 1.7\pm 0.3\,$AU. The lens-source relative proper motion is $\mu_{\rm rel} = 6.5\pm 1.1$ mas/yr. The lens (and stellar host star) is predicted to be very faint, so it is most likely that it can detected only when the lens and source stars are partially resolved. Due to the relatively high relative proper motion, the lens and source will be resolved to about $\sim46\,$mas in 6-8 years after the peak magnification. So, by 2020 - 2022, we can hope to detect the lens star with deep, high resolution images.
astro-ph.EP
astro-ph
Discovery of a Gas Giant Planet in Microlensing Event OGLE-2014-BLG-1760 A. Bhattacharya1,M , D.P. Bennett1,2,M , I.A. Bond3,M , T. Sumi4,M , A. Udalski5,O, R. Street6,R, Y. Tsapras6,7,8,R and F. Abe9, M. Freeman10, A. Fukui11, Y. Itow9, M. C. A. Li12, C. H. Ling3, K. Masuda9, Y. Matsubara9,Y. Muraki9, K. Ohnishi13, L. C. Philpott14, N. Rattenbury12, T. Saito15, A. Sharan12, D. J. Sullivan16, D. Suzuki2, P. J. Tristram17 (MOA collaboration), J. Skowron5, M. K. Szyma´nski5, I. Soszy´nski5, R. Poleski5,18, P. Mr´oz5, S. Kozlowski5, P. Pietrukowicz5, K. Ulaczyk5, L. Wyrzykowski5 E. Bachelet6,19, D. M. Bramich19,20, G. D'Ago21,22, M. Dominik23,†, R. Figuera Jaimes20,23, (OGLE collaboration), K. Horne23, M. Hundertmark24,25, N. Kains26, J. Menzies27, R. Schmidt7, C. Snodgrass28,29,I. A. Steele30, J. Wambsganss7 (ROBONET collaboration) 6 1 0 2 r a M 2 2 . ] P E h p - o r t s a [ 3 v 7 7 6 5 0 . 3 0 6 1 : v i X r a -- 2 -- ABSTRACT We present the analysis of the planetary microlensing event OGLE-2014-BLG-1760, which shows a strong light curve signal due to the presence of a Jupiter mass-ratio planet. One unusual feature of this event is that the source star is quite blue, with V − I = 1.48 ± 0.08. This is marginally consistent with source star in the Galactic bulge, but it could possibly indicate a young source star in the far side of the disk. Assuming a bulge source, we perform a Bayesian analysis assuming a standard Galactic model, and this indicates that the planetary system resides in or near the Galactic bulge 1Department of Physics, University of Notre Dame, 225 Nieuwland Science Hall, Notre Dame, IN 46556, USA; Email: [email protected] 2Code 667, NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA 3Institute of Natural and Mathematical Sciences, Massey University, Auckland 0745, New Zealand 4Osaka University, 1-1 Yamadaoka, Suita, Osaka Prefecture 565-0871, Japan 5Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland 6Las Cumbres Observatory Global Telescope Network, 6740 Cortona Drive, Suite 102, Goleta,CA 93117, USA 7Astronomisches Rechen-Institut, Zentrum fur Astronomie der Universitat Heidelberg (ZAH), 69120 Heidelberg, Germany 8School of Physics and Astronomy, Queen Mary University of London, Mile End Road, London E1 4NS, UK 9Institute for Space-Earth Environmental Research, Nagoya University, 464-8601 Nagoya, Japan 10Dept of Physics, University of Auckland, Private Bag 92019, Auckland, New Zealand 11Okayama Astrophysical Observatory, National Astronomical Observatory of Japan, Asakuchi,719-0232 Okayama, Japan 12Dept. of Physics, University of Auckland , Private Bag 92019, Auckland, New Zealand 13Nagano National College of Technology, 381-8550 Nagano,Japan 14Department of Earth, Ocean and Atmospheric Sciences, UNiversity of British Columbia, Vancouver, British Columbia, Vancouver, British Columbia, V6T 1Z4, Canada 15Tokyo Metroplitan College of Industrial Technology, 116-8523 Tokyo, Japan 16School of Chemical and Physical Sciences, Victoria University, Wellington, New Zealand 17Mt. John University Observatory, P.O. Box 56, Lake Tekapo 8770, New Zealand 18Department of Astronomy, Ohio State University, 140 West 18th Avenue, Columbus, OH 43210, USA 19Qatar Environment and Energy Research Institute (QEERI), HBKU, Qatar Foundation, Doha, Qatar 20European Southern Observatory, Karl-Schwarzschild-Str. 2, 85748 Garching bei Munchen, Germany 21Dipartimento di Fisica "E.R. Caianiello", Universit`a di Salerno, Via Ponte Don Melillo, 84084-Fisciano (SA), Italy 22Istituto Nazionale di Fisica Nucleare, Sezione di Napoli, Napoli, Italy 23School of Physics & Astronomy, University of St Andrews, North Haugh, St Andrews KY 16 9SS, UK 24Niels Bohr Institute, University of Copenhagen, Juliane Maries Vej 30, 2100, Kobenhavn, Denmark 25SUPA, School of Physics & Astronomy, University of St Andrews, North Haugh, St Andrews KY16 9SS, UK 26Space Telescope Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 27South African Astronomical Observatory, PO Box 9, Observatory 7935, South Africa 28Planetary and Space Sciences, Dept of Physical Sciences, The Open University, Milton Keynes, MK7 6AA, UK 29Max Planck Institute for Solar System Research, Justus-von-Liebig-Weg 3, 37077 Gotingen, Germany 30Astrophysics Research Institute Liverpool John Moores University, Liverpool L3 5RF, UK and Royal Society University Research Fellow †Royal Society University Research Fellow MMicrolensing Observation in Astrophysics OOptical Gravitational Lensing Experiment RRobonet -- 3 -- at DL = 6.9±1.1 kpc. It also indicates a host star mass of M∗ = 0.51±0.44M(cid:12), a planet mass of mp = 180±110M⊕, and a projected star-planet separation of a⊥ = 1.7±0.3 AU. The lens-source relative proper motion is µrel = 6.5 ± 1.1 mas/yr. The lens (and stellar host star) is predicted to be very faint, so it is most likely that it can detected only when the lens and source stars are partially resolved. Due to the relatively high relative proper motion, the lens and source will be resolved to about ∼ 46 mas in 6-8 years after the peak magnification. So, by 2020 - 2022, we can hope to detect the lens star with deep, high resolution images. Subject headings: gravitational lensing: micro, planetary systems 1. Introduction Gravitational Microlensing is the technique of detecting exoplanets using gravitational lensing. Microlensing is unique in its ability to detect planets (Gould & Loeb 1992) just outside of the snow line (Lissauer 1993) down to an earth mass (Bennett & Rhie 1996), which is difficult or impossible with other methods. According to the core accretion theory, the snow line (Ida & Lin 2005; Lecar et. al. 2006; Kennedy et. al. 2006; Kenyon & Hartmann 1995) plays a very crucial role in the planet formation process. Beyond the snow line, ices condense, increasing the density of the solid materials. Higher density of solids speeds the planet formation process in the protoplanetary disk, hence forming cold planets quickly beyond the snow line. The planets, discovered by microlensing, provide the statistics needed to understand the architecture of cold planets beyond the snow line (Suzuki et al. 2016). Since this method does not rely on the light from the host stars, it can detect planets, even when the host stars cannot be detected (Mao & Paczy´nski 1991). Most of the planets discovered so far using microlensing are ∼ 1-8 kpc away. Thus, microlensing is able to detect planets in the inner Galactic disk and bulge, where it is difficult to detect planets with other methods. Thus, microlensing has the potential to measure how the properties of exoplanets depend on the Galactic environment. For most planetary microlensing events, the angular Einstein radius, θE is measured from the finite source effect. In events where the lens star brightness (Bennett et al. 2006) or the parallax effect is measured (Gould 1992; Gaudi et al. 2008; Muraki et al. 2011), the planetary mass and distance to the planetary system can be determined . Hence this technique can be used to build statistics of planetary mass as a function of the host star mass. Since the planets detected by microlensing are ∼ 1-8 kpc away, the distance to the planetary system will also allow the determination of the planetary mass function as a function of the distance towards the galactic center. In this paper we present the discovery of a gas giant planet orbiting the lens stars for mi- crolensing event OGLE-2014-BLG-1760. The mass ratio of this planet is q = 8.64 × 10−4, which is slightly less than that of Jupiter. The paper is organized as follows: Section 2 describes the light -- 4 -- curve data collected for the event OGLE-2014-BLG-1760. The next section (Section 3) is divided in four parts: 3.1 summarizes data reduction procedures for the different data sets; Section 3.2 shows the best fit model and procedures that are used to obtain it; Section 3.3 describes how the limb darkening of the source is modeled; and section 3.4 presents our attempt to detect the microlensing parallax effect in the light curve. In Section 4, we discuss source brightness and angular radius measurement and derive the lens-source relative proper motion. Section 5 discusses an estimate of the lens properties and the future possible investigations. 2. Observation The OGLE (Optical Gravitational Lensing Experiment) collaboration operates a microlensing survey towards the galactic bulge, with the 1.3 meter Warsaw telescope from Las Campanas obser- vatory in Chile. Most of the OGLE-IV (phase 4) observations were taken in the Cousins I-band, with occasional observations in Johnson V -band (Udalski et al. 2015). In the year 2014, OGLE Early Warning System (EWS) (Udalski et al. 1994) has alerted 2049 microlensing candidates of which this event OGLE-2014-BLG-1760 was the 1760th one. The MOA (Microlensing Observation in Astrophysics) (Bond et al. 2001) collaboration also operates a microlensing survey towards the galactic bulge with the 1.8 meter MOA-II telescope from Mount John Observatory at Lake Tekapo, New Zealand. The observations are mostly taken in MOA-red wide band filter which covers the wavelengths of the standard Cousins R + I bands. For year 2014, MOA has reported about ∼33 microlensing anomalies out of which OGLE-2014- BLG-1760 planetary event is one of them. Both of these telescopes have relatively large fields-of-view, 2.2 deg2 for MOA-II, and 1.4 deg2 for OGLE-IV. These enable survey observations with cadences as high as one observation every 15 minutes, and this allows the surveys to detect the sharp light curve features of planetary light curve features, when they are only smoothed by the finite source effects of a main sequence source star. It is the high cadence observation of microlensing events of MOA-II and OGLE-IV survey that helps in detecting microlensing anomalies, including the microlensing planetary signatures. The microlensing event OGLE-2014-BLG-1760 was discovered at (RA,decl.)(2000) = (17 : 57 : 38.16,−28 : 57 : 47.37)[(l,b)=(1.3186,−2.2746)], by the OGLE EWS on August 22, 2014 around 7.25 a.m. EDT. The same event was alerted by MOA on August 31, 2014 as MOA-2014-BLG-547. Later, on September 10, 2014, the MOA collaboration detected the planetary cusp crossing and announced the anomaly in the light curve (around HJD = 2456911 in Figure 1). In response to MOA anomaly alert, the follow up groups Robonet and µFUN started collecting data on this event. The high cadence observation of MOA-II survey covered most of the cusp crossing, and the trough after the cusp crossing was well covered by follow up groups, Robonet and µFUN. The Robonet group observed the event with 1 m robotic telescopes at Cape Town, South Africa and at Siding Springs, Australia, in the Sloan I-band. The µFUN group also observed the event with the 1.3 m -- 5 -- SMARTS CTIO telescopes in both the V and I-bands. All these observations are shown in Figure 1. Although the µFUN data is pretty flat because the trough is followed by the normal single lens decrease ( see Figure 1), the µFUN data do help to exclude the wide model with s = 1.27 and contribute in constraining the planetary model ( discussed in section 3.2). It is true that the constraint from the Robonet and OGLE data is stronger because of better Robonet coverage and OGLE observations over a wider range of magnifications. 3. Data Reduction and Modeling 3.1. Data Reduction All the images were reduced to photometry using the difference image method (Tomany & Crotts 1996; Alard & Lupton 1998; Alard 2000). The MOA images were reduced using MOA difference image analysis pipeline (Bond et al. (2001)). The data was then detrended using the non-2014 data to minimize the error due to the effects of differential refraction, seeing and airmass. To avoid systematic errors due to flat field changes and changes in detector sensitivity, we removed data prior to 2011 and only use the data after HJD(cid:48) = 5596 (February, 2011) for modeling. OGLE data were reduced using OGLE Difference Image Analysis software and optimal centroid method (Udalski (2003); Wozniak (2000)). The µFUN photometry was produced using a modified version of the PySIS package (Albrow et. al. 2009). Robonet data were reduced using the DanDIA package Bramich (2008). (cid:113) σ2 i + e2 The error bar estimates obtained from the photometry codes are sufficient to find the best fit model. These error bar estimates are good measures of the relative uncertainties for measurements with the same telescope, but they are often wrong by a factor of ∼2. To determine the uncertainties for the model parameters, it is necessary to have accurate error bars which will give χ2/dof ≈ 1 for each passband. We used method of Bennett et al. (2014) to normalize the error bars using σ(cid:48) i = k i is the modified error bar. A k value is selected for each passband to give χ2/dof = 1. Our initial fits used k = 1.50 and emin = 0.003 for all passbands. Next, once the best fit model was found, the values of k was modified for each passband to satisfy χ2/dof = 1. The number of datapoints used for each passband and their corresponding k values are listed in Table 1. New fits were done with the new error bars, which resulted in very small changes to the model parameters. min. Here σi is the initial uncertainty of i-th data point; σ(cid:48) 3.2. Best Fit Model We begin our modeling of the OGLE-2014-BLG-1760 light curve with single lens microlensing model (Paczy´nski 1986). There are three non-linear model parameters for a single lens event: t0 - the time of peak magnification, u0 - the minimum separation between source and lens in Einstein -- 6 -- Fig. 1. -- The light curve of event OGLE-2014-BLG-1760. The light grey dashed line represent the best fit single lens model and the dark grey line represents the best planetary model. The data are plotted in the following colors: red for the MOA-red band, black for OGLE-I, purple for OGLE-V , (difficult to see due to overlap with the OGLE-I data), blue for the ROBONET SAAO-A I-band, gold for the ROBONET SAAO-B I-band, green for the ROBONET Siding Springs-A I-band, cyan for the ROBONET Siding Springs-B I-band, and magenta for the µFUN I-band. -- 7 -- Fig. 2. -- This figure shows the caustic configuration and source trajectory for the best fit planetary model. When the source crosses inside the planetary caustic, its magnification jumps as two new highly magnified images are created, and then the magnification drops just as abruptly when the source exits and two images disappear. The two caustic crossings are at HJD(cid:48) − 2450000 = 6911.00 and 6911.07. Since the interval between the caustic crossings is similar to the source radius crossing time, t∗, the two magnification peaks merge into one peak, as shown in Figure 1 at HJD(cid:48) ∼ 6911. Also, when the source passes between the two minor image planetary caustics, the magnification of the source brightness drops, as can be seen in the light curve after caustic crossing, between HJD(cid:48) = 6911.20 and 6914.00 in Figure 1. -- 8 -- radius units, and tE - the Einstein radius crossing time. There are also two linear parameters for each passband: the source flux fs, and the blend flux, fbl. We find the best single lens parameters as the starting point for a systematic search through parameter space to find the best binary lens solution. Because the light curve follows a single lens shape for most of its history, except in the vicinity of the planetary feature at t ≈ 6911, we can use the best fit single lens parameters as the starting point for an initial condition grid search, following Bennett (2010). To describe a binary lens, we need three additional parameters: the lens mass ratio, q, the projected separation between the lens masses, s, measured in Einstein radius units, and the angle between the source trajectory and the lens axis, θ. Also, binary events often have caustic or cusp crossings, which resolve the angular size of the source, so we need an additional parameter, the source radius crossing time, t∗, to model finite source effects. With the single lens parameters fixed to the best fit values, we using the initial condition grid search method to search over the parameter ranges 0.48 ≤ s ≤ 2.10, −4 ≤ log q ≤ −2, and −π ≤ θ ≤ π, with t∗ fixed at t∗ = 0.05. We then select ∼ 10 of the best fit values from the initial condition grid search (with very different values of q, s and θ) to use as initial conditions for full, non-linear modeling runs using the Bennett (2010) χ2 minimization recipe, which is a modification of Markov Chain Monte Carlo algorithm (Verde et al. 2003). The parameters of the best fit binary lens model and the χ2 improvement are compared with the best fit single lens model in Table 2. The binary lens best fit model has q = 8.64 × 10−4 improves the renormalized χ2 by ∆χ2 = 1218.75 compared to the best fit single lens model. Hence the binary lens model is the preferred model. High magnification planetary microlensing events usually have a "close-wide" degeneracy, in Table 1. Data Reduction Summary Telescope Filter Na/Nb k MOA OGLE OGLE µFUN µFUN ROBONET SAAO A ROBONET SAAO B ROBONET Siding springs- A ROBONET Siding springs- B red(I+R) I V I V I I I I 11655/11673 10779/10794 119/119 45/45 3/4 54/54 32/32 56/56 75/75 0.63 0.98 0.84 0.79 0.99 1.16 0.96 1.24 1.23 a Number of data points used b Number of total observations -- 9 -- that solutions with s ↔ 1/s are nearly degenerate. This is usually not the case for low magnification planetary events because the caustic structure for the major and minor image caustics is quite different. However, the major and minor image caustics with s ↔ 1/s are encountered at the same single lens magnification, and with specific source trajectories, it is possible to produce similar light curves with both s > 1 and s < 1, particularly if the light curves aren't very well sampled. Therefore, we have searched for models with s ∼ 1/0.83 = 1.20. We found a best fit s > 1 model with s = 1.27, but this is a worse fit than the best fit model (with s = 0.83) by ∆χ2 ∼ 215. This is because the OGLE-2014-BLG-1760 is well sampled, so the s = 1.27 model is excluded, and the s = 0.83 is the only viable solution. 3.3. Limb Darkening Effect The photometric calibrations and the extinction toward the red clump stars in the vicinity of the source are discussed in Section 4. The magnitude and color of the source are indicated with the cyan point in Figure 4, and the extinction corrected color is (V − I)S,0 = 0.34, as discussed in section 4. This color implies that the source is an ∼A9 star with Teff ∼ 7352 K from Kenyon & Hartmann (1995). We use a linear limb darkening model, and from Claret (2000), we select the limb darkening parameters u to be 0.4204, 0.5790 and 0.46035 corresponding to the V , I and MOA-Red bands, respectively. These corresponds to the temperature of Teff = 7352 K and a surface gravity of log g = 4.5. The parameters of the best fit model are insensitive to the precise limit darkening parameters. 3.4. Search for a Microlensing Parallax Signal The microlensing parallax effect has been detected in a number of planetary microlensing events (Gaudi et al. 2008; Bennett et al. 2010; Muraki et al. 2011) where it has allowed the lens system masses to be measured. Due to the Earth's orbital motion about the Sun, the apparent lens - source relative motion deviates from uniform linear motion. This phenomenon is known as microlensing parallax effect (Gould 1992; Alcock et. al. 1995). The parallax effect can be described with the parallax vector πE = (πE,N , πE,E), where direction of πE is same as lens-source relative proper motion, µrel. The amplitude of πE is the inverse of the Einstein radius projected to the observer's plane, πE = AU/rE. When both the microlensing parallax effect and the angular Einstein radius (as described in section 4) are measured, we can determine the lens system mass and distance from the following equations: ML = DL = θE kπE AU πEθE + AU DS , (1) (2) -- 10 -- where k = 4G/(c2AU) = 8.14 mas M−1(cid:12) , ML is the total mass of host star and planet and DL and DS ∼ 8 kpc are the distances to the lens system and source stars, respectively. The host star mass, M∗, is given by M∗ = ML/(1 + q). In order to include the microlensing parallax effect, we must add two new model parameters πE and φE, the magnitude and direction angle of the parallax vector. (The north and east components of πE are given by πE,N = πE cos(φE) and πE,E = πE sin(φE).) Our best fit parallax model has an unusually large πE value of πE = 5.86, which would imply a very nearby and low mass lens if it is an accurate measurement, but the improvement in χ2 is only ∆χ2 = 9.90 over the best fit without parallax. We can examine the origins of this parallax signal by examining the cumulative ∆χ2 between the non-parallax and parallax models as a function of time, which we show in Figure 3. The cumulative ∆χ2 is displayed for the MOA and OGLE data separately, and we can see that only the OGLE data favors the parallax model. Penny et al. (2016) consider the distribution of published planetary microlensing events and show the there an implausibly large number with >∼ 1). This suggests that some of the published planetary "high" microlensing parallax values (πE events have spurious large πE values (Han et al. 2016). We conclude that this microlensing parallax measurement for OGLE-2014-BLG-1760 is also spurious. 4. Calibration and Source Properties The OGLE data were taken in the OGLE-IV I and V -bands, which we calibrate to the OGLE- III catalog Cousins I and Johnson V band (Szyma´nski et al. 2010). 151 bright (I ≤ 16.50) and isolated stars were matched in both passbands and used for this calibration. The following calibration relations are used to convert OGLE V I magnitudes to OGLE III Catalog Cousins I and Johnson V magnitude: Table 2. Model Parameters parameter units binary lens best fit single lens best fit tE t0 u0 s θ q t(cid:63) fit χ2 days HJD−2450000 radian 10−4 days 15.87 ± 0.41 6905.856 ±0.026 0.1806 ± 0.0074 0.8269 ± 0.0047 -0.3977 ±0.0086 8.64 ± 0.89 0.0366 ±0.0044 15.47 6905.9541 0.19 22818.05 24036.80 -- 11 -- Fig. 3. -- The difference in the cumulative χ2 values between the best fit models with and without parallax are shown for MOA data (blue), OGLE data (red) and total MOA + OGLE data (black). The negative cumulative ∆χ2 means parallax model is supported. Also MOA data does not support parallax whereas OGLE data supports parallax model. The χ2 improvement from MOA + OGLE data for parallax model is ∼ 3.60. The overall χ2 improvement for parallax model from all the telescopes is ∼ 9.90. Hence the parallax signal is dubious and probably a false signal. -- 12 -- IOGLEIV = −0.0471 + 0.99867IOGLEIII + 0.00133VOGLEIII VOGLEIV = −0.3444 − 0.10068IOGLEIII + 1.10068VOGLEIII (4) The uncertainty in the brightness due to these calibration relations is ∼ 0.01 magnitude. The source brightness in the OGLE III catalog scale is IS = 19.07± 0.14 , VS = 20.51± 0.26. The errors in brightness are calculated from MCMC averages over all the MCMC fits and from the uncertainty in the calibration relations. (3) There is a single magnified OGLE V -band observation at t = 6896.50 (HJD(cid:48)), and this results in a relatively large, 10% uncertainty in the V -band source flux for the best fit model. As a result, the OGLE V I source colors given a source star color of V − I = 1.45 ± 0.11. The color of the source star is bluer than the average main sequence star color as can be seen in figure 4. Because we have only this single V -band measurement, we also calculate the source star color from MOA- Red and OGLE-I band following Gould et al. (2010a) and Bennett et al. (2012). Since we have a large number of MOA-Red and OGLE-I observations when the source is significantly magnified, we expect a robust measurement of the RMOA − I source color. About 140 bright, isolated stars with I-band magnitude I < 16 and 1.0 < V − I < 2.6 are matched between the MOA-II and OGLE III reference images, and we used these to find the following calibration relation: RMOA − IOGLEIII = 0.18143 × (V − I)OGLEIII , where the RMOA passband has been calibrated to give RMOA = IOGLEIII when (V − I)OGLEIII = 0. The RMS error for this calibration relation is 0.0343, for a formal uncertainty of 0.0343/ N = 0.0029, but we assume a calibration uncertainty of 0.02 for RMOA − IOGLEIII. √ Our models give a source brightnesses of IS = 19.07 ± 0.14 and RMOA,S = 19.34 ± 0.15 and a calibrated source color of V − I = 1.52 ± 0.11. This color is less than 1-σ away from the color of the source derived from OGLE V -band data, so we conclude that the blue color from the OGLE data is probably real. However, the average color from both measurements is V − I = 1.48 ± 0.08, which still has a relatively large error bar. We take this average as the correct color of the source. This color is plotted in blue point in the color magnitude diagram of all stars within 140(cid:48)(cid:48) of the source star, Figure 4. The green points in the CMD figure 4 show the HST CMD plot (Holtzman et al. 1998) shifted to have the same red clump centroid position as the OGLE-III stars . The source color is slightly bluer than the main sequence star distribution shown with green points in the HST CMD in figure 4. The average color of main sequence bulge stars in the same magnitude range of the source (I = 19.07± 0.14) in in the HST CMD is V − I = 1.75. We can estimate the probability that a star with the measured color of the source (V − I = 1.48 ± 0.08) is drawn from this HST distribution by integrating the Gaussian describing the measured source color with its error bar over the distribution of stars in the same magnitude range from the HST CMD. We then divide this result by the result of the integral with the same error bar, but centered on the average color of V − I = 1.75. The ratio of integral centered on the measured color and the integral centered on -- 13 -- Fig. 4. -- The (V − I, I) color magnitude diagram (CMD) of the stars in the OGLE-III catalog Szyma´nski et al. (2011) within 140(cid:48)(cid:48) of OGLE-2014-BLG-1760. The red spot indicates red clump centroid, and the cyan spot indicates the source magnitude and color with error bars. The green points show HST CMD of Baade's window transformed to have the same red clump centroid as this field. The source star is slightly bluer than the HST main sequence stars. This might be explained by a young, blue source in the Milky Way disk on the far side of the bulge. -- 14 -- the average color is 0.05, indicating that this source color is marginally consistent with being drawn from the known bulge star population. With nearly, 50 planetary microlensing events published to date, we would expect one or two to be found with such a blue source star. On the other hand, it is also possible that blue stars are preferentially lensed because they are at a greater distance. This can be the case if the source is a young blue (late A or early F) star on the far side of the Galactic disk, assuming negligible extinction beyond the bulge on this line-of-sight. Such stars would have a much larger microlensing event rate, so they could be preferentially lensed by bulge stars. On the other hand, it is possible that the CMD in this field is noticeably different than that of Baade's window where the Holtzman et al. (1998) image was taken. We calculate the extinction in the direction of the source using the method described by Bennett et al. (2014). The position of the centroid of red clump in the OGLE III catalog is found to be (V − I, I)RC = (2.20, 15.84) as shown in figure 4. From Bensby et al. (2011) and Nataf et al. (2013), the dereddened red clump centroid is determined to be (V − I, I)RC = (1.06, 14.39). Hence, the extinction at this galactic coordinate is : (E(V − I), AI )RC = (1.14, 1.45). The extinction to the source star is assumed to be same as the average extinction for the red clump stars within 140(cid:48)(cid:48) since most of the extinction is thought to be in the foreground. Thus, the extinction corrected source color and brightness are (V − I, I)S,0 = (0.34, 17.62), as shown in blue in figure 4. The source radius is calculated from the dereddened source magnitude and color using the following formula, obtained from a private communication with Tabetha Boyajian: log10(2θ∗) = 0.50141358 + 0.41968496(V − I) − 0.2I (5) For the source color or V − I = 1.48, the calculated source radius is θ∗ = (6.57 ± 0.11) × 10−4 mas. The relative proper motion and Einstein radius are calculated from: µrel = θ∗ t(cid:63) θE = θ∗ × tE t(cid:63) Since the best fit model has tE = 15.87 days and t∗ = 0.04 days, we find: µrel = 6.55 ± 1.12 mas/yr θE = 0.29 ± 0.05 mas . (6) (7) (8) (9) There is a 5% uncertainty in source radius relation (equation 5). We assume 1-2% error in the calibration relations (equations 3 and 4). 5. Lens Properties and Discussion If we assume a standard Galactic model (Bennett et al. 2014) with a source star in the bulge and we assume that all stars and brown dwarf have an equal probability of hosting a planet with -- 15 -- the measured properties, then we can perform a Bayesian analysis to estimate the lens system properties. We ran an MCMC run with about 180,000 links to obtain the posterior distributions presented in Table 3 and Figure 5. The mass of the host star is only approximately determined to be M∗ = 0.51 +0.44−0.28 M(cid:12), so it could be an M, K, or G star. The 1-sigma range of the planet mass, mp = 182 +137−83 spans the range from the mass of Saturn to that of Jupiter. The distance to the lens system is more precisely determined, due to the relatively small angular Einstein radius, θE = 0.29 ± 0.05 mas. Our analysis predicts a lens system distance of DL = 6.86 ± 1.11 . This implies that the lens system is very likely to be in the Galactic bulge. Penny et al. (2016) argue that the published planetary microlensing events show a dearth of planets orbiting Galactic bulge stars. This OGLE-2014-BLG-1760L lens system would seem to be a counterexample, along with a number of other planetary microlens systems, such as OGLE-2005- BLG-380L (Beaulieu et al. 2006), OGLE-2008-BLG-092L (Poleski et al. 2014), MOA-2008-BLG- 310L (Janczak et al. 2010), OGLE-2008-BLG-355L (Koshimoto et al. 2014), MOA-2011-BLG-353L (Rattenbury et al. 2015) and MOA-2011-BLG-293L (Yee et al. 2012; Batista et al. 2014). Actually, the problem with the Penny et al. (2016) analysis is pretty easy to understand. The Einstein radius crossing times (tE) for bulge events are significantly smaller than the tE values for disk events, but the detection efficiencies for microlensing events and planetary signals are substantially higher for events with large tE values. The statistical analysis of Suzuki et al. (2016) shows that the planet detection efficiency weighted median tE value is ≈ 42 days. This effect occurs for high magnification events (Gould et al. 2010b) because high magnification is much easier to predict when tE is large, and for low magnification events (Suzuki et al. 2016) because the planetary signal duration (for fixed q) is proportional to tE. This effect can be accounted for with accurate detection efficiency calculations, but Penny et al. (2016) use a detection efficiency calculation for an advance ground- based survey, that is much more sensitive than any ground-based survey undertaken to date. One of us (Bennett 2004) has previously studied a variety of ground-based microlensing surveys with a wide range of telescope options, and these simulations show that this detection efficiency bias Table 3. Lens System Parameters parameter Host star mass, M∗ Planet mass, mP Host star - Planet 2D separation, a⊥ Host star - Planet 3D separation, a3D Lens distance, DL Lens magnitude, ILens Lens magnitude, HLens units M(cid:12) M⊕ AU AU kpc Cousins I H (V − I)S = 1.48 ± 0.08 0.51+0.44−0.28 182+137−83 1.75+0.34−0.33 2.57+1.12−0.45 6.86 ± 1.11 23.42+1.89−2.92 20.80+1.54−2.31 -- 16 -- with tE is much stronger with the less capable surveys that can approximate the sensitivity of the observing programs that have discovered the published events. It should also be noted that the color of the source star V − I = 1.48 ± 0.08 implies that this star is bluer than and only marginally consistent with bulge main sequence shown in Figure 4. This source star could be a "blue straggler" which is an old, bright blue star that probably formed from a merger of two smaller stars, but these stars are rare, or it could be that this field has a larger fraction of bluer, metal poor stars than the HST Baade's window stars plotted in Figure 4. But this also seems unlikely. Another possibility is that this star is a relatively young blue star that resides in the Galactic disk on the far side of the bulge with minimal additional extinction between the bulge and this source star. If so, HST observations taken separated by 2 years could reveal the characteristic source star proper motion of a far side disk star. Since the galactic coordinates of the source, (l, b) = 1.3186,−2.2746 are pretty close to the nominal WFIRST exoplanet microlensing survey fields (Spergel et al. 2015) field, so understanding the source distance distribution in this area of the bulge is important for WFIRST. The probability distribution for the brightness of the lens star in I and H-bands is shown in Figure 6. The blend magnitude from best fit is found Ibl = 17.94 which is brighter than the lens star magnitude predicted in Table 3 and Figure 6. Since the lens-source relative proper motion is µrel = 6.55 ± 1.12 mas/yr, the source and lens will be separated by ∼ 1 HST pixel (39.6 ± 6.2) by 2020. If the lens system is observed directly in high resolution images, then the brightness measurement of lens leads to the precise calculation of the mass and distance to the lens system (Bennett et al. 2006, 2015; Batista et al. 2015). But in the I-band, the source star is brighter than the median prediction for the lens by ∼ 4 magnitudes (see Figure 6), which makes it difficult to observe the lens in I band with the high resolution follow-up images unless the lens and source are completely resolved. The extinction corrected (I − H) color of the source is obtained from (V − I) color using Kenyon & Hartmann (1995). The H-band extinction AH is also obtained from AI and AV using the Cardelli et al. (1989) formula. From the I-band magnitude, the (I − H) color and the H-band extinction, AH , the H-band magnitude of the source is calculated to be HS = 17.96. From Figure 6, we see that the H-band magnitudes of source and median prediction for the lens differ by ∼ 2.0-2.5 magnitude. The uncertainty in the source brightness is measured from the Markov chain links. We also include the uncertainty in the source color and magnitude. There is 5% uncertainty in the color conversion from the extinction corrected (V − I) to (I − H). The red lines in Figure 6 is such that if lens is at least as bright as the red lines (i.e. it lies in the red-shaded region) then lens+source brightness is brighter than the 3-σ upper limit on the source brightness. This would mean that the lens could be detected when it is still unresolved from the source. Since most of the lens flux histogram lies to the right of the red-shaded region, the lens is likely to be too faint to be detected in high resolution images unless lens and source are partially resolved. Since the relative lens-source proper motion, µrel ∼ 6.5 mas/yr, the lens and source are expected to be resolvable by JWST, HST or adaptive optics imaging some in 6-8 years after peak magnification, in 2020 - 2022. Suzuki et al. (2016) derived a mass ratio function from planets detected in MOA-II survey -- 17 -- Fig. 5. -- The probability distribution of planet mass, host star mass, planet - host star separation and distance to lens system, derived using Bayesian statistical analysis. The central 68.3% of the distribution is shaded dark grey and the remaining central 95.4% of the distribution is shaded light grey. The vertical black line marks the median of the probability distribution of respective parameters. -- 18 -- Fig. 6. -- The probability distribution for the I and H-band lens star magnitudes given the source radius estimate based on the source magnitude and color given in Figure 4. As in Figure 5, central 68.3% is shaded dark grey and the remaining central 95.4% of the distribution is shaded light grey. The blue lines indicate the source magnitude, and the left and right panels show the I and H magnitude distributions, respectively. The vertical red lines mark the lens brightness corresponding to the 3-σ upper limit on the source brightness, so if the lens is in the light red shaded region it will be bright enough to push the combined lens+source brightness above the brightness from the light curve models. In this case, the excess brightness from the lens will be detectable even when the lens and source remain unresolved. If the lens is fainter and located to the right of the red shaded region, we will not be able to identify the lens with high angular resolution JWST, HST or adaptive optics observations until the lens star begins to separate from the source 5-8 years after peak magnification. -- 19 -- data, and discovered a break in the mass ratio function at q ∼ 10−4. The OGLE-2014-BLG-1760Lb planetary mass ratio of q = 8.6 ± 0.9 is above the mass ratio break qbr ∼ 10−4. This is close to Jupiter's mass ratio, but the host star probably has a mass of M∗ <∼ M(cid:12), so this planet is probably a low-mass gas giant, like Saturn. Follow up observations with JWST, HST or adaptive optics in 2020-2022 should be able to measure the lens brightness and determine the planetary mass and distance using the methods of Bennett et al. (2006, 2007, 2015) and Batista et al. (2015). Later with WFIRST (Spergel et al. 2015), a similar kind of study will provide the statistics needed to determine the planetary mass function as a function of the host star mass and distance. A.B., D.P.B. and D.S. were supported by NASA through grants NASA-NNX12AF54G and NNX13AF64G. I.A.B. and P.Y. were supported by the Marsden Fund of Royal Society of New Zealand, contract no. MAU1104. N.J.R. was supported through Royal Society of New Zealand Rutherford Discovery Fellowship. A.S, M.L. and M.D. were supported through Royal Society of New Zealand. T.S. received support from JSPS23103002, JSPS24253004 and JSPS26247023. MOA project received grants from JSPS25103508 and 23340064. A.F. was supported by the Astrobiology Project of the Center for Novel Science Initiatives (CNSI), National Institutes of Natural Sciences (NINS) (Grant Number AB261005). The OGLE project received funding from National Science Centre, Poland, grant MAESTRO 2014/14/A/ST9/00121 to AU. OGLE team thanks Profs. M. Kubiak and G. Pietrzy´nski, former members of the OGLE team, for their contribution to the col- lection of the OGLE photometric data over the past years. DMB was supported by NPRP grant X-019-1-006 from the Qatar National Research Fund (a member of Qatar Foundation). This work makes use of observations from the LCOGT network, which includes three SUPAscopes owned by the University of St Andrews. The RoboNet programme is an LCOGT Key Project using time allo- cations from the University of St Andrews, LCOGT and the University of Heidelberg together with time on the Liverpool Telescope through the Science and Technology Facilities Council (STFC), UK. This research has made use of the LCOGT Archive, which is operated by the California In- stitute of Technology, under contract with the Las Cumbres Observatory. We thank Prof. Andrew Gould and µFUN team for allowing us to use their data and acknowledge their hard work and contribution in collecting the µFUN data. REFERENCES Alard C., 2000, A&AS, 144, 363 Alard C., Lupton R. H., 1998, ApJ, 503, 325 Albrow, M.D., et al., 2009, MNRAS, 397, 2099 Alcock, C., et. al., 1995, ApJ, 454, 125 -- 20 -- Batista, V., Beaulieu, J.-P., Bennett, D.P.,et al., 2015, ApJ, 808, 170 Batista, V., Beaulieu, J.-P., Gould, A., et al. 2014, ApJ, 780, 54 Beaulieu, J.-P., Bennett, D. P., Fouqu´e, P., et al. 2006, Nature, 439, 437 Bennett, D. P. 2004, Extrasolar Planets: Today and Tomorrow, 321, 59 (arXiv:astro-ph/0404075) Bennett, D. P., 2010, ApJ, 716, 1408 Bennett, D. P. & Rhie, S. H., 1996, ApJ, 472, 660 Bennett, D. P., Anderson, J., et al., 2006, ApJ, 647, L171 Bennett, D. P., Anderson, J., 2007, Gaudi, S. B.,ApJ, 660, 781 Bennett, D. P., Rhie, S., et al., 2010, ApJ, 713, 837 Bennett, D. P., Sumi, T., Bond, I. A., et al., 2012, ApJ, 757, 119B Bennett, D. P., Batista, V., Bond, I. A., et al., 2014, ApJ, 785, 155 Bennett, D. P., Bhattacharya, A., Anderson J., et al., 2015, ApJ, 808, 169B Bensby, T., Ad´en, D., Melendez, J.,et al.., 2011, A&A, 533, A134 Bessell, M. S, & Brett, J. M., 1988, PASP, 100, 1134 Bond, I. A., Abe, F., Dodd, R. J., et al., 2001, MNRAS, 327, 868 Bramich, D. M., 2008, MNRAS, 386, L77b Cardelli, J. A., Clayton G. C., & Mathis J. A., 1989, ApJ, 345, 245C Claret, A., 2000, A&A, 363, 1081C Gaudi, S., Bennett, D. P., Udalski, A., et al., 2008, Science, 319, 927 Gould, A., 1992, ApJ, 392, 442 Gould, A., Loeb, A., 1992, ApJ, 396, 104G Gould, A., Dong, S., Bennett, D. P., et al. 2010a, ApJ, 710, 1800 Gould, A., Dong, S., Gaudi, B.S., et al. 2010b, ApJ, 720, 1073 Han, C., Bennett, D.P., Udalski, A., et al., 2016, ApJ, submitted Holtzman, J. A., Watson, A. M., Baum, W. A., et al., 1998, ApJ, 115, 1946H Ida, S., & Lin, D. N. C., 2005, ApJ, 626, 1045 -- 21 -- Janczak, J., et al., 2010, ApJ, 711, 731 Kennedy, G. M., Kenyon, S. J., & Bromley, B. C., 2006, ApJL, 650, L139 Kenyon, S. J., & Hartmann, L., 1995, ApJSS, 101, 117 Koshimoto, N., Udalski, A., Sumi, T., et al. 2014, ApJ, 788, 128 Lecar, M., Podolak, M., Sasselov, D., & Chiang, E, 2006, ApJ, 640, 1115 Lissauer, J.J. 1993, Ann. Rev. Astron. Ast., 31, 129 Mao, S., Paczy´nski, B., 1991, ApJ, 374L, 37M Muraki, Y., et al., 2011, ApJ, 741, 22 Nataf, D. M., Gould, A., Fouqu´e, P., et al., 2013, ApJ, 769, 88 Penny, M., Henderson, C. B., Clanton, C., 2016, arXiv:1601.02807 Paczy´nski, B., et al, 1986, ApJ, 304, 1P Poleski, R., Skowron, J., Udalski, A., et al. 2014, ApJ, 795, 42 Rattenbury, N. J., Bennett, D. P., Sumi, T., et al. 2015, MNRAS, 454, 946 Spergel, D., Gehrels, N., Baltay, C., et al. 2015, arXiv:1503.03757 Smith, M. C., Mao, S. & Wo´zniak, P. R., 2002, MNRAS, 332, 962 Sumi, T., Kamiya, K., Bennett, D. P., et al., 2011, Nature, 473, 349 Suzuki, D., Bennett, D. P.,Sumi, T., et al., 2016, ApJ, submitted. Szyma´nski, M. K., Udalski, A., Soszy´nski, I., et al., 2011, Acta Astron., 60, 295 Szyma´nski, M. K., Udalski, A., Soszy´nski, I., et al., 2011, Acta Astron., 61, 83 Tomany, A. B. & Crotts, A. P., 1996, AJ, 112, 2872 Udalski, A., 2003, Acta Astron., 53, 291 Udalski, A., Szyma´nski, M., Ka(cid:32)luzny, J., Kubiak, M., Mateo, M., Krzemi´nski, W., & Paczy´nski , B. 1994, Acta Astron., 44, 227 Udalski, A., Szyma´nski, M. K., & Szyma´nski, G., 2015, Acta Astron., 65, 1 Verde, L., Reiris, H. V., & Spergel, D. N., 2003, ApJS, 148,195 Wozniak P. R., 2000, Acta Astron., 50, 421 Yee, J. et al., 2012, ApJ, 755, 102 -- 22 -- This preprint was prepared with the AAS LATEX macros v5.2.
1803.06187
1
1803
2018-03-16T12:18:04
Pre-discovery transits of the exoplanets WASP-18 b and WASP-33 b from Hipparcos
[ "astro-ph.EP", "astro-ph.SR" ]
We recover transits of WASP-18 b and WASP-33 b from Hipparcos (1989-1993) photometry. Marginal detections of HAT-P-56 b and HAT-P-2 b may be also present in the data. New ephemerides are fitted to WASP-18 b and WASP-33 b. A tentative (~1.3 sigma) orbital decay is measured for WASP-18 b, but the implied tidal quality factor (Q' ~ 5 x 10^5) is small and survival time (<10^6 years) is too short to be likely. No orbital decay is measured for WASP-33 b, and a limit of Q' > 2 x 10^5 is placed. For both planets, the uncertainties in published ephemerides appear underestimated: the uncertainty in the period derivative of WASP-18 b would be greatly reduced if its current ephemeris could be better determined.
astro-ph.EP
astro-ph
MNRAS 000, 1–4 (2018) Preprint 19 March 2018 Compiled using MNRAS LATEX style file v3.0 Pre-discovery transits of the exoplanets WASP-18 b and WASP-33 b from Hipparcos I. McDonald1⋆, E. Kerins1 1Jodrell Bank Centre for Astrophysics, Alan Turing Building, Manchester, M13 9PL, UK Accepted XXX. Received YYY; in original form ZZZ ABSTRACT We recover transits of WASP-18 b and WASP-33 b from Hipparcos (1989–1993) pho- tometry. Marginal detections of HAT-P-56 b and HAT-P-2 b may be also present in the data. New ephemerides are fitted to WASP-18 b and WASP-33 b. A tentative (∼1.3σ) orbital decay is measured for WASP-18 b, but the implied tidal quality fac- ∼ 5 × 105) is small and survival time (< 106 years) is too short to be likely. tor (Q′ No orbital decay is measured for WASP-33 b, and a limit of Q′ > 2 × 105 is placed. For both planets, the uncertainties in published ephemerides appear underestimated: the uncertainty in the period derivative of WASP-18 b would be greatly reduced if its current ephemeris could be better determined. Key words: planets and satellites: dynamical evolution and stability - planets and satellites: gaseous planets - planets and satellites: individual: WASP-18 b - planets and satellites: individual: WASP-33 b - planet–star interactions - stars: variables: δ Scuti 1 INTRODUCTION Exoplanetary science is a relatively young field, hence many long-term evolutionary characteristics of planetary systems remain unknown. Pre-discovery archival data can provide, e.g., more precise orbital properties. Changes in these prop- erties may come from transit timing variations (TTVs) caused by a second planet in the system (e.g. Steffen et al. 2013), or by long-term orbital expansion or decay, due to stellar mass loss or tidal inspiral (e.g Mustill & Villaver 2012). In particular, historical data lets us constrain the tidal quality factor of exoplanet hosts, allowing us to model tidal effects from stars more generally (e.g. Penev et al. 2012). Few historical observations have sufficient sensitivity or cadence to detect exoplanets. Photometric accuracy of bet- ter than ∼0.01 mag is generally required, while duty cycles of transits are typically only a few per cent of the orbit, so dozens of repeated visits are necessary to secure a transit. Of the literature data available, only the Hipparcos satel- lite (Perryman & ESA 1997; van Leeuwen 2007) has suffi- cient accuracy and cadence to reliably search for exoplan- ets en masse. Hipparcos operated between 1989 and 1993, and returned broadband photometry to an accuracy of a few millimagnitudes on around 120 000 nearby stars. Tran- sits of HD 209458 b and HD 189733 b have had their Hip- parcos photometry published already (Robichon & Arenou ⋆ E-mail: [email protected] c(cid:13) 2018 The Authors 2000; H´ebrard & Lecavelier Des Etangs 2006). In this arti- cle, we search for transits of other known exoplanets in the original Hipparcos data1. 2 TRANSITING EXOPLANETS IN THE HIPPARCOS DATASET Exoplanets in the Hipparcos dataset were selected from the Exoplanets Data Explorer (EDE2; Han et al. 2014), us- ing the parameters "TRANSIT == 1 && HIPP > 0". This re- turned 17 unique systems. We further restricted our criteria to a transit depth >5 mmag (DEPTH > 0.005), returning the 11 systems listed in Table 1. For HAT-P-56 and HD 189733, outliers in the Hippar- cos data were removed using a κσ-clipping routine: i.e., an iterative pass of the data was performed, removing points more than κ standard deviations from the mean. A cutoff of κ = 3.5 was applied, which was chosen so as not to re- move points in the expected transit regions. As stars have between 54 and 187 data points, any choice of κ & 2.7 is not expected to remove valid data from the fit. The photometric data were folded on literature or- bit ephemerides (Table 1). Four transiting planets were 1 VizieR catalogue I/311 2 http://exoplanets.org 2 I. McDonald et al. Table 1. Hipparcos stars exhibiting transits of >5 mmag. Name HIP d WASP-18 b WASP-33 b HD 17156 b KELT-7 b KELT-2 A b HAT-P-56 b HD 80606 b GJ 436 b HAT-P-2 b HD 189733 b HD 209458 b 7562 11397 13192 24323 29301 32209 45982 57087 80076 98505 108859 (pc) 126 ± 5 118 ± 3 79.8 ± 1.6 138 ± 5 134 ± 6 319 ± 23 65.2 ± 1.1 10.1 ± 0.2 129 ± 4 19.8 ± 0.1 48.9 ± 0.5 T0 (TJD) (d) 4644.90531 4163.22373 4756.7313 6223.9592 5974.60335 6553.61645 4876.344 4415.62074 4397.49375 4279.43671 2826.62851 P (d) 0.94145299 1.21986975 21.21663 2.7347749 4.113791 2.7908327 111.43670 2.643850 5.6334729 2.21857567 3.52474859 T14 (d) 0.09089 0.11224 0.1338 0.14630 0.2155 0.09463 0.504 0.03170 0.1787 0.0760 0.1277 Depth (mmag) 9.16 11.36 5.29 8.28 5.21 11.11 11.17 6.96 5.22 24.12 14.61 Hip. rms N Expected detection Observed depth (mmag) 18.8 12.9 19.0 16.7 21.0 15.1 16.7 75.0 19.7 15.1 14.8 (σ) 1.38 3.29 0.00 0.00 0.25 0.74 0.00 0.00 0.75 2.76 1.97 9 15 0 0 2 6 0 0 9 4 5 (σ) 2.81 3.12 · · · · · · –0.79 0.83 · · · · · · 0.94 3.73 3.56 (mmag) 14.6 10.8 · · · · · · –16.6 12.5 · · · · · · 6.6 32.6 26.4 Notes: Distances come from Gaia Data Release 1 (Gaia Collaboration et al. 2016), with the exception of GJ 436, which comes from van Leeuwen (2007). Transit parameters are sourced from the EDE (values for WASP-18 b and WASP-33 b explicitly come from Wilkins et al. (2017) and Zhang et al. (2017)); truncated Julian dates are given as TJD = JD − 2 450 000 days. N is the number of observations expected during transit. expected to be detected (>1σ): WASP-18 b, WASP-33 b, HD 189733 b and HD 209458 b and all four were recovered. Transits of KELT-2 A b were not recovered due to the low signal-to-noise ratio. Transits of HAT-P-56 b and HAT-P-2 b were expected just below the 1σ detection limit, and mea- surements of the recovered transit depth are close to the 1σ limit. Since this measurement effectively uses a boxcar tran- sit, and since the Hipparcos photometric transmission curve is relatively blue (λeff ≈ 5275 A), a limb-darkened model is expected to recover these transits at just above 1σ. How- ever, since the photometry would be of insufficient quality to model further, they are neglected for the remainder of this paper. WASP-18 b and WASP-33 b have never previously been recovered from Hipparcos data. Their lightcurves are shown in Figure 1, folded on the empherides from Table 1. Data sampling is sparse: 132 points over 1190 days for WASP-18 b and 113 points over 930 days for WASP-33 b (one point has been cleaned by κσ-clipping from the latter). Consequently, a blind search for planets in the Hipparcos data would have been liable to miss these transits, which are not apparent in the unfolded lightcuves. 3 ORBITAL SOLUTIONS AND EVOLUTION previous analyses of (Robichon & Arenou in photometry As Hippar- 2000; cos H´ebrard & Lecavelier Des Etangs 2006), we note that fitting a two-parameter ephemeris (mid-transit epoch and period, T0 and P ) to data of this quality is less accurate than taking an established ephemeris and providing a refined period. In each case, T0, t14 and Rp/R∗ were held fixed to the values in Table 1, and the Hipparcos data were folded on a range of periods spanning 0.000015 days either side of these ephemerides. The transit was represented by a trapezoid ingress and egress, based on the above parameters. The impact of in- cluding limb darkening on the precision of the resulting fit was found to be significant, but the exact treatment of limb darkening was not. Hence, the transit between second and third contact was modelled as a point source crossing a limb- WASP-18 b / HIP 7562 WASP-33 b / HIP 11397 9.3 9.32 9.34 9.36 9.38 9.4 9.42 9.44 9.46 9.48 8.16 8.18 8.2 8.22 8.24 8.26 8.28 e d u t i n g a m s o c r a p p H i e d u t i n g a m s o c r a p p H i 8.3 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 Orbital phase Figure 1. Hipparcos photometry, phase-folded on a modern ephemeris. Lines show the expected transit position, width and depth. darkened star, with limb-darkening co-efficients taken from jktld (Southworth 2008): inputs of Teff = 6400 and 7430 K, log(g) = 4.367 and 4.300 dex and [Fe/H] = 0.0 and 0.1 dex were assumed for WASP-18 and WASP-33, respectively, while a microturbulent velocity of 2 km s−1 and a quadratic law with Claret (2004) models was assumed for both, and the Hipparcos filter was approximated by Sloan g′. A χ2 minimisation performed to identify allowed periods for the Hipparcos data. The reduced χ2 minimum is close to unity in both cases (Figure 2), so the periods where χ2 ≤ χ2 min + 1 can be used to approximate the period uncertainty. The dif- ferences between light curves with transiting planets and flat light curves are ∆χ2 = 12 and 16 for WASP-18 b and WASP-33 b, respectively, so the transits are detected with clear significance. The fitted periods and corresponding mid- transit times for this two-epoch fit are • P = 0.941 454 55 +0.000 000 87 −0.000 001 32 days, and MNRAS 000, 1–4 (2018) 10 5 0 -5 ) g a m m ( i e s o n r o h t p e D -10 -15 WASP-18 b HIP 7562 -20 0.94144 10 0.94145 0.94146 Period (days) 5 0 ) g a m m ( i e s o n -5 r o h t p e D -10 WASP-33 b HIP 11397 -15 1.21985 1.21986 1.21987 1.21988 Period (days) 1.04 1.02 1 0.98 0.96 2 χ r 0.94 0.92 0.9 0.88 0.94147 1.4 1.35 1.3 1.25 1.2 1.15 1.1 1.05 1 0.95 0.9 1.21989 2 χ r Figure 2. Goodness-of-fit of Hipparcos-derived periods. Dark, black lines show the mean transit depth (across t14) at that pe- riod; thin, red lines show the out-of-transit noise level. Dashed, blue lines show the reduced χ2 using the limb-darkened model (right axis). • T0 = 2 448 436.2359 +0.0125 −0.0082 for WASP-18 b and • P = 1.219 869 98 +0.000 000 79 • T0 = 2 448 472.5334 +0.0040 −0.0055 −0.000 000 57 days, and for WASP-33 b. each planet These mid-transit times represent observations taken 16 years (6145 and 4665 orbits) before those in the discovery papers of (Hellier et al. 2009; Collier Cameron et al. 2010), and more than double the length of their observational record to 24 and 23 years, re- spectively. To these transit times, we added the literature transit photometry collated for both WASP-18 b and WASP- 33 b (Wilkins et al. 2017 and Zhang et al. (2017), respec- tively), and created O − C diagrams for each planet (Figure 3). Unfortunately, the low cadence of the Hipparcos com- pared to modern data means that they do not provide con- straints greatly better than those available in the current lit- erature (Turner et al. 2016; Zhang et al. 2017; Wilkins et al. 2017). To fit the orbits, we ran two-parameter (T0, P ) and three-parameter (T0, P, δP/P ) Monte-Carlo χ2 fits to the observed mid-transit times. A two-parameter fit for this en- tire dataset formally provides • P = 0.941 452 67 ± 0.000 000 11 days, • T0 = 2 457 319.80197 ± 0.00021, and • χ2 r = 5.14 for WASP-18 b and • P = 1.219 870 61 ± 0.000 000 15 days, • T0 = 2 456 934.77020 ± 0.00010, and • χ2 r = 2.50 MNRAS 000, 1–4 (2018) Pre-discovery Hipparcos transits of exoplanets 3 ) s y a d ( e m i t t i s n a r t d e t u p m o c - d e v r e s b O ) s y a d ( e m i t t i s n a r t t d e u p m o c - d e v r e s b O WASP-18 b / HIP 7562 P = 0.94145267 d T0 = 57319.80197 dP/P = +/- 3 x 10-10 dP/P = -6 x 10-10; P = 0.94145186 d 50000 55000 JD - 2 400 000 (days) WASP-33 b / HIP 11397 P = 1.21987061 d T0 = 56934.77020 dP/P = +/- 3 x 10-10 0.015 0.01 0.005 0 -0.005 -0.01 -0.015 -0.02 -0.025 -0.03 0.002 0.001 0 -0.001 -0.002 0.01 0.008 0.006 0.004 0.002 0 -0.002 -0.004 -0.006 -0.008 -0.01 0.003 0.002 0.001 0 -0.001 -0.002 -0.003 50000 55000 JD - 2 400 000 (days) Figure 3. O − C diagrams for WASP-18 b and WASP-33 b, modelled against the best-fit ephemeris. Curves show models with period changes (dP/P ), as indicated on each plot. for WASP-33 b. These fits are shown in the O − C diagrams in Figure 3. A three-parameter fit formally provides: • δP/P = −6 ± 2 × 10−10, • P = 0.941 451 86 ± 0.000 000 23 days, • T0 = 2 457 319.80167 ± 0.00026, and • χ2 r = 4.64 for WASP-18 b and • δP/P = 2 ± 3 × 10−10, • P = 1.219 870 93 ± 0.000 000 50 days, • T0 = 2 456 934.77090 ± 0.00017, and • χ2 r = 2.78 for WASP-33 b. The fit for WASP-18 b is shown as the dotted line in Figure 3. 4 DISCUSSION AND CONCLUSIONS The reduced χ2 minimum of these fits is substantially greater than unity: in both bodies, an unmodelled scatter of around 0.001 days (1.44 minutes) is seen in the O − C di- agrams. This suggests that the errors quoted above are likely to be underestimates, either due to physical or unmodelled instrumental sources (cf. Adams et al. 2010; Barros et al. 2013). It also implies that either the photometric uncer- 4 I. McDonald et al. tainties on the input data are underestimated, or that an undetected third body in the system is causing TTVs. We used ttvfaster (Agol & Deck 2016) to model a third-body TTV signal to the data, assuming a circular or- bit, co-planar to the relevant planet. Unfortunately, the only ranges of parameters that can produce a sufficiently strong signal (∆TTTV & 0.0005 days) are of dynamically unstable systems, or those where a companion would be spectroscopi- cally detectable (e.g. a 0.25 M⊙ star in a 5.7-day orbit). Un- less cyclical variation of the planets' orbits are being driven by tidal interaction with their host stars, it appears that the uncertainties on the published transit times have been under-estimated in several cases, which could be due in part to microvariability on the host stars (von Essen et al. 2014). Given these under-estimated uncertainties, and the pos- sibility of other physical sources of TTV, the significance of the orbital change of either exoplanet cannot be precisely computed. Taking only the Hipparcos data at face value, we have a ∼1.0σ measurement of orbital expansion in WASP-33 b, and a ∼1.3σ measurement of orbital decay in WASP-18 b, depending on the exact period adopted. These are not significant detections. Strong orbital decay is not expected for these planets, as their host stars are relatively warm and have thin con- vective envelopes in which tides can be generated. The tidal quality factors for these stars are expected to be Q′ ∼ 108 (Barker & Ogilvie 2009). Due to its spin-orbit misalign- ment (Collier Cameron et al. 2010), non-radial changes to the orbit of WASP-33 may also be expected (Iorio 2011; Lin & Ogilvie 2017), causing more complex TTV signals over long periods. Using Equations 4 and 5 of (Wilkins et al. 2017), δP/P ∼ −6 × 10−10 implies Q′ ∼ 5 × 105 for WASP- 18. This is a much smaller value than nominally ex- pected (cf. Collier Cameron & Jardine 2018; Penev et al. 2018), but similar to that proposed for WASP-12 b by Maciejewski et al. (2016). However, it also implies a sur- vival time of < 106 years, thus the mere observable pres- ence of WASP-18 b means this value of δP/P is likely to be erroneously high. For WASP-33, δP/P < −1 × 10−10 implies Q′ > 2 × 105, which is not very limiting, but inter- esting given the visible tides the planet generates on its star (von Essen et al. 2014). A significant uncertainty driving the difference between the two- and three-parameter fits for WASP-18 b is the pe- riod in the current epoch, which differs by ∼ 8 × 10−7 days. A few high-precision measurements of transit times in the current epoch could greatly constrain these uncertainties, determining whether the offset of the Hipparcos datapoint in the O − C diagram is significant. We therefore strongly encourage monitoring of WASP-18 b, to more accurately de- termine its orbital period in the present epoch. ACKNOWLEDGEMENTS The authors acknowledge support from the UK Science and Technology Facility Council under grant ST/P000649/1. REFERENCES Adams E. R., L´opez-Morales M., Elliot J. L., Seager S., Osip D. J., 2010, ApJ, 714, 13 Agol E., Deck K., 2016, ApJ, 818, 177 Barker A. J., Ogilvie G. I., 2009, MNRAS, 395, 2268 Barros S. C. C., Bou´e G., Gibson N. P., Pollacco D. L., Santerne A., Keenan F. P., Skillen I., Street R. A., 2013, MNRAS, 430, 3032 Claret A., 2004, A&A, 428, 1001 Collier Cameron A., Jardine M., 2018, preprint, (arXiv:1801.10561) Collier Cameron A., et al., 2010, MNRAS, 407, 507 Gaia Collaboration et al., 2016, A&A, 595, A2 Han E., Wang S. X., Wright J. T., Feng Y. K., Zhao M., Fakhouri O., Brown J. I., Hancock C., 2014, PASP, 126, 827 H´ebrard G., Lecavelier Des Etangs A., 2006, A&A, 445, 341 Hellier C., et al., 2009, Nature, 460, 1098 Iorio L., 2011, Ap&SS, 331, 485 Lin Y., Ogilvie G. I., 2017, MNRAS, 468, 1387 Maciejewski G., et al., 2016, A&A, 588, L6 Mustill A. J., Villaver E., 2012, ApJ, 761, 121 Penev K., Jackson B., Spada F., Thom N., 2012, ApJ, 751, 96 Penev K., Bouma L. G., Winn J. N., Hartman J. D., 2018, preprint, (arXiv:1802.05269) Perryman M. A. C., ESA eds, 1997, The HIPPARCOS and TYCHO catalogues. Astrometric and photometric star cata- logues derived from the ESA HIPPARCOS Space Astrometry Mission ESA Special Publication Vol. 1200 Robichon N., Arenou F., 2000, A&A, 355, 295 Southworth J., 2008, MNRAS, 386, 1644 Steffen J. H., et al., 2013, MNRAS, 428, 1077 Turner J. D., et al., 2016, MNRAS, 459, 789 Wilkins A. N., Delrez L., Barker A. J., Deming D., Hamilton D., Gillon M., Jehin E., 2017, ApJ, 836, L24 Zhang M., et al., 2017, preprint, (arXiv:1710.07642) van Leeuwen F., 2007, A&A, 474, 653 von Essen C., et al., 2014, A&A, 561, A48 This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1–4 (2018)
1103.0010
1
1103
2011-02-28T21:00:06
A Search for Additional Planets in Five of the Exoplanetary Systems Studied by the NASA EPOXI Mission
[ "astro-ph.EP" ]
We present time series photometry and constraints on additional planets in five of the exoplanetary systems studied by the EPOCh (Extrasolar Planet Observation and Characterization) component of the NASA EPOXI mission: HAT-P-4, TrES-3, TrES-2, WASP-3, and HAT-P-7. We conduct a search of the high-precision time series for photometric transits of additional planets. We find no candidate transits with significance higher than our detection limit. From Monte Carlo tests of the time series using putative periods from 0.5 days to 7 days, we demonstrate the sensitivity to detect Neptune-sized companions around TrES-2, sub-Saturn-sized companions in the HAT-P-4, TrES-3, and WASP-3 systems, and Saturn-sized companions around HAT-P-7. We investigate in particular our sensitivity to additional transits in the dynamically favorable 3:2 and 2:1 exterior resonances with the known exoplanets: if we assume coplanar orbits with the known planets, then companions in these resonances with HAT-P-4b, WASP-3b, and HAT-P-7b would be expected to transit, and we can set lower limits on the radii of companions in these systems. In the nearly grazing exoplanetary systems TrES-3 and TrES-2, additional coplanar planets in these resonances are not expected to transit. However, we place lower limits on the radii of companions that would transit if the orbits were misaligned by 2.0 degrees and 1.4 degrees for TrES-3 and TrES-2, respectively.
astro-ph.EP
astro-ph
A Search for Additional Planets in Five of the Exoplanetary Systems Studied by the NASA EPOXI Mission Sarah Ballard1, Jessie L. Christiansen2, David Charbonneau1, Drake Deming3, Matthew J. Holman1, Michael F. A'Hearn4, Dennis D. Wellnitz4, Richard K. Barry3, Marc J. Kuchner3, Timothy A. Livengood3, Tilak Hewagama3,4, Jessica M. Sunshine4, Don L. Hampton5, Carey M. Lisse6, Sara Seager7, and Joseph F. Veverka8 ABSTRACT We present time series photometry and constraints on additional planets in five of the exoplanetary systems studied by the EPOCh (Extrasolar Planet Observation and Characterization) component of the NASA EPOXI mission: HAT-P-4, TrES-3, TrES- 2, WASP-3, and HAT-P-7. We conduct a search of the high-precision time series for photometric transits of additional planets. We find no candidate transits with sig- nificance higher than our detection limit. From Monte Carlo tests of the time series using putative periods from 0.5 days to 7 days, we demonstrate the sensitivity to detect Neptune-sized companions around TrES-2, sub-Saturn-sized companions in the HAT- P-4, TrES-3, and WASP-3 systems, and Saturn-sized companions around HAT-P-7. We investigate in particular our sensitivity to additional transits in the dynamically favor- able 3:2 and 2:1 exterior resonances with the known exoplanets: if we assume coplanar orbits with the known planets, then companions in these resonances with HAT-P-4b, WASP-3b, and HAT-P-7b would be expected to transit, and we can set lower limits on the radii of companions in these systems. In the nearly grazing exoplanetary systems TrES-3 and TrES-2, additional coplanar planets in these resonances are not expected to transit. However, we place lower limits on the radii of companions that would transit if the orbits were misaligned by 2.0◦ and 1.4◦ for TrES-3 and TrES-2, respectively. Subject headings: eclipses -- stars: planetary systems -- techniques: image processing -- techniques: photometric 1Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138; sbal- [email protected] 2NASA Ames Research Center, Moffett Field, CA 94035 3NASA/Goddard Space Flight Center, Greenbelt, MD 20771 4University of Maryland, College Park, MD 20742 5University of Alaska Fairbanks, Fairbanks AK 99775 6Johns Hopkins University Applied Physics Laboratory, Laurel, MD 20723 7Massachusetts Institute of Technology, Cambridge, MA 02139 8Cornell University, Space Sciences Department, Ithaca, NY 14853 -- 2 -- 1. Introduction EPOXI (EPOCh + DIXI) is a NASA Discovery Program Mission of Opportunity using the Deep Impact flyby spacecraft (Blume 2005). From January through August 2008, the EPOCh (Extrasolar Planet Observation and Characterization) Science Investigation used the HRI camera (Hampton et al. 2005) with a broad visible bandpass filter to gather precise, rapid cadence photo- metric time series of known transiting exoplanet systems. The majority of these targets were each observed nearly continuously for several weeks at a time. In Table 1 we give basic information about the seven EPOCh targets and the number of transits of each that EPOCh observed. One of the EPOCh science goals is a search for additional planets in these systems. Such planets would be revealed either through the variations they induce on the transits of the known exoplanet, or directly through the transit of the second planet itself. The search for additional planets in the EPOCh observations of the M dwarf GJ 436 was presented in Ballard et al. (2010). Because GJ 436 is a nearby M dwarf, it is the only EPOCh target for which we are sensitive to planets as small as 1.25 R⊕. In this work, we conduct a search for photometric transits of additional planets; the transit times of the known exoplanets observed by EPOCh are presented in Christiansen et al. (2011). The search for transiting planets in the EPOCh light curves is scientifically compelling be- cause the discovery of two transiting bodies in the same system permits the direct observation of their mutual dynamical interactions. This enables constraints on the masses of the two bodies independent of any radial velocity measurement (Holman & Murray 2005; Agol et al. 2005), as has been done for the multiple transiting planet system Kepler 9 (Holman et al. 2010). There are also separate motivations for searches for additional planets around the EPOCh targets. The search for additional transits in the EPOCh observations is complementary to existing constraints on ad- ditional planets from photometric observations, radial velocity measurements, and transit timing analyses of the known exoplanet. Here we briefly summarize such work to date for our five targets. Smith et al. (2009) in- vestigated 24 light curves of known transiting exoplanets, including the EPOXI targets HAT-P-4, TrES-2, WASP-3, and HAT-P-7, and found that they were sensitive to additional transits of Saturn- sized planets with orbital periods less than 10 days with greater than 50% certainty, although that probability is less for HAT-P-4 (Kov´acs et al. 2007) because of decreased phase coverage. Transit timing analyses of TrES-3b (O'Donovan et al. 2007) have ruled out planets in interior and exterior 2:1 resonances (Gibson et al. 2009), although the transit times obtained by Sozzetti et al. (2009) for TrES-3b may suggest a deviation from a constant period that could be attributed to an additional body. Freistetter et al. (2009) found that a broad range of orbits around TrES-2 (O'Donovan et al. 2006) would be dynamically stable for additional planets, although the constraints presented by Rabus et al. (2009) for TrES-2 have ruled out an 5 M⊕ planet in the 2:1 resonance specifically, and Holman et al. (2007) found no deviations in the transit timing residuals from the predicted ephemeris. Additionally, Raetz et al. (2009) observed a candidate transit in their photometry of -- 3 -- TrES-2 which might be attributed to an additional body in the system in an external orbit to TrES- 2b. However, Kipping & Bakos (2010) investigated the TrES-2 Kepler observations and found no unexpected photometric decrements and no significant transit timing or transit duration variation. Maciejewski et al. (2010) performed an analysis of the transit times of WASP-3b (Pollacco et al. 2008), and found evidence for planet with a mass of 15 R⊕ in a orbit close to a 2:1 resonance with the known planet. In the HAT-P-7 (P´al et al. 2008) radial velocity measurements, Winn et al. (2009) found a drift that provides evidence for a third body. This radial velocity trend is consistent with any period longer than a few months. Finally, the light curves obtained by the Kepler Mis- sion (Borucki et al. 2010) will ultimately enable exquisite constraints on the presence of additional planets in two of the systems which were also observed by EPOXI: TrES-2 and HAT-P-7. The remainder of this paper is organized as follows. In Section 2, we describe the photometry pipeline we created to produce the time series. In Section 3, we describe the search we conduct for additional transiting planets. We present a Monte Carlo analysis of the EPOCh observations of HAT-P-4, TrES-3, TrES-2, WASP-3, and HAT-P-7, and demonstrate the sensitivity to detect additional planets in the Neptune-sized and Saturn-sized radius ranges. In Section 4, we present our best candidate transit signals, and from the search for additional transits we place upper limits on the radii of additional putative planet in these systems in the exterior 3:2 and 2:1 resonances with the known exoplanets. 2. Observations and Data Reduction The photometric pipeline we developed for the EPOCh data is discussed at length in Ballard et al. (2010) (concerning GJ 436 in particular), and is summarized in Christiansen et al. (2010) (concern- ing HAT-P-7) and Christiansen et al. (2011) (concerning HAT-P-4, TrES-3, TrES-2, and WASP-3). We outline here the basic steps we undertake to produce the final EPOCh time series. We acquired observations of the five EPOCh targets presented here nearly continuously for approximately two- week intervals. These intervals were interrupted for several hours at approximately 2-day intervals for data downloads. We also obtained for TrES-2, WASP-3, and HAT-P-7 approximately one day of "pre-look" observations, implemented to optimize pointing for each target, that predate the contin- uous observations by a week. The basic characteristics of the targets and observations are given in Tables 1 and 2. Observations of this type were not contemplated during development of the original Deep Impact mission; the spacecraft was not designed to maintain very precise pointing over the timescale of a transit (Table 2). Furthermore, the available onboard memory precludes storing the requisite number of full-frame images (1024×1024 pixels). Hence, the observing strategy during the later observations (TrES-2, WASP-3, and HAT-P-7) used 256×256 sub-array mode for those times spanning the transit, and 128×128 otherwise. This strategy assured complete coverage at transit, with minimal losses due to pointing jitter exceeding the 128×128 sub-array at other times. We elected to exclude the following EPOCh data from the search for additional transits: first, the observations of XO-2, for which we gathered only partial transits and had relatively sparse phase -- 4 -- coverage due to pointing jitter and data transfer losses. Second, we did not use the observations from the second EPOCh campaign for HAT-P-4 (from 29 Jun - Jul 7 2008), which we could not calibrate to the same level of precision as the original observations for reasons explained below. Our sensitivity to additional transit signals in the HAT-P-4 light curve, which should theoretically have improved with additional observations removed in time, was in reality diminished due to the in- creased correlated noise in the second campaign HAT-P-4 observations. For this reason, we elected to use only the original 22 days of observations in the search for additional transits. We used the existing Deep Impact data reduction pipeline to perform bias and dark subtrac- tions, as well as preliminary flat fielding (Klaasen et al. 2005). We first determined the position of the star on the CCD using PSF fitting, by maximizing the goodness-of-fit (with the χ2 statistic as an estimator) between an image and a model PSF (oversampled by a factor of 100) with variable position, additive sky background, and multiplicative brightness scale factor. We then processed the images to remove sources of systematic error due to the CCD readout electronics. We first scaled down the two central rows by a constant value, then we scaled down the central columns by a separate constant value. Finally, in the case of 256×256 images, we scaled the entire image by a multiplicative factor to match the 128×128 images (the determination of the optimal correction values is performed independently for each target). We performed aperture photometry on the corrected images, using an aperture radius of 10 pixels, corresponding to twice the HWHM of the PSF. To remove remaining correlated noise due to the interpixel sensitivity variations on the CCD, we fit a 2D spline surface to the brightness variations on the array as follows. We randomly drew a subset of several thousand out-of-transit and out-of-eclipse points from the light curve (from a data set ranging from 11,000 total points in the case of TrES-3 to 20,000 points in the case of HAT-P-4), recorded their X and Y positions, and calculated a robust mean of the brightness of the 30 nearest spatial neighbors for each selected point. To determine the robust mean, we used an iterative sigma-clipping routine that recalculates the mean after excluding outliers further than 3 sigma from the mean estimate at each iteration (the routine concludes after the iteration when no new outliers are identified). Given the set of X and Y positions and the average brightness values of the 30 points which lie nearest those positions, we fit a spline surface to the brightness Table 1. EPOCh Targets Name V Magnitude Number of Transits Observeda Dates Observed [2008] HAT-P-4 TrES-3 XO-2 GJ 436 TrES-2 WASP-3 HAT-P-7 11.22 11.18 12.40 10.67 11.41 10.64 10.50 aSome transits are partial. 10 7 3 8 9 8 8 Jan 22 -- Feb 12, Jun 29 -- Jul 7 Mar 8 -- March 10, March 12 -- Mar 18 Mar 11, Mar 23 -- Mar 28 May 5 -- May 29 Jun 27 -- Jun 28, Jul 9 -- Jul 18, Jul 21 -- Aug 1 Jul 18 -- Jul 19, Aug 1 -- Aug 9, Aug 11 -- Aug 17 Aug 9 -- Aug 10, Aug 18 -- Aug 31 -- 5 -- variations in X and Y using the IDL routine GRID_TPS. This spline surface has the same resolution as the CCD, and approximates a flat field of the CCD which has been convolved by a smoothing kernel with width equal to the average distance required to enclose 30 neighboring points. We then corrected each data point individually by bilinearly interpolating on the spline surface to find the expected brightness of the star at each X and Y position. We then divide each observation by its expected brightness to remove the effects of interpixel sensitivity variations. We used only a small fraction of the observations to create the spline surface in order to minimize the potential transit signal suppression introduced by flat fielding the data with itself. To produce the final time series, we iterated the above steps, fitting for the row and column multiplicative factors, the sub-array size scaling factor, and the 2D spline surface that minimized the out-of-transit standard deviation of the photometric time series. We include two additional steps in the reduction of these data that were not included in the Ballard et al. (2010) reduction of the GJ 436 EPOCh observations. First, during the second cam- paign observations of HAT-P-4 and TrES-2, we observed an increase in brightness when the position of the star was located in the lower right-hand quadrant of the CCD. At the image level, we ob- served a bright striping pattern in this quadrant that caused the measured brightness of the star to increase as soon as the PSF entered this quadrant. We found that the dependence of the brightness increase in this quadrant was correlated with the Instrument Control Board Temperature value recorded for for each image in the FITS header. For the HAT-P-4 second campaign and TrES-2 observations, we first fit a spline to the dependence of the photometric residuals (after the boot- strap flat field was applied, described below) on the Instrument Control Board Temperature, using residuals for which the entire PSF of the star fell into the CCD quadrant in question. The most egregious brightness increase is 4 mmag, when this effect was most prominent. We then performed aperture photometry again on these targets and corrected each image by interpolating the Instru- ment Control Board Temperature for that image onto the spline, multiplying this correction value by the fraction of the PSF core that fell into the quadrant in question, and dividing this value from the photometry. We found that this iterative procedure largely removed this quadrant-dependent effect. In the latter half of TrES-2 observations, we no longer observed the brightness increase in this CCD quadrant. Therefore, we found that the correction procedure was only necessarily for the second campaign HAT-P-4 and the first portion of TrES-2 observations. However, because of the 6-month separation of the second campaign HAT-P-4 observations from the original HAT-P-4 observations, the behavior of the CCD had sufficiently altered to disallow the combination of the data into a single 2D spline surface. The separate 2D spline correction of the second campaign observations (spanning only 8 days), coupled with the residual striping artifacts, sufficiently de- creased the precision of the second campaign observations that we elected to exclude them from the search for additional transiting planets around HAT-P-4. Secondly, we include one final correction after we have removed the interpixel brightness vari- ations with the 2D spline, which is to perform an additional point-by-point correction to the data taken during transit and secondary eclipse of the known exoplanets. The bootstrap flat field ran- -- 6 -- domly selects a set of points to create the spline surface, instead of using all the data to create this surface; this minimizes the suppression of additional transits. Our sensitivity to additional transits is sufficiently diminished during the transits of the known planet that we are concerned more with removing correlated noise, and less concerned about avoiding additional transit suppression. The two reasons for the diminished sensitivity during transit windows are, first, that we fit a slope with time to the points immediately outside of the transit of the known exoplanet (from 3 minutes to 30 minutes before and after transit) and divide by the slope in order to normalize each transit before we fit for the system parameters (this procedure is also detailed in Ballard et al. 2010). This could have the possibility of removing a decrement due to an additional transit. Second, there is also the possibility of an occultation of one planet by another. We therefore elected to perform a point-by-point correction after the 2D spline for points occurring during the transit and eclipse of the known exoplanet, wherein we find a robust mean of the 30 nearest neighbors to each point (using the same iterative routine described above) and divide this value individually for each point in transit or eclipse. This has the benefit of removing additional correlated noise during transit and eclipse, while still minimizing signal suppression of additional putative planets outside of these time windows After we take these steps to address the systematics associated with the observations, we achieve a precision for the unbinned observations which is approximately twice the photon noise limited precision for all five targets. We estimate the photon limited precision at the image level, by converting the stellar flux from watts per meter squared per steradian per micron (W/m2·sr·µm) to electron counts as follows. We first divide by the conversion factor in the FITS header (keyword RADCALV = 0.0001175 W/m2·sr·µm per DN/s), then multiply by the exposure time (INTTIME = 50.0005 s), and finally multiply by the gain (28.80 e/DN, per Klaasen et al. 2008 for the HRI camera). We then estimate the photometric error by calculating 1/√N , where N is the number of electrons. We have excluded read noise, bias, and dark current from the estimation of the photon limited precision because these quantities contribute negligibly to the total number of electrons measured within the aperture; we briefly summarize our reasoning here. Using the results of the calibration tests on the HRI instrument shown in Klaasen et al. (2008), we estimate the read noise and dark current (given the CCD temperature of 160 K, as recorded in the image headers) to contribute less than a DN. Calculating the median bias value per pixel from the overclocked pixels associated with each image, and then multiplying this median value by the number of pixels contained within the aperture, we determine that the bias contributes less than 500 DN. When compared to the total measured DN flux contained in the aperture, which is of order 105 DN for the dimmest target star, TrES-3, we conclude that read noise, bias, and dark current are negligible. We repeat the photon limited precision calculation on 50 images for each target, and take the mean of these values to be our estimate for the photon limited precision. Our precision of 1.21 mmag for HAT-P-4 is 94% above the limit, 2.17 mmag for TrES-3 is 106% above the limit, 1.62 mmag for TrES-2 is 136% above the limit, 0.97 mmag for WASP-3 is 106% above the limit, and 0.86 mmag for HAT-P-7 is 91% above the limit. The EPOCh precision for GJ 436 of 0.51 mmag was only 56% above the photon noise limit, which we attribute to the longer baseline of observations with -- 7 -- fewer gaps in phase coverage, both of which enabled us to create a higher quality 2D spline flat field (Ballard et al. 2010). Figure 1 shows five EPOCh time series after the 2D spline correction is applied; these light curves are identical to the ones presented in Christiansen et al. (2010, 2011). In the right panel adjacent to each time series, we show how the time series, after the 2D spline is applied, bins down as compared to Gaussian noise over timescales of 7 hours (512 points) or less. We selected the longest contiguous portion of the lightcurve between transits (and excluding secondary eclipse) to calculate the standard deviation as a function of binsize -- this unbinned portion typically comprises about 2500 points. We compare the expected Gaussian scatter for a bin size of 1 hour (assuming that sigma decreases as N −1/2, and normalizing to the observed rms of the unbinned data) to the measured scatter, and find that the presence of correlated noise inflates the 1σ error bar by a factor of 1.86 for HAT-P-4, 1.58 for TrES-3, 1.90 for TrES-2, 2.77 for WASP-3, and 3.14 for HAT-P-7 for 1 hour timescales. We also investigate the transit signal suppression introduced by using a flat field created from the out-of-transit and out-of-eclipse data itself. We avoid the suppression of known transits in each data set by iteratively excluding those observations (using an ephemeris for the known planet derived from the EPOCh observations) from the points used to generate the flat field surface, so that we only use the presumably constant out-of-transit and out-of-eclipse observations to sample the CCD sensitivity. However, if the transit of an additional planet occurs while the stellar PSF is lying on a part of the CCD that is never visited again, the 2D spline algorithm instead models the transit as diminished pixel sensitivity in that CCD location. To quantify the suppression of additional transits that result from using the 2D spline, we inject transit light curves with periods ranging from 0.5 days to 7 days in intervals of 30 minutes in phase (ranging from a phase of zero to a phase equal to the period) into the EPOCh light curve just prior to the 2D spline step. After performing the 2D spline, we then phase the data at the known injected period and fit for the best radius, using χ2 as the goodness-of-fit statistic. We show in Figure 2 the radius suppression as a function of period for five EPOCh targets. The HAT-P-4 observations occur over a longer duration with less gaps in phase coverage, so even at a period of 7 days, we have 95% confidence that the radius of an additional transiting planet will not be suppressed to less than 60% its original value. For example, an additional 8 R⊕ planet orbiting HAT-P-4 will appear no smaller than 0.6×8 R⊕, or 4.8 R⊕, with 95% confidence. However, for a target with sparser phase coverage, such as WASP-3, we have 95% confidence that the radius will not be suppressed to less than 45% its original value. The same 8 R⊕ planet orbiting WASP-3 will therefore appear no smaller than 0.45×8 R⊕, or 3.6 R⊕, with the same confidence. The average (50% confidence) suppression value of 75% across all periods and for all targets reflects the average density of points on the CCD (and thus the quality of the 2D spline), which is indicative of the pointing jitter of the instrument. We describe our incorporation of signal suppression into our search for additional planets in greater detail in Section 3.2. -- 8 -- Fig. 1. -- Left panels: EPOXI time series for targets HAT-P-4, TrES-3, TrES-2, WASP-3, and HAT-P-7. For TrES-3 (second panel from top), we show the light curve with original modulation due to star spots at bottom. Right panels: The standard deviation versus bin size for each target, compared to the ideal Gaussian limit (shown with a line, normalized to match the value at N=1). -- 9 -- Fig. 2. -- The 50% and 95% confidence values for suppression of additional transits as a function of orbital period in the EPOCh observations. We have 50% confidence that the transit signal will not be suppressed more than the value of the dashed line at that period, and 95% confidence that the transit signal will not be suppressed more than the value of the solid line. -- 10 -- 3. Analysis 3.1. Search for Additional Transiting Planets We search the EPOXI time series for evidence for additional shallow transits. We developed software to search for these additional transits, which is discussed at length in Ballard et al. (2010). The steps involved in the procedure are summarized in this section. We conduct a Monte Carlo analysis to assess how accurately we could recover an injected planetary signal in each of the EPOCh light curves. We evaluate our sensitivity to transit signals on a grid in radius and period space sampled at regular intervals in R2 P and regular frequency spacing in P . We create an optimally spaced grid as follows: for the lowest period at each radius, we determine the radii at which to evaluate the adjacent periods by solving for the radius at which we achieve equivalent signal-to- noise (for this reason, we expect significance contours to roughly coincide with the grid spacing). We use the Mandel & Agol (2002) routines for generating limb-darkened light curves given these parameters to compute a grid of models corresponding to additional possible planets. If we make the simplifying assumptions of negligible limb darkening of the host star, a circular orbit, and an orbital inclination angle i of 90◦, the set of light curves for additional transiting bodies is a three parameter family. These parameters are radius of the planet Rp, orbital period of the planet P , and orbital phase φ. To generate these light curves, we also use the stellar radii values determined by Christiansen et al. (2010, 2011) from the EPOXI data, with the corresponding stellar masses, from the literature, that were used to calculate those radii. Those radius and mass values are 1.60 R⊙ and 1.26 M⊙ (Kov´acs et al. 2007) for HAT-P-4, 0.82 R⊙ and 0.93 M⊙ (Sozzetti et al. 2009) for TrES-3, 0.94 R⊙ and 0.98 M⊙ (Sozzetti et al. 2007) for TrES-2, 1.35 R⊙ and 1.24 M⊙ (Pollacco et al. 2008) for WASP-3, and 1.82 R⊙ and 1.47 M⊙ (P´al et al. 2008) for HAT-P-7. At each test radius and period, we inject planetary signals with 75 randomly assigned phases into the residuals of EPOCh light curves with the best transit model divided out, and then attempt to recover blindly the injected signal by minimizing the χ2 statistic. The period range of injected signals is selected for each target individually, to ensure the injected transit signal comprises at least two transits in most cases. For a target with high phase coverage, like HAT-P-4, we inject signals with periods up to 7 days, but for targets with observations of a shorter duration and sparser phase coverage, like TrES-3 or WASP-3, we inject signals up to 3.5 and 2.5 days, respectively. For planets with periods higher than these values, we may detect the planet, but with a single transit, we expect only a very weak constraint on the period. We first conduct a coarse χ2 grid search in radius, period, and phase. We select the spacing of this grid to minimize processing time while ensuring that the transit was not missed; we polish the parameters with a finer χ2 search after the initial coarse search. We sample the χ2 space at 300 points in period space (at even frequency intervals between 0.5 and 8.5 days), 50 points in radius space (between 0.5 and 5.5 Earth radii) and a variable number of points in phase space set by the resolution of the transit duration for each period. We use an expression for the transit duration τ given by Seager & Mall´en-Ornelas (2003): -- 11 -- sin i sin(cid:16) πτ P (cid:17) =s(cid:18) R⋆ + RP a (cid:19)2 − cos2i. (1) For each test model, we compute the χ2, using the out-of-transit standard deviation to estimate the error in each point. After the grid χ2 minimum is determined, we use the amoeba minimization routine (Nelder & Mead 1965) to more finely sample the χ2 space in order to find the χ2 minimum from the specified nearest grid point. We also investigate whether aliases of the best-fit period from the χ2 grid improve the fit. We find that roughly half of the best solutions from the grid are aliases of the injected period, most at either half or twice the value of the injected period, but we test aliases at every integer ratio from 1/35 to 35 times the given period (although aliases other than 1:2, 2:1, 3:1, 1:3, 2:3, or 3:2 occurs less than 3% of the time for all targets). We also repeat the finer grid search at the three next lowest χ2 minima, in case the best solution (or an alias of the best solution) lies closer to that grid point. For all injected signals, we recover a solution which is a better fit (in the χ2 sense) than the injected signal. For this reason, we are confident that we are sampling the χ2 space sufficiently finely to locate the best solution. We quantify the success of this analysis by how well the search blindly recovers the known injected transit signal. We define the error on the recovered parameter, for instance period, to be Pinjected − Pobserved /Pinjected. Figure 3 shows this relative error in radius, with 95% confidence, for all searches. As we note in the last paragraph of Section 2, we anticipate suppression of additional transit signals from the bootstrap flat field treatment of the EPOXI data. We evaluate the suppression we expect at the period values used in the Monte Carlo analysis, using the results shown in Figure 2. We incorporate this expected suppression by vertically shifting the effective radius values of the grid points at which we evaluate our sensitivity to additional transits. For example, for HAT-P-4 at 1.63 days, all grid points have been shifted upward in radius by a value of 1/0.7, or 1.42, because we anticipate that the radius will be suppressed to no smaller than 70% its original value. For this reason, the recovery statistics corresponding to a 3.0 R⊕ transit depth in the final light curve would be accurate for an original transit signal of a 4.3 R⊕ planet once we fold our expectation of signal suppression. We also evaluate the overall detection probability for putative transiting planets. Given the cadence and coverage of the EPOXI observations, we determine the number of in-transit points we expect for a given radius, period, and phase (where the phase is evaluated from 0 to 1 periods, in increments of 30 minutes). We then evaluate the expected significance of the detection, assuming a boxcar-shaped transit at the depth of (RP /R⋆)2, and the standard deviation of the time series. At each phase and period individually, we scale down RP to incorporate the signal suppression at that ephemeris. We use the improvement in the χ2 over the null hypothesis to define a positive detection, after we have removed the best candidate transit signal (described in Section 4.1). If we do not first remove this signal, then we are a priori defining a "detectable" signal to be any signal more prominent than the best candidate signal, and we would be unable to evaluate this signal's authenticity. We set our detection limit at an improvement in χ2 over the null hypothesis that -- 12 -- Fig. 3. -- Constraints on radius from the Monte Carlo analysis. For each point in radius and period, we create 75 light curves with random orbital phases, inject them into the EPOCh residuals, and attempt to recover them blindly. The diamonds indicate the grid of radii and periods at which we evaluate our sensitivity; the contours are produced by interpolating between these points. The contours indicate the relative error in radius (absolute value of recovered-injected/injected radius) that encloses 95% of the results. -- 13 -- signifies a correctly recovered period (which we define as a period error of <1%). This number is variable among the EPOCh targets due to the precision of the observations and the contamination of correlated noise. The detection probability of additional transiting planets, as a function of their radius and orbital period, is shown in Figure 4. For the HAT-P-4, TrES-3, and TrES-2, the ∆χ2 cutoff is set at 250, 200, and 200 respectively. For WASP-3, the ∆χ2 criterion for detection is 400, and for HAT-P-7, the cutoff is 500. There are five exceptions here for these threshold values across the five targets: in our analysis of WASP-3, we find two instances of a significance higher than 400, but an incorrect period value: these signals both comprise 4 full transit events, and are recovered at a 4:1 alias of the true period of 2.34 days. Due to the same instances of correlated noise that produce two positive deviations during two of the four transit events that decrease the depths by 2 mmag, a better solution is found at an alias of 4:1 that at the true period. For HAT-P-7, we find three similar cases of a 2:1 alias providing a better solution than the injected period, although the significance of the detection is above the threshold vale of 500. For these three injected signals, which comprise three transits, two of the transits are recovered correctly, and the third overlays a single instance of correlated noise that decreases the depth of the transit by 0.5 mmag. We investigated the true signal-to-noise ratio (including correlated noise) associated with a single detectable transit with the cutoff ∆χ2 significance for each target. We use a method similar to the one described by Winn et al. (2008) to determine the contribution of correlated noise to the standard deviation over a transit duration timescale. We first solved for the transit depth associated with the cutoff ∆χ2 value, assuming a single boxcar transit with standard deviation equal to the out-of-transit and out-of-eclipse standard deviation of the unbinned time series. We next found the standard deviation at a bin size corresponding to a transit duration for each target. We assume an edge-on transit (which assumption we also used for the Monte Carlo analysis) and the shortest period where we expect mostly single transits (this period is slightly larger than the largest period used for the Monte Carlo analysis; which period range was selected so that we would expect at least two transits in nearly all cases). This approximate orbital period is 7.5 days for HAT-P-4, 4.0 days for TrES-3, 5.0 days for TrES-2, 3.0 days for WASP-3, and 5.0 days for HAT-P-7. Using the cutoff transit depth and the standard deviation associated with the transit duration for each target, we find that the signal-to-noise ratio associated with the detection criteria is approximately constant across the targets, ranging between 5 and 8. The variation in the ∆χ2 value can be attributed in part to the varying presence of correlated noise in the different data sets (and also to the number of points associated with each transit, which depends on the transit duration). We confirm empirically that planets of these radii are detectable by examining the detection probability as a function of radius and orbital period shown in Figure 4. We convert the cutoff transit depth to a planetary radius, assuming the stellar radius derived from the EPOCh observations and average suppression of the radius to 0.75 its original value (roughly constant for all EPOCh targets, as shown in Figure 2). This radius value physically corresponds to the minimum planetary radius detectable by EPOCh from a single transit. This value is 7.1 R⊕ for HAT-P-4, 6.2 R⊕ for TrES-3, -- 14 -- 5.3 R⊕ for TrES-2, 6.5 R⊕ for WASP-3, and 7.9 R⊕ for HAT-P-7. Comparing to the nearest radius value in Figure 4, we find that indeed, at the shortest orbital period where we should expect to see single transits, we can detect a planet with radius associated with the detection criteria at high significance. At longer orbit periods, we still expect single transits, but the likelihood that the single transit occurs during a gap in the phase coverage increases. 4. Discussion 4.1. Best Candidate Transit Signals We present our best candidate transits here, for each of the five EPOCh targets. Figure 5 shows each of the individual candidate transit events that comprise the best candidate signal, as well as the entire phased and binned signal. For HAT-P-4, the best candidate is a 2.7 R⊕ planet in a 3.1 day orbit; the ∆χ2 significance is 61 (as compared to a detection criterion of 250). For TrES-3, the best candidate is a 2.9 R⊕ planet in a 2.63 day orbit; the ∆χ2 significance is 87 (as compared to a detection criterion of 200). For TrES-2, the best candidate is a 3.6 R⊕ planet with a period of 7.22 days; the significance is ∆χ2 of 269 (as compared to a detection criterion of 200). For WASP-3, the best candidate is a 4.2 R⊕ with a period of 5.9 days; the ∆χ2 significance is 232 (as compared to a detection criterion of 400). For HAT-P-7, the best candidate is a 4.4 R⊕ planet with a 3.9 day orbit; the significance of the detection is a ∆χ2 of 201 (as compared to a detection criterion of 500). The only candidate signal above the ∆χ2 detection threshold is the one in the TrES-2 light curve; this candidate signal comprises two transit events (the other predicted events occur during gaps in the phase coverage). One of the candidate transit events occurs in a portion in the CCD that is never visited afterward; these data are therefore uncalibrated by the 2D spline flat field and are unreliable. Without the transit signal that occurs in the uncalibrated area of the CCD, the ∆χ2 significance of the remaining transit is 80, which is well below the detection threshold of 200. 4.2. Radius constraints From the results of our Monte Carlo analysis and phase coverage analysis, we can rule out tran- siting planets in the sub-Saturn radius range for HAT-P-4, TrES-3, and WASP-3, the Saturn-sized radius range for HAT-P-7, and Neptune-sized radius range for TrES-2. We consider in particular our sensitivity to additional planets in the dynamically favorable 3:2 and 2:1 resonance orbits with the known exoplanets. In Figure 4, we show the detection probability as a function of period for planets ranging in size from 3 to 10 R⊕, with positions of the exterior 3:2 and 2:1 resonances marked by vertical lines. We also indicate in Figure 4 the regions not guaranteed to be stable by Hill's criterion, per the formula given in Gladman (1993). Assuming an eccentricity of zero for both the known and putative additional planet, the planetary orbits are assumed to be stable if the following -- 15 -- condition holds, where µ1 = m1/Mstar, µ2 = m2/Mstar, α = µ1 + µ2, and δ =pa2/a1: µ1µ2 α4/3 , (2) µ2 α−3(cid:16)µ1 + δ2(cid:17) (µ1 + µ2δ)2 > 1 + 34/3 · We solve numerically for the boundaries in δ of the stable region, using the stellar masses and masses for the known planets given by Kov´acs et al. (2007) for HAT-P-4b, Sozzetti et al. (2009) for TrES-3b, Sozzetti et al. (2007) for TrES-2b, Pollacco et al. (2008) for WASP-3, and P´al et al. (2008) for HAT-P-7, and conservatively using a putative mass for the second body equal to the mass of Saturn. This results in an overestimate of the extent of the Hill-unstable region for the planets with masses smaller than Saturn; while we find we are sensitive to planets with radii well below that of Saturn, the mass of putative additional planets depends on their composition and is uncertain. However, the critical δ values vary slowly with increased mass of the putative additional planet, so that increasing the mass to that of Jupiter changes the periods associated with the closest Hill-stable orbits by only 7% at most for these systems. In some cases, the 3:2 orbital resonance is not guaranteed to be Hill-stable (though it may be stable); the exact boundary of the stable region depends on the mass we assume for the additional planet. From the detection probabilities shown in Figure 4, in the HAT-P-4 system, we are sensitive to planets as small as 8 R⊕ in the 3:2 and 2:1 resonance with HAT-P-4b (with a period of 3.06 days) with 95% confidence. In the TrES-3 system, with the known exoplanet in a 1.3 day orbit, we would have detected a 5 R⊕ planet in the 3:2 and 1:2 resonance with 70% and 50% probability, respectively, and an 8 R⊕ in either orbit with nearly 100% probability. Around TrES-2 we are sensitive to the smallest planets, and would have detected a 4 R⊕ planet with 65% probability in the 3:2 resonance with TrES-2b (which has a period of 2.47 days). In both the 3:2 and 2:1 resonances, we had a high probability of detecting a 5.0 R⊕ planet: >95% in the case of the 3:2 resonance, and 90% in the 2:1 resonance. Around WASP-3, we had 50% chance of detecting a 5.0 R⊕ planet in the 3:2 resonance with WASP-3b (which has a period of 1.85 days), and would have seen a planet as small as 8 R⊕ in either the 3:2 or 2:1 resonance with >95% probability. Around HAT-P-7, we would have detected a Saturn-sized 10 R⊕ planet in either the 3:2 or 2:1 resonances with 95% probability, and had a 70% chance of detecting an 8 R⊕ planet. If we assume an inclination equal to that of the known exoplanet, we can rule out additional transiting planets of HAT-P-4, WASP-3, and HAT-P-7 in the 3:2 and 2:1 resonances of the sizes stated above, as we still expect additional planets to transit at those orbital distances. However, the known exoplanets in both the TrES-3 and TrES-2 systems are already in grazing orbits, so additional planets in the exterior 3:2 and 2:1 resonances would not be expected to transit if they were strictly coplanar with the known exoplanet. However, if the orbits of additional planets were misaligned by 2.0◦ in the case of TrES-3 and 1.4◦ in the case of TrES-2 (using the planetary inclinations and stellar radii from Christiansen et al. 2010, 2011) then we would observe a transit in both of the 3:2 and 2:1 exterior resonances. The orbital inclinations of the ice and gas giants in our solar system vary by up to nearly 2◦ (Cox 2000), so it is feasible that an additional planet in these systems could transit. -- 16 -- 5. Acknowledgments We are extremely grateful to the EPOXI Flight and Spacecraft Teams that made these difficult observations possible. At the Jet Propulsion Laboratory, the Flight Team has included M. Abra- hamson, B. Abu-Ata, A.-R. Behrozi, S. Bhaskaran, W. Blume, M. Carmichael, S. Collins, J. Diehl, T. Duxbury, K. Ellers, J. Fleener, K. Fong, A. Hewitt, D. Isla, J. Jai, B. Kennedy, K. Klassen, G. LaBorde, T. Larson, Y. Lee, T. Lungu, N. Mainland, E. Martinez, L. Montanez, P. Morgan, R. Mukai, A. Nakata, J. Neelon, W. Owen, J. Pinner, G. Razo Jr., R. Rieber, K. Rockwell, A. Romero, B. Semenov, R. Sharrow, B. Smith, R. Smith, L. Su, P. Tay, J. Taylor, R. Torres, B. Toyoshima, H. Uffelman, G. Vernon, T. Wahl, V. Wang, S. Waydo, R. Wing, S. Wissler, G. Yang, K. Yetter, and S. Zadourian. At Ball Aerospace, the Spacecraft Systems Team has included L. Andreozzi, T. Bank, T. Golden, H. Hallowell, M. Huisjen, R. Lapthorne, T. Quigley, T. Ryan, C. Schira, E. Sholes, J. Valdez, and A. Walsh. Support for this work was provided by the EPOXI Project of the National Aeronautics and Space Administration's Discovery Program via funding to the Goddard Space Flight Center, and to Harvard University via Co-operative Agreement NNX08AB64A, and to the Smithsonian Astro- physical Observatory via Co-operative Agreement NNX08AD05A. Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS, 359, 567 REFERENCES Ballard, S., et al. 2010, ApJ, 716, 1047 Blume, W. H. 2005, Space Science Reviews, 117, 23 Borucki, W. J., et al. 2010, Science, 327, 977 Christiansen, J. L., et al. 2011, ApJ, 726, 94 -- . 2010, ApJ, 710, 97 Cox, A. N. 2000, Allen's Astrophysical Quantities, ed. A. N. Cox Freistetter, F., Suli, ´A., & Funk, B. 2009, Astronomische Nachrichten, 330, 469 Gibson, N. P., et al. 2009, ApJ, 700, 1078 Gladman, B. 1993, Icarus, 106, 247 Hampton, D. L., Baer, J. W., Huisjen, M. A., Varner, C. C., Delamere, A., Wellnitz, D. D., A'Hearn, M. F., & Klaasen, K. P. 2005, Space Science Reviews, 117, 43 Holman, M. J., et al. 2010, Science, 330, 51 -- 17 -- Holman, M. J., & Murray, N. W. 2005, Science, 307, 1288 Holman, M. J., et al. 2007, ApJ, 664, 1185 Kipping, D. M., & Bakos, G. ´A. 2010, ArXiv e-prints Klaasen, K. P., et al. 2008, Review of Scientific Instruments, 79, 091301 Klaasen, K. P., Carcich, B., Carcich, G., Grayzeck, E. J., & McLaughlin, S. 2005, Space Science Reviews, 117, 335 Kov´acs, G., et al. 2007, ApJ, 670, L41 Maciejewski, G., et al. 2010, MNRAS, 407, 2625 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Nelder, J. A., & Mead, K. 1965, Computer Journal, 7, 308 O'Donovan, F. T., et al. 2007, ApJ, 663, L37 -- . 2006, ApJ, 651, L61 P´al, A., et al. 2008, ApJ, 680, 1450 Pollacco, D., et al. 2008, MNRAS, 385, 1576 Rabus, M., Deeg, H. J., Alonso, R., Belmonte, J. A., & Almenara, J. M. 2009, A&A, 508, 1011 Raetz, S., et al. 2009, Astronomische Nachrichten, 330, 459 Seager, S., & Mall´en-Ornelas, G. 2003, ApJ, 585, 1038 Smith, A. M. S., et al. 2009, MNRAS, 398, 1827 Sozzetti, A., Torres, G., Charbonneau, D., Latham, D. W., Holman, M. J., Winn, J. N., Laird, J. B., & O'Donovan, F. T. 2007, ApJ, 664, 1190 Sozzetti, A., et al. 2009, ApJ, 691, 1145 Winn, J. N., et al. 2008, ApJ, 683, 1076 Winn, J. N., Johnson, J. A., Albrecht, S., Howard, A. W., Marcy, G. W., Crossfield, I. J., & Holman, M. J. 2009, ApJ, 703, L99 This preprint was prepared with the AAS LATEX macros v5.2. -- 18 -- Table 2. Characteristics of the EPOCh Observations Telescope aperture Spacecraft memory Bandpass Integration time Pointing jitter Defocus Pixel scale 30 cm 300 Mb 350-1000 nm 50 seconds ± 20 arc-sec per hour 4 arc-sec FWHM 0.4 arc-sec per pixel Subarray size 256×256 pixels spanning transit, 128×128 otherwisea aWith the exception of the HAT-P-4 observations during 2008 January and February and TrES-3 observations, which were conducted entirely in 128×128 subarray mode. -- 19 -- Fig. 4. -- Detection probability versus period for planets ranging in size from 3 to 10 R⊕. The detection criteria is set by the percentage of phases at a given period for which the number of points observed in transit produces a χ2 improvement of the cutoff significance, compared to the null hypothesis (∆χ2 of 250 for HAT-P-4, 200 for TrES-3 and TrES-2, 400 for WASP-3, and 500 for HAT-P-7). We assume a boxcar-shaped transit at the depth of (RP /R⋆)2. The vertical lines show the positions of the 3:2 and 2:1 resonances with the known planet, and the cross-hatching shows the location of orbits which are not guaranteed to be stable by Hill's criterion per Gladman (1993). -- 20 -- Fig. 5. -- The best candidate transits for the five EPOCh targets. Each of the individual candidate transit events comprising the signal are shown at left; the phased and binned signal is shown at right. A time of zero on each X axis corresponds to the time of the first transit of the known planet observed by EPOCh. The ∆χ2 significance of the HAT-P-4, TrES-3, WASP-3, and HAT-P- 7 candidate signals fall below the detection criteria. While the significance of the TrES-2 candidate is above the detection criteria, one of the candidate transits (shown in the leftmost panel) occurs in a sparsely sampled part of the CCD, so the observations are uncalibrated and unreliable. Excising this candidate event, the significance of the remaining signal falls below the detection threshold.
1310.3851
1
1310
2013-10-14T20:54:17
Sustainability and the Astrobiological Perspective: Framing Human Futures in a Planetary Context
[ "astro-ph.EP" ]
We explore how questions related to developing a sustainable human civilization can be cast in terms of astrobiology. In particular we show how ongoing astrobiological studies of the coupled relationship between life, planets and their co-evolution can inform new perspectives and direct new studies in sustainability science. Using the Drake Equation as a vehicle to explore the gamut of astrobiology, we focus on its most import factor for sustainability: the mean lifetime <L> of an ensemble of Species with Energy-Intensive Technology (SWEIT). We then cast the problem into the language of dynamical system theory and introduce the concept of a trajectory bundle for SWEIT evolution and discuss how astrobiological results usefully inform the creation of dynamical equations, their constraints and initial conditions. Three specific examples of how astrobiological considerations can be folded into discussions of sustainability are discussed: (1) concepts of planetary habitability, (2) mass extinctions and their possible relation to the current, so-called Anthropocene epoch, and (3) today's changes in atmospheric chemisty (and the climate change it entails) in the context of pervious epochs of biosphere-driven atmospheric and climate alteration (i.e. the Great Oxidation Event).
astro-ph.EP
astro-ph
  Sustainability  and  the  Astrobiological   Perspective:   Framing  Human  Futures  in  a  Planetary  Context   Adam  Frank1  &  Woodruff  Sullivan2          We  explore  how  questions  related  to  developing  a  sustainable  human  civilization  can  be   Abstract   cast  in  terms  of  astrobiology.    In  particular  we  show  how  ongoing  astrobiological  studies  of   the  coupled  relationship  between  life,  planets  and  their  co-­‐evolution  can  inform  new   perspectives  and  direct  new  studies  in  sustainability  science.    Using  the  Drake  Equation  as   a  vehicle  to  explore  the  gamut  of  astrobiology,  we  focus  on  its  most  import  factor  for   sustainability:  the  mean  lifetime  𝐿  of  an  ensemble  of  Species  with  Energy-­‐Intensive   Technology  (SWEIT).    We  then  cast  the  problem  into  the  language  of  dynamical  system   theory  and  introduce  the  concept  of  a  trajectory  bundle  for  SWEIT  evolution  and  discuss   how  astrobiological  results  usefully  inform  the  creation  of  dynamical  equations,  their   constraints  and  initial  conditions.    Three  specific  examples  of  how  astrobiological   considerations  can  be  folded  into  discussions  of  sustainability  are  discussed:  (1)  concepts   of  planetary  habitability,  (2)  mass  extinctions  and  their  possible  relation  to  the  current,  so-­‐ called  Anthropocene  epoch,  and  (3)  today's  changes  in  atmospheric  chemisty  (and  the   climate  change  it  entails)  in  the  context  of  pervious  epochs  of  biosphere-­‐driven   atmospheric  and  climate  alteration  (i.e.  the  Great  Oxidation  Event).                                                                                                                     1  Department  of  Physics  and  Astronomy,  University  of  Rochester,  Rochester  NY.   [email protected]    2  Dept.  of  Astronomy  and  Astrobiology  Program,  University  of  Washington,  Seattle,   WA.  [email protected]     1        Anthropogenic  global  climate  change  is  currently  recognized  as  a  significant,  perhaps   1.    Introduction   fundamental,  issue  facing  human  civilization  (Solomon  et  al  2007).    The  chemical   composition  of  the  Earth’s  atmosphere  has  been  significantly  altered  by  human  activity.     Moreover,  detailed  analysis  of  global  data  sets  has  implied  the  potential  for  driving  the   climate  system  into  a  state  quite  different  from  the  one  in  which  human  civilization  has   emerged  and  flourished  (Parry  et  al  2007).    Recognition  of  the  likelihood  of  profound   climate  change  has  thus  led  to  the  desire  to  sustain,  to  some  degree,  the  current  climate   state.    Climate  science  is,  however,  just  one  domain  in  which  discussions  of  “sustainability”   have  emerged.    It  has  also  gradually  become  apparent  that  human  activity  has  been  driving   many  other  changes  in  the  coupled  "earth  systems”  of  atmosphere,  hydrosphere,   cryosphere,  geosphere  and  biosphere  in  ways  that  could  threaten,  or  at  least  strongly   stress,  the  so-­‐called  “project  of  civilization.”    Such  changes  to  the  earth  systems  include:  (a)  the  depletion  of  natural  fisheries  where  it  is   estimated  that  95%  of  all  fish  stocks  have  suffered  some  form  of  collapse  over  the  last  half   century  (Worm  et  al  2006);  (b)  diminishing  supplies  of  fresh  water  (Gleick  2003);  (c)  loss   of  rain  forest  habitat  (Williams  2003);  and  (d)  continuing  acidification  of  the  oceans   (Cicerone  et  al  2004).    In  all  cases  human  activity,  integrated  over  time  and  location,  have   led  to  substantial  changes  in  the  state  of  the  coupled  earth  systems.    These  changes  have   been  dramatic  enough  for  some  researchers  to  begin  speaking  of  the  beginning  of  the   Anthropocene,  a  new,  geological  epoch  succeeding  the  Holocene  (the  current  interglacial   period;  Zalasiewicz  et  al  2010).        The  field  of  sustainability  science  has  emerged  in  the  wake  of  this  recognition  seeking  to   understand  the  interactions  “between  natural  and  social  systems”  (Kates  2011).    In   particular  this  discipline  studies  how  such  interactions  can  lead  to  new  modalities  of   human  development  that  meet  “the  needs  of  the  present  and  future  generations.”     Sustainability  science,  bridging  disparate  domains  such  as  sociology  and  Earth  Systems     2   Science,  has  grown  rapidly.    More  than  20,000  articles  have  appeared  addressing   sustainability  science  over  the  last  40  years  with  a  doubling  in  the  number  of  articles  every   8  years  (Kates  2011).    Although  sustainability  science  often  focuses  on  place-­‐specific  issues,   by  its  very  nature  it  requires  a  global  perspective  to  address  issues  associated  with  the   need  to  “conserve  the  planet's  life  support  systems”  for  future  generations  (Kates  2010).       This  trend  is  clear  in  recent  studies  exploring  planetary-­‐scale  tipping  points  (Lenton  &   Williams  2013)  or  the  existence  of  planetary-­‐scale  “boundares”  as  safe  operating  limits  for   civilization  (Rockstrom  et  al  2009).    In  this  way  sustainability  science  and  the  theoretical   perspective  it  takes  on  the  trajectory  of  human  culture  is  necessarily  global,  or  better  yet   planetary.    It  is  from  that  perspective  that  sustainability  science  overlaps  with  the  domain   of  another  young  and  rapidly  growing  field:  astrobiology.        Astrobiology  is  essentially  the  study  of  life  in  an  astronomical  context  (Sullivan  &  Baross   2007).    The  NASA  Astrobiology  Institute,  for  example,  defines  its  subject  as  “the  study  of   the  origin,  evolution,  distribution,  and  future  of  life  in  the  universe”.  More  specifically,  since   most  extrapolations  involve  a  planetary  context  for  the  origin  and  evolution  of  life,   astrobiology  is  concerned  with  planetary  issues  just  like  sustainability  science.    Astrobiology  faces  an  obvious  "N=1  dilemma"  in  that  we  have  only  one  known  example  of   life  in  the  universe  (more  on  this  in  Section  2).    Nevertheless,  since  the  1990s  there  has   been  an  explosion  in  new  studies  and  new  results  relevant  to  astrobiology’s  core  questions.     For  the  purposes  of  this  paper  we  break  these  advances  into  three  research  domains   (though  there  are  others  that  are  of  broader  import):       1. Exoplanets:  the  discovery  of  planets  orbiting  other  stars  and  the  characterization  of   exoplanetary  systems  in  terms  of  habitability  for  life  (Lineweaver  &  Chopra  2012,   Udry  2007).   2. Solar  System  Studies:  the  detailed  (often  in  situ)  exploration  of  planets,  moons  and   other  bodies  in  our  own  planetary  system  with  a  focus  on  the  evolution  and  history   of  habitable  locations  (i.e.,  liquid  water,  sources  of  free  energy  for  metabolism,  etc.)   (Lineweaver  &  Chopra  2012,  Arndt  &  Nisbet  2012).   3     3. Earth  System  Studies:  the  detailed  investigation  of  the  Earth’s  history  including  the   history  of  the  3.5-­‐Gyr-­‐old  biosphere  and  its  coupled  interactions  with  atmosphere,   oceans,  ice  regions  and  land  masses  (Azua-­‐Bustos  et  al  2011,  Coustenis  &  Blac  et  al   2011).    The  most  notable  discovery  in  astrobiology  relevant  to  exoplanets  (domain  1)  has  been  the   recognition  that  planets  are  quite  common  in  the  Galaxy,  with  more  than  a  billion  Earth-­‐ mass-­‐like  worlds  expected  to  exist  on  orbits  within  the  habitable  zones  of  their  stars   (Seager  2012).    Relevant  to  solar  system  studies  (domain  2)  we  now  recognize  that  Mars   once  hosted  liquid  water  on  its  surface  and  that  many  of  the  moons  of  the  gas  and  ice  giant   planets  harbor  subsurface  liquid  oceans  (Castillo-­‐Rodgez  &  Lunine  2012).    Relevant  to   earth  systems  science  (domain  3),  we  realize  that  the  biosphere  and  non-­‐biological  Earth   systems  have,  at  least  during  some  epochs,  co-­‐evolved,  meaning  that  significant  feedbacks   have  led  to  substantial  changes  in  the  evolution  of  the  entire  earth  system.    The   development  of  an  oxygen-­‐rich  atmosphere  due,  in  part,  to  the  respiration  of  anaerobic   bacteria  is  one  example  of  such  co-­‐evolution  (Kasting  &  Knoll  2012).      Thus  astrobiology  takes  an  inherently  large-­‐scale  and  long-­‐term  view  of  the  evolution  of   life  and  planets.    In  this  way  the  data,  the  perspective  and  conceptual  tools  of  astrobiology   may  cast  the  global  problems  of  sustainability  science  into  a  different  and,  perhaps,  useful   light.    In  particular,  the  astrobiological  perspective  allows  the  opportunities  and  crises   occurring  along  the  trajectory  of  human  culture  to  be  seen  more  broadly  as,  perhaps,   critical  junctures  facing  any  species  whose  activity  reaches  significant  level  of  feedback  on   its  host  planet  (whether  Earth  or  another  planet).    In  this  way,  the  very  question  of   sustainability  may  be  seen  not  solely  through  the  lens  of  politics  and  policy  decisions   (Kates  2010),  but  also  as  an  essential  evolutionary  transformation  that  all  (or  at  least   many)  technological  species  must  experience.        In  this  paper  we  explore  the  argument  that  perspectives  developed  through  astrobiological   studies  can  usefully  inform  sustainability  science  by  broadening  its  understanding,   providing  case  studies,  and  suggesting  different  modes  of  conceptualization.  In  particular,   we  seek  to  frame  a  research  program  that  might  allow  researchers  to  develop  a  better     4   understanding  of  the  types  of  trajectories  a  biosphere  might  follow  once  a  generic  Species   with  Energy-­‐Intensive  Technology  (SWEIT)  emerges.      We  begin  in  Section  2  with  a  discussion  of  the  relevance  of  astrobiology,  using  the  standard   Drake  Equation  as  a  vehicle  for  framing  our  questions.  In  addition  we  use  the  Drake   equation  to  address  the  concept  of  a  statistically  relevant  ensemble  SWEITs.    We  then   discuss  in  Section  3  possible  theoretical  tools  for  modeling  sustainability  from  an   astrobiological  perspective  with  an  emphasis  on  dynamical  system  theory.    In  Section  4  we   present  three  examples  of  specific  astrobiological  topics  that  can  inform  sustainability:   definitions  of  habitability  across  time,  the  occurrence  of  mass  extinctions,  and  the   biosphere  driven  climate  change  as  a  consequence  of  SWEIT  activity.    Finally,  in  Section  5   we  summarize  our  findings  as  well  and  discuss  possible  directions  for  future  research.      Historically  the  Drake  Equation  has  been  instrumental  in  framing  discussions  of     2.    The  Drake  Equation  and  its  Longevity  Factor  L   astrobiology.    Originally  proposed  by  Frank  Drake  in  1962  as  a  means  for  estimating  the   present  number  (N)  of  radio-­‐transmitting  cultures,  the  equation  cleanly  parses  the   question  of  life  and  its  evolution  into  astronomical,  biological  and  sociological  factors   (Tarter  2007).    Note  that  the  original  intention  of  the  Drake  equation  was  to  estimate  the   number  of  civilizations  detectable  today.    Since  we  are  interested  in  the  question  of  SWEIT   lifetimes,  detectability  is  not  our  concern.    Instead  our  first  goal  is  to  use  equation  to   estimate  the  number  of  SWEITs  that  exist  now  or  have  already  gone  extinct.      In  its  traditional  form  the  Drake  Equation  is       (1)           𝑁 = 𝑅∗𝑓!𝑛! 𝑓! 𝑓! 𝑓! 𝐿    ,         5   where  𝑅∗ represents  the  rate  of  star  formation  in  the  Galaxy,  𝑓!  is  the  fraction  of  stars  that   host  planets,  𝑛!   is  the  mean  number  of  planets  in  the  so-­‐called  habitable  zone3  of  those   stars  with  planets,  𝑓!  is  the  fraction  of  those  planets  where  life  forms,  𝑓!  is  the  fraction  of   life-­‐bearing  worlds  that  evolve  intelligence,  𝑓!    is  the  fraction  of  intelligent  species  that   develop  radio  transmissions,  and  L    is  the  mean  lifetime  of  such  a  transmitting   technological  species.    In  this  paper  we  broaden  the  usual  definitions  of    𝑓!  and  L  beyond   solely  radio  transmission  to  consideration  of  the  emergence  and  longevity  of  any  SWEIT,   whether  or  not  radio  technology  is  involved.    Many  analyses  have  been  made  attempting  to  produce  estimates  of  N  relevant  to  radio-­‐ based  searches  for  other  technological  civilizations  (Wallenhorst  1981,  Pena-­‐Cabrera  &   Durand-­‐Manterola  2004).    Although  early  work  on  this  problem  constituted  a  kind  of   educated  guess  work,  those  efforts  were  nevertheless  useful  in  helping  to  structure  debate   about  factors  leading  to  the  emergence  of  technological  species.  When  Drake  first  proposed   the  equation  only  the  first  term  𝑅∗  could  be  estimated  at  all  (the  current  best  value  is   𝑅∗   ≈ 50  stars/yr    Prantozs  2013,  Watson,  private  communication).    In  the  last  two  decades,   however,  two  more  terms  in  the  Drake  Equation  have  become  well-­‐characterized.     Beginning  with  the  discovery  in  1995  of  an  exoplanet  orbiting  the  star  51  Peg,  astronomers   have  now  discovered  roughly  1000  planets  orbiting  other  stars  (Howard  2013,  Seager   2013).    Most  importantly,  the  current  sample  of  exoplanets  now  allows  a  good  estimate  of   the  fraction  of  planet-­‐bearing  stars  with  estimates  trending  towards  𝑓!~  1  (Seager  2013).     In  addition,  studies  of  transiting  planets,  especially  by  NASA's  Kepler  mission,  have  also   provided  constraints  on  the  number  of  exoplanetary  systems  (stars  with  multiple  planets),   their  architecture,  and  the  nature  of  the  discovered  planets.  From  this  work  the  number  of   Earth  sized  worlds  per  system  is  estimated  to  be  14%    (Dressing  &  Charbonneau  2013,   Howard  et  al  2013).  Grouping  all  terms  dependent  purely  on  astronomical  investigations,   i.e.,    𝑁! = 𝑅∗𝑓!𝑛!  ,  one  estimates  𝑁! = 7  habitable  planets  formed  per  year  ,  which  implies   that  there  now  exist  ~1010  potentially  habitable  planets  in  the  Galaxy.                                                                                                                     3  In  this  paper  we  use  the  traditional  definition  of  habitable  zone:  the  range  of  orbital   distances  from  its  host  star  in  which  a  planet's  surface  could  have  stable  liquid   water  (taken  as  the  sine  qua  non  for  life).     6    The  next  term  in  the  Drake  equation,  𝑓!  (the  fraction  of  habitable  worlds  in  which  life   actually  arises),  remains  completely  unconstrained.    There  is,  however,  an  expectation  that   over  the  coming  decades  estimates  of  its  value  may  be  possible.    Studies  of  Mars  in   particular  have  yielded  evidence  for  an  early  epoch  in  which  water  flowed  on  its  surface.     Future  in  situ  explorations  of  Mars  may  provide  evidence  of  fossil  (or  the  lack  thereof)  for   active  life  on  the  planet.    In  either  case  this  will  provide  constraints  on  𝑓! .    In  addition,   spectroscopic  studies  of  exoplanet  atmospheres  also  hold  the  possibility  of  yielding   evidence  for  “biosignatures”  in  the  form  of  non-­‐equilibrium  concentrations  of  atmospheric   constituents  that  can  be  linked  to  an  active  biosphere  (Seager  2013).    The  possibility  of  empirically  deriving  constraints  on  Drake  Equation  factors  ends,  however,   with  the  next  two  terms,  𝑓!  and  𝑓! .    Both  the  fraction  of  life-­‐bearing  planets  that  evolve   intelligence  and  the  fraction  of  those  that  evolve  technological  cultures  involve  questions  of   evolutionary  biology  and  sociology  that  are  unlikely  to  be  constrained  without  either  in  situ   explorations  of  exoplanets  (unlikely  for  centuries)  or  direct  contact  with  a  SWEIT.    Even   here,  however,  our  knowledge  of  Earth’s  own  evolutionary  history  allows  some  forms  of   inference  to  be  attempted  using  probability  theory  (Lineweaver  &  Davis  2002,  Spiegel  &   Turner  2012).    Recognition  that  the  evolutionary  trajectory  of  life  on  our  planet  has  passed   through  a  number  of  critical  steps,  each  of  low  probability,  has  allowed  some  authors  such   as  Carter  (1983,  Watson  2008)  to  argue  that  probability  distributions  for  𝑓!  and  𝑓! ,  and   hence,  broad  limits  on  their  values,  can  be  derived.        Consider,  for  example,  that  astronomical  and  geophysical  factors  give  a  planet  a  habitable   lifetime  of  𝑡! .  As  shown  in  Carter  (1983),  if  there  are  n  critical  steps  leading  to  some   property  such  as  intelligence  or  technological  capability  and  each  kth  step  has  low   probability  𝜆!  (𝜆! 𝑡! ≪ 1),  then  the  expectation  for  the  time  at  which  the  kth  step  will  occur   is         (2)             𝑡!/! = !!!! 𝑡!      .         7   Using  a  more  detailed  model  of  the  Earth’s  evolutionary  history,  Watson  (2008)  enlarged   Carter’s  argument  to  estimate  that  there  have  been  n  =  7  critical  steps  for  the  emergence  of   intelligence.    This  implies  that  intelligence,  on  average,  appears  only  at  the  very  end  of  a   planet's  era  of  habitability.    For  stars  like  the  sun  this  conclusion  leads  to  a  low  value  of   both  𝑓!  and  𝑓! .        The  discussion  in  these  papers  demonstrates  the  ways  in  which  considerations  of  Earth's   evolutionary  history  allow  for  broad  astrobiological  reasoning.  For  our  purposes  -­‐   exploring  the  relevance  of  astrobiology  to  key  issues  in  sustainability  -­‐  the  debate  over  the   correct  value  of  the  still  unknown  terms  in  the  Drake  Equation  is  less  important  than  the   existence  of  data  and  methods  that  advance  that  debate.    In  particular  there  is  now  a   significant  body  of  empirically  derived  knowledge  about  the  planetary  context  of  life  either   as  it  exists  on  Earth  or  its  potential  on  other  worlds.  This  may  make  it  possible  to  address   the  single  most  important  aspect  of  the  Drake  equation  for  sustainability  science:  L,  the   final  factor.    We  therefore  argue  that  a  key  question  in  sustainability  science  can  be  stated   in  explicitly  astrobiological  terms:       average  lifetime  𝐿     Given  an  ensemble  of  N  species  with  energy  intensive  technology  (SWEITs),  what  is  their    This  is  equivalent  to  asking:  if  we  could  rerun  Earth’s  past  (and  future)  history  many  times   and  select  those  trajectories  leading  to  SWEIT  (note  they  need  not  be  human),  then  what   value  of  𝐿  would  be  obtained  across  that  ensemble  of  histories?  While  we  do  not  know  the   actual  value  of    𝐿,  using  the  Drake  Equation  we  can  point  to  a  methodology  and  start  to   answer  this  question.      First,  we  should  check  whether  our  proposed  ensemble  of  SWEIT  makes  sense.  We  can   calculate  how  large  a  volume  of  space  is  needed  to  contain  a  large  enough  SWEIT  sample   size  K  such  that  averages  like  𝐿  become  meaningful.    For  K, we  consider  a  value  of  1000  to   constitute  a  statistically  relevant  sample  since  the  deviations  around  averages  will,  in  a   binomial  distribution,  go  as  1 √𝐾  ,or  just  3%.    Since  we  are  interested  in  the  time-­‐ integrated  number  of  SWEITs  (meaning  we  include  SWEITs  already  extinct),  we  can  use  an     8   alternative  form  of  the  Drake  Equation,  considering  it  to  be  a  probability  distribution   𝑑𝑁!  of  the  number  of  SWEITs  in  a  volume  associated  with  a  relevant  cosmic  scale  (a  galaxy,   a  cluster  of  galaxies,  or  observable  Universe)  relative  to  the  lifetime  𝑇!  of  structures  at  that   scale.    Thus  we  write  the  Drake  Equation  as,       𝑑𝑁! = 𝑁!"𝐹!"  (𝑑𝐿/𝑇! )  ,             (3)    where  𝑁!"  is  the  total  number  of  habitable  planets  in  the  volume  associated  with  the  scale   considered  (  𝑁!" = 𝑁∗! 𝑓!𝑛!  ),  and  𝐹!" = 𝑓! 𝑓! 𝑓!  is  the  combination  of  biological  and   technological  evolution  factors  which  we  call  the  “bio-­‐technological  probability”.    The  total   number  of  SWEITs  (currently  alive  or  extinct)  at  the  chosen  scale  is  the  integral  over  all   SWEIT  lifetimes,  or  𝑁! = 𝑁!"𝐹!"  .    Assuming  a  constant  density  of  stars  (at  whatever  scale  we  are  interested  in),  we  can  use   this  alternative  form  to  derive  a  simple  expression  for  what  we  call  the  enclosure  radius  𝑅!  ,   which  measures  the  size    of  a  region  containing  a  statistically  relevant  sample  of   technological  species.  Considering  𝑛! = 𝑁! /(!! 𝜋𝑅!! )  to  be  the  number  density  of  SWEIT  over   the  given  scale  𝑅!  we  have  !! 𝜋𝑅!!𝑛! =  𝐾  or,       𝑅! =   𝑅! !!!"!!" !/!             (4)    For  for  galactic  scales  (𝑁!" ≡ 𝑁!" )  we  assume    𝑓! = 0.5, 𝑛! = 0.5,  𝑁∗! = 10!! ,  and  𝑅! ≅ 10!  ly  .    This  yields  𝑁!" ≅ 10!"  and  𝑅!" =  220  lt_yr   𝐹!" !!/! .    For  the  scale  of  the   observable  Universe  we  assume  the  same  values  of  𝑓!  𝑎𝑛𝑑  𝑛! ,  𝑅! ≅ 10!"  ly  and  𝑁∗! = 10!" .     This  yields  𝑅! =  2154  lt_yr   𝐹!" !!/! .    In  Fig  1  we  plot  the  enclosure  radius  as  a  function  of   𝐹!"  .         Using  these  expressions  we  ask  what  is  the  minimum  value  of  the  bio-­‐technological   probability  𝐹!" ∗  that  yields  a  statistically  revelant  sample  in  the  local  Univese  𝑅!" .    Taking   𝑅!" = 0.01  𝑅!  we  have  𝐹!" ∗ = 10!!" .    Thus  even  with  the  odds  of  evolving  a  SWEIT  on  a     9   given  habitable  planet  being  one  in  one  million  billion  at  least  1000  species  will  still  have   passed  through  the  transition  humanity  faces  today  within  our  local  region  of  the  cosmos.   Of  course  for  less  restrictive  values  of  𝐹!"  a  representative  sample  of  species  can  be  found   within  the  local  neighborhood  of  our  Galaxy.    For  example  assuming  𝐹!" = 10!!    yields   𝑅! = 3  ×  10!  ly.       This  analysis  demonstrates  that  from  an  astrobiological  perspective  the  concept  of  a   representative  SWEIT  ensample  is  reasonable  even  if  the  evolution  of  these  species  is   highly  unlikely.    Over  the  course  of  cosmic  evolution  enough  of  SWEITs  should  have  arisen   to  allow  one,  in  principle,  to  meaningfully  inquire  about  average  properties  of  their   developmental  trajectories  (such  as  𝐿).    Note  that  we  are  explicitly  not  asking  if  any  of  these   SWEIT  could  be  contacted.    They  may  already  be  long  extinct.    The  exercise  here  is  simply   to  determine  if  such  an  ensemble  is  meaningful  to  consider  within  a  “local”  volume  of  the   Universe.     statistically  significant  (𝑲  =  1000)  SWEIT  enseamble  versus  𝑭𝒃𝒕 ,  the  combined  biological  and   Figure  1.    Enclosure  radius  vs  bio-­‐technological  probability.    Radius  of  sphere  Re  enclosing  a   evolutionary  factors  in  the  Drake  equation.    Blue  line  is  appropriate  for  intergalactic  distances  (mean   densities  of  stars  smoothed  over  many  galaxies)  and  is  valid  for  distances  above  Re  >106  light-­‐years.       Red  line  is  appropriate  for  densities  of  stars  within  a  galaxy  and  is  valid  for  distances  Re  <106  light-­‐ years.     10   Given  the  plausibility  of  the  existence  of  a  SWEIT  ensemble,  we  now  turn  to  exploration  of   how  their  trajectories  of  development  can  be  conceptualized.   We  now  consider  the  general  framework  in  which  average  “trajectories”  for  SWEIT   evolution  might  be  explored.    We  use  the  formalism  of  dynamical  systems  theory  (Beltrami   3.  Trajectories  of  Technological  Energy-­‐Intensive  Species   1987)  in  which  a  set  of  governing  differential  equations  is  invoked  to  represent  the   evolution  of  a  single  SWEIT.        These  equations,  with  appropriate  initial  conditions  and  constraints,  determine  the   trajectory  of  the  system  through  the  multi-­‐dimensional  solution  space  defined  by  the   system’s  independent  variables.    Invoking  the  concept  of  an  ensemble  of  systems,  each  with   different  initial  conditions  and  constraints,  allows  us  to  explore  suitably  averaged   evolutionary  characteristics  such  as  a  mean  lifetime  𝐿.    The  application  of  dynamical   systems  theory  in  this  context  is  similar  to  that  used  in  studies  of  systems  ecology  (May   1977,  Petraitis  &  Hoffman  2010)  and  ecological  economics  (Krutilla  &  Reuveny  2006).    The  SWEIT  solution  space  is  likely  to  be  hyper-­‐dimensional  depending  on  conditions   inherent  to  the  planet  on  which  the  species  evolved  (ocean  coverage,  existence  of  plate   tectonics,  availability  of  various  classes  of  resources  such  as  fossil  fuels,  etc.)  as  well  as   biological  and  social  characteristics  (lifespans  of  individual  organisms,  degree  of  social   cooperation,  etc.).    For  our  purposes  in  articulating  a  broad  program  of  research,  we   consider  a  simplified  set  of  equations  and  a  concomitant  solution  space  that  may  capture   essential  characteristics  of  the  problem    As  a  toy  model  to  illustrate  the  proposed  method,  consider  an  intelligent  species  -­‐  on  the   path  to  developing  energy  intensive  technology  -­‐  that  can  harvest  some  form  of  biomass  for   energy  (such  as  trees  in  Earth’s  example).    Thus  the  energy  resource  can  be  expressed  as  a   "population”  𝐸 .    Let  us  assume  that  this  renewable  resource’s  own  growth  is  limited  by   environmental  factors  which  lead  to  a  carrying  capacity  𝐾 .    Thus  the  population  of  the   energy  bearing  resource  E  is  finite.  The  coupling  between  the  growth  of  the  SWEIT     11   population  𝑁  and  the  energy  resource  population  𝐸  can  then  be  described  by  a  modified   (logistic)  form  of  a  simple  Predator-­‐Prey  system  (Brauer  &  Castillo-­‐Chavez  2011)              !"!" = 𝑏𝐸 1 − !! − 𝑎𝐸𝑁             (5)    (6)                          !"!" = 𝑐𝑎𝐸𝑁 − 𝑑𝑁    where  𝑏  is  the  growth  rate  of  the  energy  resource,    𝑎  is  the  SWEIT  “predation  rate”  of   energy  (its  rate  of  energy  harvesting),    𝑐  is  the  rate  at  which  the  resource  can  be  used  to   increase  the  SWEIT  population,  and  𝑑  is  the  SWEIT  mortality  rate.        The  behavior  of  such  as  system  in  its  2-­‐D  solution  space  (𝑁 , 𝐸)    is  presents  a  textbook   example  of  a  stable  dynamical  system  in  which  an  initial  population  𝑁!  and  resource  𝐸!   experiences  oscillatations  with  decreasing  in  amplitude  until  a  steady  state  is  achieved  (at   𝑁! = 𝑏 𝑎 (1 − 𝑑 𝑐𝑎𝐾)  and  𝐸! = 𝑑 𝑐𝑎).    Figure  2  shows  a  representative  solution  to  the   system  given  by  equations  5  and  6.       SWEIT  and  Energy  resource  (𝑵𝒐 , 𝑬𝒐 )  and  evolves  to  a  stable  solution  as  demonstrated  by  its   Figure  2.  Trajectory  for  coupled  SWEIT/energy-­‐resource  populations  described  by  modified   trajectory  approaching  a  fixed  point  (𝑵𝒔 , 𝑬𝒔 ).       predator-­‐prey  system  (equations  5  &  6).    The  systems  begins  with  low  populations  of  both       12   The  system  of  equations  5  and  6  represent  a  simple  example  of  the  idea  of  a  solution  space.     In  reality  we  would  need  a  much  more  general  description  of  SWEIT  evolution.      The  lowest   dimensional  solution  space  appropriate  for  the  evolution  of  a  SWEIT  would  however  have   to  include  some  form  of  feedback  on  the  planetary  system.      Thus  we  need  at  least  one   independent  variable  that  represents  the  “forcing”  of  the  planetary  systems  by  SWEIT   activity.      This  forcing  would  likely  represent,  in  some  form,  the  energy  released  into  the   coupled  planetary  systems  due  to  the  technological  use  of  harvested  energy.    Thus  we   define    𝑟! = 𝐷! 𝐷!  to  be  the  ratio  of  the  planetary  system  forcing  driven  by  SWEIT  growth   (𝐷! )  to  that  produced  without  the  technological  activity  of  the  species  (i.e.  the  planets  own   “natural”  forcings,  𝐷! ).    As  an  example  consider  that  at  the  present  epoch  of  human   evolution  the  rate  of  energy  trapping  by  anthropogenic  CO2  emission  represents  a  form   of  𝐷! ,  while  the  rate  of  energy  trapped  as  a  result  of  volcanic  outgassing  of  CO2  represents  a   component  of  𝐷! .    Thus  we  suggest  that  a  minimum  solution  space  vector  𝑆  for  an  individual  SWEIT  is   𝑆 = (𝑁 , 𝑒! , 𝑟! )  where  𝑁  is  the  SWEIT  population,  𝑒!  is  the  energy  harvested  per  captia,   and  𝑟!  represents  SWEIT  feedback  on  planetary  systems.    Note  that  the  variable  𝑁  captures   the  inherent  success  of  the  species  defined  from  a  purely  Darwinian  perspective;  the   variable  𝑒!  captures  the  success  of  the  species’  technological  capacities  as  it  harvests  and   utilizes  more  energy  than  would  be  possible  without  technology;  the  variable  𝑟!  captures   reality  of  thermodynamic  feedback  such  that  energy  used  for  work  demands  entropy   generation.    In  particular,  if  𝑟! ≫ 1  then  the  combined  planetary  systems  will  be  driven  into   new  states  on  timescales  shorter  than  bio-­‐evolutionary  time  scales  for  the  SWEIT.           13     Figure  3.    Solution  space  for  SWEIT  evolution.    Shown  is  a  schematic  of  possible  stability  domains  in  a   single  N  (population),  ec  (energy  harvest  rate  per  capita)  plane.    As  shown  schematically  in  figure,    A  specific  SWEIT  evolutionary  model  takes  the  general  form  !"!" = 𝑀 𝑁 , 𝑒! , 𝑟! , …  where  the   region  III  is  most  likely  to  drive  unstable  increases  in  the  direction  of  increased  planetary  forcing,   rf  ,which  may  degrade  the  planet’s  ability  support  a  large  SWEIT  population.   vector  of  functions  𝑀 𝑁 , 𝑒! , 𝑟! , …  reflects  explicit  couplings  between  terms.    The  research   program  is  embodied  in  the  exploration  of  different  functional  forms  of  𝑀 𝑁 , 𝑒! , 𝑟! , … .     These  different  models  should  be  informed  by  insights  gained  from  astrobiological  models   (see  Sec.  4).    Those  insights  will  also  be  expressed  through  the  choice  of  initial  conditions   for  the  systems  and  additional  constraints  on  variables  or  coupling  constants  between   variables.        For  a  given  specific  model  (i.e.  a  choice  of  𝑀 𝑁 , 𝑒! , 𝑟! , … )  dynamical  systems  theory  then   allows  analysis  of  regions  of  instability  as  well  the  presence  of  quasi-­‐stable  limit  cycles  or   stable  fixed  points  similar  to  what  occurs  in  the  simple  Predator-­‐Prey  model  presented   above,  (i.e.  the  stable  fixed  point  at  (𝑁! , 𝐸! )  in  Fig. 3).    For  a  more  complete  model  such   global  classes  of  behavior,  i.e  the  existence  of  limit  cycles,  attractor  basins,  saddle  points  etc,     14   should  allow  broad  classes  of  behavior  in  SWEIT  trajectories  to  be  classified  and   understood.    Considering  that  the  growth  of  the  population  N  is  likely  to  be  coupled  to  the   harvested  energy  (i.e.,  there  is  a  positive  feedback  between  the  two  variables)  one   should  be  able  to  map  regions  of  stability  and  instability  in  the  solution  space.     Figure  3  presents  a  schematic  of  possible  regions  of  stability/instability  in  (𝑁 , 𝑒! )   plane.    Region  I,  with  low  population  and  low  𝑒! ,  is  likely  stable  in  that  a  population   could  remain  in  that  region  for  long  periods  relative  to  timescales  inherent  to  the   environment.    Given  the  low  energy  harvesting  capabilities  this  region  would  not  be   a  SWEIT  but  something  like  human  civilization  many  millennia  ago.    Region  II,  with   high  population  and  low  𝑒!  ,  would  likely  be  unstable,  as  the  SWEIT  would  consume   the  resources  provided  by  the  environment  on  timescales  short  compared  with  their   natural  regeneration  time  and  would  lack  the  energy  harvest  capacities  to  enhance   the  productivity  of  the  environment.    Region  III,  corresponding  schematically  to   human  society's  current  location,  could  in  principle  be  stable  as  it  maintains  high   populations  through  a  large  capacity  to  harvest  energy.    The  difficulty  for  a  SWEIT  in   this  region,  however,  is  that  feedback  on  planetary  systems  generated  as  that  energy   is  used  will  inevitably  change  𝑟! .    This  would  mean  that  region  III  is  unstable  leading   to  movement  perpendicular  to  the  (𝑁 , 𝑒! )  plane  and  eventually  into  regions  where   the  environment  is  driven  into  new  states  on  time  scales  short  compared  with   natural  responses.    Finally,  Region  IV  appears  potentially  stable  (sustainable)  as  a   smaller  population  with  relatively  high  𝑒!  might  not  be  capable  of  maintain  stable   values  of  𝑟!  over  long  time  scales.    Note  that  details  of  the  locations  of  these   quadrants  will  vary  depending  on  the  details  of  the  system  which,  in  turn,  determine   the  limits  at  which  instabilities  set  in.    The  purpose  of  an  explicit  modeling  program   would  be  to  articulate  such  details  including  the  size  of  the  regions  (shown   schematically  to  be  of  equal  area  in  Fig  3).    Figure  4  shows  two  classes  of  behavior  in  the  full   𝑁 , 𝑒! , 𝑟!  solution  space.    The  red  line   represents  a  “collapse”  trajectory  whereby  the  species  population  initially  increases  along     15   with  its  technology  (energy  harvesting  capacities)  at  low  𝑟! .    Once  a  threshold  of  𝑁  and  𝑒!  is   reached,  however,  entropy  generation  pushes  the  solution  rapidly  towards  higher  𝑟!  and   global  instability,  thus  leading  to  systemic  failures  of  technological  systems  and  negative   feedbacks  from  the  planetary  systems  on  which  those  systems  depend.    Collapse  of  the   population  (low  values  of  N)  then  follows.    The  blue  line  represents  a  sustainabile   trajectory  in  which  smaller  levels  of  population  with  high  technology  are  achieved  before   high  values  of  𝑟!  negatively  impact  the  ability  of  the  species  to  maintain  its  energy   harvesting  systems.    We  represent  the  end  state  of  such  a  solution  as  a  limit  cycle  given  the   inherent  time-­‐dependence  of  such  sustainable  habitats  (see  Sec.  4.1).  We  expect  that  the   topology  of  the  solution  space  will  dictate  the  actual  form  of  these  solutions  in  that  both   collapse  and  sustainable  trajectories  will  be  defined  through  attractor  basins  and  unstable   terrains  such  as  saddle  points  between  them  (Brauer  &  Castillo-­‐Chavez  2011).     Figure  4.    Schematic  of  two  classes  of  trajectories  in  SWEIT  solution  space.    Red  line  shows  a   trajectory  representing  population  collapse  whereby  development  of  energy  harvesting   technologies  allows  for  rapid  population  growth  which  then  drives  increases  in  planetary   forcing.    As  planetary  support  systems  change  state  the  SWEIT  population  is  unable  to   maintain  its  own  internal  systems  and  collapses.    Blue  line  shows  a  trajectory  representing   sustainability  in  which  population  levels  and  energy  use  approach  levels  that  do  not  push   planetary  systems  into  unfavorable  states.           16   In  this  section  we  outline  three  specific  examples  where  astrobiological  studies  might  drive   new  research  aimed  at  developing  a  broader  understanding  of  sustainable  human  cultures.   4.    Areas  of  Astrobiology  Relevant  for  Sustainability  Science       A  central  concept  in  astrobiology  is  habitability,  defined  broadly  as  the  ability  of  a  planet  to   support  life  (Sullivan  &  Carney  2007);  note  that  a  planet  may  be  habitable  at  a  certain  time   4.1  Planets,  Moons  and  Exoplanets:  Habitability,  Sustainability  and  Time   even  if  no  life  actually  exists.    Habitability  may  also  be  restricted  to  include  those   conditions  in  which  abiogenesis  (the  formation  of  life  from  non-­‐life)  can  occur  (Lineweaver   &  Chopra  2012),  though  this  is  not  necessary  because  life  could  start  elsewhere  and  be   transported  to  a  planet  (panspermia)  by  a  number  of  means  (e.g.,  debris  from  planetary   impacts  occurring  elsewhere  in  a  planetary  system).        Note  that  “Habitability”  in  astrobiology  is  a  broader  concept  than  “sustainability”  in   sustainability  science.  Instead  of  astrobiology  asking  if  any  form  of  life  is  possible  on  an  any   given  planet,  sustainability  science  asks  if,  on  a  particular  planet  (Earth),  a  particular  kind  of   human  civilization  (ours)  is  possible  over  long  time  scales.    Thus  sustainability  becomes  a   subset  of  habitability,  albeit  the  subset  with  the  greatest  urgency  to  our  society  today.     Given  such  a  connection  it  is  sensible  to  understand  how  progress  in  astrobiological  studies   of  habitability  can  inform  planetary  perspectives  on  sustainability.    One  of  the  most  basic  definitions  of  habitability  corresponds  to  a  planet  having  surface   temperature    𝑇!  such  that  water  molecules  on  the  surface  would  be  in  a  liquid  state   273  K <   𝑇! < 373    K  for  a  pressure  of  1  atm .    For  a  planet  at  a  distance  D  from  a  star   with  temperature  𝑇!  and  radius  𝑅!  this  yields,       𝑇! = 1 − 𝑎 !/! !! !!! !/! 𝑇!             (7)    where  𝑎  measures  the  planetary  albedo  (reflectivity)  of  incident  starlight.    The  dependence   of  𝑇!  on  distance  means  one  can  define  a  band  of  orbits  (a  range  of  D)  called  the  Habitable   Zone  (HZ),  where  surface  temperatures  allow  water  to  exist  in  the  liquid  state  (Brownlee  &     17   Kress  2007).    Tp  (and  thus  the  width  and  location  of  the  HZ)  can  also  be  significantly  altered   by  the  radiative  properties  of  the  planet's  atmosphere,  e.g.,  Tp  is  higher  in  the  presence  of   greenhouse  gases.    The  discovery  of  subsurface  “oceans”  beneath  the  water-­‐ice  surfaced  moons  of  giant   planets  in  our  solar  system  (e.g.,  Europa,  Enceladus,  and  perhaps  Titan)  has  enlarged  our   understanding  and  definition  of  habitability.    Despite  their  huge  distance  from  the  sun,   these  moons  maintain  water  in  the  liquid  state  due  to  tidal  heating  by  their  host  planet   (Chyba  &  Phillips  2007,  Forgan  et  al  2013).    Studies  of  Earth’s  oceans  have  also  shown  that   purely  chemotrophic  ecosystems  can  develop  around  deep-­‐sea  hydrothermal  vents   without  any  need  for  energy  from  sunlight  (Baross  et  al  2007).  Thus  it  is  possible  that   subsurface  oceans  on  giant  planet  satellites  might  also  be  habitable  (McKay  2007).    The   discovery  of  exoplanets  in  a  variety  of  orbital  configurations  has  also  served  to  broaden   definitions  of  habitability  as  astronomers  consider  a  far  wider  range  of  orbital  properties   (e.g.,  rocky  planets  in  tight  orbits  around  cool  dwarf  stars)  than  those  provided  by  our  own   solar  system  (Howard  2013).    Thus  the  spatial  domains  of  habitability  have  been   significantly  broadened  through  recent  astrobiological  studies.      For  sustainability,  however,  a  far  more  important  domain  may  be  astrobiological  insights   into  the  temporal  domain  of  habitability.  The  capacity  for  a  planet  to  support  life  is,   fundamentally,  a  time-­‐dependent  property.    This  can  be  most  easily  seen  in  equation  7   where  each  factor  in  the  equation  can  be  expected  to  vary  (though  on  different  timescales).     For  example,  the  variation  of  the  sun’s  temperature  and  radius    𝑇! , 𝑅!  has  been  moving  its   habitable  zone  outward  over  billion-­‐year  timescales.    Based  on  solar  evolution  models  it  is   estimated  that  increases  in  the  sun's  luminosity  of  ~10%/Gyr  will  render  the  Earth   uninhabitable  by  most  forms  of  present  life  as  “soon”  as  ~1.0  Gyr  in  the  future  (Caldiera   and  Kasting  1991,  Watson  2008,  Kopparapu  et  al  2012)   .        The  study  of  Mars  provides  the  most  direct  and  important  example  of  time-­‐dependent     habitability.    Studies  by  a  variety  of  orbiters,  landers  and  rovers  over  the  last  two  decades   have  built  a  conclusive  case  that  Mars  once  hosted  a  warmer,  wetter  climate  (Azua-­‐Bustos,   et  al  2012).    In  particular,  early  conditions  on  Mars  may  have  allowed  abiogenesis  to  occur     18   (Jakosky  et  al  2007).    The  period  of  Martian  habitability  lasted  less  than  1  Gyr  and  ended   perhaps  ~4  Gyr  ago.    The  cause  of  the  loss  of  Martian  habitability  is  still  debated,  but  one   favored  theory  invokes  the  early  escape  of  its  atmosphere’s  lighter  elements  and  molecular   species  into  space  via  collisions  with  the  solar  wind.     changes  with  time.    The  same  principle  must  surely  apply  to  sustainability.  Note  that   sustainability  requires  conditions  appropriate  not  just  for  life  but  for  a  particular  kind  of   One  principal  lesson  to  be  drawn  from  astrobiological  studies  is  that  planetary  habitability   social  organization  or  a  particular  kind  of  species  (a  technological  energy-­‐intensive  one).     Thus  naively  one  may  expect  that  for  cases  like  our  own  in  which  the  appearance  of  strong   SWEIT  induced  planetary  forcing  is  just  beginning,  (𝑟! ≫ 1)  ,  sustainable  states  may  be   inherently  more  delicately  balanced  .    This  would  be  the  case  because  of  competing  forces   where  cultural  dynamics  are  just  as  important  as  the  response  of  the  planet’s  biological  and   physical  systems.    Thus  we  should  expect  that  sustainability  will  be  inherently  easier  to   unbalance  than  just  habitablity.    In  other  words  one  can  expect  the  existence  of  multiple   modes  of  instability  amongst  the  coupled  systems  facing  any  SWEIT  (Helbing  2012).      Modeling  trajectories  in  systems  coping  with  rapidly  changing  climate,  for  example,  might   draw  on  understandings  derived  from  the  history  of  Mars,  Venus  or  Earth’s  own  past,  such   as  during  the  aftermath  of  mass  extinctions  (Sec.  4.2).        A  more  detailed  understanding  of  habitability  has  already  changed  understandings  of   sustainability  in  terms  of  climate  states.    For  example,  studies  of  our  own  solar  system   (Kopparapu  et  al  2012)  show  that  the  Earth’s  orbit  is  now  located  just  beyond  the  inner   edge  of  the  solar  system's  HZ  (0.99  AU)  .    This  means  it  is  possible  that  our  present  climate   is  closer  to  the  limiting  case  of  a  so-­‐called  “moist  greenhouse”  in  which  water  vapor  (an   effective  greenhouse  gas)  from  ocean  evaporation  plays  a  significant  role  relative  to  CO2   and  other  greenhouse  gases.    Had  we  found  ourselves  in  the  center  of  the  Sun's  HZ,  the   possibility  of  such  a  moist  greenhouse  would  be  more  remote.  It  has  been  estimated  that  if   we  continue  current  anthropogenic  CO2  deposition  rates  into  the  atmosphere,  Earth  might   enter  the  moist  greenhouse  phase  by  as  early  as  the  year  2300.       19   To  summarize,  studies  of  sustainability  on  a  planetary  scale  can  be  seen  as  a  subset  of   astrobiological  concerns  with  habitability.    Consideration  of  astrobiological  studies  of   habitability  demonstrate  the  inherent  time  dependence  that  equally  applies  to   sustainability  issues.  Deeper  consideration  of  astrobiological  studies  concerning  the  time   dependence  of  habitability  may  thus  prove  also  useful  for  sustainability  studies.     Astrobiology  is  concerned  with  the  long-­‐term  evolution  of  life  on  any  planet  where  it  might   arise.  The  only  example  we  have  to  date  is,  of  course,  that  of  Earth  and  so  astrobiologists   4.2    Mass  Extinctions:  Constraints  on  Responses  to  the  Anthropocene   are  deeply  concerned  with  the  “major”  events  in  the  history  of  Earth's  life.    Particular  issues   are:  What  constitutes  a  major  event?  When  did  these  happen?  What  forces  drove  these   events  and  would  they  be  likely  or  inevitable  for  life  on  another  planet  (whether  exotic  life   or  our  own  form)?  One  particular  class  of  major  events  includes  mass  extinctions  (Ward   2007).    The  history  of  Earth’s  flora  and  fauna  over  the  past  550  Myr  provides  evidence  of  five   distinct  mass  extinctions.  These  were  events  in  which  the  rate  of  extinction  rose  above  the   rate  of  speciation  and  more  than  75%  of  all  species  were  removed  from  the  biosphere  over   a  relatively  short  duration  (Barnosky  et  al  2011).  The  “Big  Five”  mass  extictions  are:  End-­‐   Cretaceous  (KT)  Event,  65  Ma  (Myr  ago)  with  an  estimated  76%  of  all  species  lost;  Triassic   Event,  200  Ma,  80%  lost;  Permian  Event,  251  Ma,  96%  lost;  Devonian  Event,  359  Ma,  75%   lost;  and  Ordovician  Event,  443  Ma,  86%  lost.    In  these  events  significant  fractions  of  both   land  and  marine  species  were  rapidly  driven  into  extinction,  followed  in  each  case  by  a   significantly  greater  diversity  of  species.    Furthermore,  it  has  been  argued  that  human   activity  is  now  driving  the  biosphere  into  a  new,  sixth  mass  extinction,  the  Anthropocene   Event    (Barnosky  et  al  2011).    Thus  mass  extinctions  represent  another  potential  area  in   which  astrobiological  considerations  can  address  both  specific  and  foundational  questions   in  sustainability  science.        The  best  characterized  mass  extinction  is  the  End-­‐Cretaceous  (KT)  event  which  eliminated   the  dinosaurs  and  three-­‐quarters  of  all  other  species,  in  the  process  eventually  allowing   mammals  to  gain  a  dominant  foothold  (Ward  2007).    Most  lines  of  evidence  point  to  an     20   asteroid  or  comet  impact  as  the  primary  cause  of  this  mass  extinction.    The  KT  event  is,   however,  the  only  mass  extinction  that  can  definitely  be  associated  with  an  impact.        The  most  significant  mass  extinction  in  Earth’s  history  is  currently  ascribed  to  causes   internal  to  earth  systems  coupling.    The  End-­‐Permian  Mass  Extinction  (251  Ma),  involved   the  greatest  loss  of  biodiversity  in  Earth's  history.    Within  only  0.2  Myr,  more  than  56%  of   all  genera  and  96%  of  all  species  were  lost.    Current  research  links  the  cause  of  the  Permian   extinction  event  to  large-­‐scale  volcanic  magma  flows  associated  with  the  formation  of  the   Siberian  Traps  (Payne  and  Clapham  2012).    The  release  of  CO2  from  the  flows  triggered   enhanced  global  warming  which  then  initiated  a  strong  response  from  the  climate  system.     Changes  in  ocean  levels  of  CO2  (and  therefore  acidity)  and  subsequent  changes  in  ocean   circulation  led  to  deep  marine  anoxia.    The  alterations  in  ocean  circulation  also  led  to   changes  in  ocean  stratification.  Purple  algae  brought  into  contact  with  sunlight-­‐rich  layers   may  then  have  driven  production  of  high  levels  of  hydrogen  sulfide  both  at  sea  and  over   land,  further  enhancing  extinction  levels.        Thus  it  appears  that  climate  change,  driven  by  enhanced  release  of  greenhouse  gases,  was   the  agent  driving  the  most  powerful  mass  extinction  in  Earth’s  history.      In  fact,  other  than   the  KT  event,  it  may  well  be  that  climate  change  driven  by  increased  greenhouse  gas   concentrations  through  volcanism  played  a  major  role  in  all  mass  extinction  events  (Payne   and  Clapham  2012,  Fuelner  2008).  The  implication  of  rapidly  changing  climate  as  either  a   direct  or  secondary  cause  of  previous  mass  extinctions  holds  obvious  lessons  for  our  own   situation  (i.e.  environmental  stresses  driving  cascades  of  extiction,  Newman  1997).     Understanding  the  ways  climate  has  coupled  to  significant  changes  in  biodiversity  in  the   past  is  one  direct  application  of  astrobiology  to  sustainability  studies.    This  is  particularly   true  as  such  understanding  can  be  applied  to  current  conditions  in  which  anthropogenic   climate  change  and  other  anthropogenic  drivers  are  forcing  the  coupled  earth  systems.      The  record  of  previous  mass  extinctions  has  other  uses  for  sustainability  studies,  including,   for  example,  characterization  of  21st  century  trends  in  biodiversity.    Consideration  of  the   fossil  record,  for  example,  leads  to  estimates  that  the  time  to  a  75%  reduction  in  species   (i.e.,  an  "official"  sixth  mass  extinction)  will  be  just  200-­‐600  yr  from  now  (assuming  that     21   current  loss  rates  in  biodiversity  hold,  Barnosky  et  al  2011).    Such  extrapolations  into  the   future  remain  uncertain,  but  can  be  improved  by  looking  both  backward,  in  terms  of  better   accounts  of  previous  extinction  events,  and  forward,  in  terms  of  better  modeling  of  the   impact  of  the  Anthropocene  on  the  biosphere.    Such  modeling  can  be  done  within  the   general  context  of  SWEIT  trajectory  determination  in  order  to  set  limits  on  the  sensitivity   of  a  SWEIT  to  rapid  changes  in  biodiversity  .         Another  series  of  key  events  in  the  history  of  Earth's  life  involved  profound  changes  in  the   composition  of  the  atmosphere.  In  particular  the  concentration  of  oxygen  in  both  the  early   4.3    The  Great  Oxidation  Event:  Predicting  Climate  Change  For  SWEIT   Earth’s  atmosphere  and  oceans  was  far  lower,  with  atmospheric  values  probably  < 10!!  of   current  values  (Catling  &  Kasting  2007).    The  emergence  of  an  oxygen  rich  atmosphere  was   a  key  event  in  the  history  of  the  planet  (Kasting  &  Kirschvink  2012).    Furthermore,  it  was   an  event  driven  by  activity  within  the  biosphere  itself  through  respiration  by  phototrophic   prokaryotic  single-­‐celled  organisms  (anaerobic  photosynthesis  REF).    The  most  dramatic   increase  in  oxygen  levels,  by  a  factor  of  at  least  10! ,  occurred  relatively  rapidly  2.4  Ga  in  the   so-­‐called  Great  Oxidation  Event  (GOE)    The  GOE  (and  secondary  later  increases)  represents   one  of  the  most  important  examples  of  co-­‐evolution  between  life  and  the  planet  and,  as   such,  has  important  implications  for  astrobiology  and  sustainability  (Arndt  &  Nisbit  2012).        While  the  ultimate  source  of  increasing  O!  levels  is  universally  recognized  to  have   originated  in  the  biosphere  (i.e.  non-­‐oxygen  based  photosynthesis),  the  reasons  remain   debated  for  the  relatively  rapid  (on  geologic  timescales)  increase  in  atmospheric  O!  levels   in  the  GOE.    Proposed  causes  include  (Catling  &  Kasting  2007):  changes  in  the  chemical   state  of  outgassed  mantle  material;  a  decrease  in  atmospheric  methane  levels  (Kasting  et  al   1983,  Pavlov  2000);  changes  in  ocean  sinks  for  O! .    While  significant  uncertainty  remains   as  to  why  the  O!  levels  increased  when  they  did,  in  all  scenarios  strong  effects  on  the   coupled  Earth  systems  occur.    For  example,  an  anoxic  atmosphere  would  likely  have  had   high  concentrations  of  methane,  a  gas  that  on  a  molecule-­‐by-­‐molecule  basis  is  ~25  times   more  potent  as  a  greenhouse  absorber  than  CO! .    Increases  in  O!  would  remove  methane   from  the  atmosphere,  decreasing  average  global  temperatures.    The  possible  occurrence  of     22   several  “Earth”  episodes  (near  total  glacial  coverage)  occuring  during  or  just  after  the  GOE   may  also  be  related  to  this  removal  of  methane  (Kirschvink  et  al  2000).        Just  as  important  was  the  biosphere’s  own  reaction  to  the  increase  in  O!  levels.    The  GOE   was  a  double-­‐edged  sword  for  life  as  the  energetics  of  oxygen  chemistry  are  destructive  for   biological  activity,  yet  also  allow  for  more  efficient  metabolic  pathways  (Leigh,et  al  2007).     Aerobic  respiration,  for  example,  is  ~16  times  more  efficient  than  anaerobic  processes  at   generating  ATP,  the  primary  energy  carrier  for  metabolism  (Gaidos  &  Knoll  2012).    Thus   those  species  that  did  not  evolve  a  means  for  oxygen  detoxification  became  trapped  in   anaerobic  niches,  while  the  remainder  evolved  new  behaviors  that  allowed  them  to  thrive   in  the  newly  oxygen-­‐rich  atmosphere  and  oceans.    The  GOE  has  significant  implications  for  the  astrobiological  perspective  on  sustainability.  In   particular,  it  shows  that  in  at  least  one  instance  life  has  significantly  altered  the  chemical   composition  of  a  planet’s  atmosphere.    From  this  vantage  point,  the  current  era  of   anthropogenic  climate  change  can  be  seen  more  broadly  than  as  an  anomalous  byproduct   of  human  technological  progress.      From  the  GOE  we  see  that  at  least  one  time  in  the  past  life  strongly  forced  the  coupled  earth   systems.  In  light  of  this  the  alteration  of  atmospheric  chemistry  might  be  expected  to  occur   as  a  consequence  of  rapid  technologically-­‐driven  energy  harvesting  within  some  subset  of   SWEIT  trajectories.    Capturing  this  feedback  would  be  part  of  any  program  studying  such   trajectories  even  in  restricted  models  considering,  say,  only  coupling  between  𝑟!  and  𝑒! ,  i.  e.,         !!!!"   ∝  𝑓(𝑒! , !!!!" , … )    ,    where  𝑓(𝑒! , !!!!" , … )  models  the  links  between  energy  harvesting,  greenhouse  gas   production  and  and  other  quantities  which  drive  forcing  changes  in  𝑟! .      Consider  for  example  trajectory  of  our  own  species  over  the  last  10,000  years  of  its   evolution.    Fig  5a  tracks  the  human  population,  total  energy  consumption  and  as  a  proxy     23   for  planetary  forcing,  the  atmospheric  abundance  of  carbon  dioxide  (CO2).    The  plot  shows   the  rapid  coupled-­‐increase  in  all  3  quantities  within  the  last  century.    In  Fig  5b  we  present   the  same  data  as  a  trajectory  in  the  SWEIT  solution  space  introduced  in  section  3.  The  rapid   coupled  increase  in  population,  energy  consumption  and  planetary  forcing  mirrors  the   initial  phases  of  the  schematic  trajectories  shown  in  Fig  4.    The  question  we  hope  our   approach  might  address  would  be  to  explicate  the  various  forms  of  behavior  that  can   expected  in  the  future:  sustainability,  collapse  or  some  middle  ground.  As  we  have   demonstrated  the  astrobiological  perspective  allows  the  broad  considerations  of  links   between  planetary  systems  and  biospheric  activity  (in  this  case  due  to  a  single  SWEIT   species)  to  be  articulated  and,  hopefully,  modeled.     Figure  5a.    Plot  of  human  population,  total  energy  consumption  and  atmospheric  CO2     concentration  from  10,000  BCE  to  today.    Note  the  coupled  increase  in  all  3  quantities  over  the   last  century.     24      Consideration  of  Figures  5a  and  5b  indicates  that  modeling  the  trajectory  of  SWEIT   Figure  5b.    Same  data  as  figure  5a  but  plotted  as  trajectory  in  SWEIT  solution  space   ensembles  would  necessarily  include  the  evolution  of  atmospheric  changes.    One  outcome   of  such  modeling  may  be  to  find  that  a  greenhouse  phases  and  a  sustainability  “crisis”  be  a   generic  feature  of  at  least  some  subset  of  the  ensemble.    The  question  then  becomes  how   common  and  enduring    such  a  phase  is,  and,  more  importantly,  what  characterizes   successful  paths  out  of  such  a  crisis,  i.e.,  what  leads  to  long  mean  SWEIT  lifetimes  𝐿?   Sustainability  science  and  astrobiology  both  seek  to  understand  the  intimate,  symbiotic,     and  continually  evolving  connections  between  life  and  host  planets.  Sustainability  science   5.    Discussion  and  Summary   is  focused  on  the  effects  of  one  particular  species  during  one  particular  epoch,  whereas   astrobiology  broadens  its  purview  to  all  possible  species  on  Earth  or  elsewhere.  Both  fields   are  rapidly  changing  and  in  their  infancy.  Astrobiology's  concept  of  planetary  habitability  is   of  particular  relevance  to  sustainability  science,  while  the  latter's  concern  with  rapid   changes  in  the  biosphere  caused  by  a  single  intelligent  species  (ours)  informs  those   astrobiologists  considering  the  possible  existence  of  Species  with  Energy-­‐Intensive   Technology  (SWEITs)  on  other  planets.       25   In  this  paper  we  have  suggested  the  beginnings  of  a  research  program  that  we  hope  will   benefit  both  sustainability  science  and  astrobiology,  emphasizing  in  particular  how  recent   developments  in  astrobiology  are  of  direct  relevance  to  sustainability  science.  One   promising  avenue  comes  through  considering  an  ensemble  of  SWEITs  that  exist  now  or  in   the  past  (located  well  outside  of  our  solar  system  and  perhaps  even  our  Galaxy).  Each   SWEIT’s  history  defines  a  trajectory  in  a  multi-­‐dimensional  solution  space  with  axes   representing  quantities  such  as  energy  consumption  rates,  population  and  planetary   systems  forcing  from  causes  both  "natural"  and  driven  by  the  SWEIT  itself.    Using   dynamical  systems  theory,  these  trajectories  can  be  mathematically  modeled  in  order  to   understand,  on  the  astrobiology  side,  the  histories  and  mean  properties  of  the  ensemble  of   SWEITs,  as  well  as,  on  the  sustainability  science  side,  our  own  options  today  to  achieve  a   viable  and  desirable  future.  We  note  that  dynamical  systems  theory  as  we  have  presented  it   represents  only  one  theoretical  methodology  possible.    For  example,  the  use  of  network   theory  with  its  emphasis  on  multiple  cascading  paths  to  system  failures  (or  system   resiliency)  may  also  prove  useful  (Helping  2012).   Modeling  SWEIT/planet  evolution  in  the  way  we  have  described  may  allow  for  broad   classes  of  behavior  to  be  articulated.    A  future  research  project  may,  for  example,  explore  if   the  development  of  enhanced  greenhouse  forcing  is  an  expected  outcome  of  SWEIT   evolution.    This  could  occur  based  both  on  the  most  likely  energy  sources  harvested  early   in  a  species’  technological  development  and/or  planetary  changes  driven  a  results  of  other   SWEIT  activity.   We  note  in  conclusion  that  in  Payne  &  Clapham’s  (2012)  review  of  the  End-­‐Permian  mass   extinction  (subtitled  “An  Ancient  Analog  for  the  21st  Century?”)  the  authors  state:  “The   geological  record  is  increasingly  essential  as  an  archive  of  past  experiments  in  global   change…  The  best  -­‐  and  most  sobering  -­‐  analogs  for  our  near  future  may  lie  deeper  in   Earth's  past."    The  purpose  of  this  paper  has  been  to  broaden  such  a  conclusion  to  include   the  whole  of  astrobiological  studies.    Recent  studies  exploring  planetary-­‐scale  tipping   points  (Lenton  &  Williams  2013)  or  the  existence  of  planetary-­‐scale  “boundares”  as  safe   operating  limits  for  civilization  (Rockstrom  et  al  2009)  emphasize  that  a  astrobiological   perspective  is  already  present,  if  not  fully  recognized  in  many  modern  sustainability   studies.     26   Thus  the  evidence  is  indeed  strong  that  during  our  present  Anthropocene  epoch  the   coupled  Earth  systems  are  being  altered  on  an  extremely  rapid  time  scale.  Although  such   rapid  changes  are  not  a  new  phenomenon,  the  present  instance  is  the  first  (we  know  of)   where  the  primary  agent  of  causation  is  knowingly  watching  it  all  happen  and  pondering   options  for  its  own  future.  In  this  paper  we  have  argued  that  it  is  unlikely  that  this  is  the   first  time  such  an  event  as  occurred  in  cosmic  or  even  galactic  history.  The  point  is  to  see   that  our  current  situation  may,  in  some  sense,  be  natural  or  at  least  a  natural  and  generic   consequence  of  certain  evolutionary  pathways.    Given  that  fact  it  may  be  possible  to  use  the   data  and  perspectives  of  astrobiology  to  tell  us  something  about  optimal  pathways  forward.       One  point  is  clear,  both  astrobiology  and  sustainability  science  tell  us  that  the  Earth  will  be   fine  in  the  long  run.    The  prospects  are,  however,  less  clear  for  Homo  sapiens.   Acknowledgements:    We  wish  to  thank  Bruce  Balick,  Don  Brownlee,  David  Catling,  Dan   Watson,  Marcelo  Glieser,  John  Tarduno,  Marina  Alberti  and  James  Kasting  for  helpful     discussions.    Eric  Blackman  provided  helpful  comments  on  an  early  draft.    This  work  was   supported  in  part  with  funds  from  the  University  of  Rochester.           27     Bibliography   Alberti,  M.  Russo,  Tenneson.    (2013).    Snohomish  Basin  2060  Scenarios:    Adapting  to  an   Uncertain  Future.    Urban  Ecology  Research  Laboratory,  University  of  Washington,  Seattle.      Arndt,  Nicholas  T.,  Nisbet,  Euan  G.  (2012).    “Processes  on  the  Young  Earth  and  the  Habitats   of  Early  Life.”  Annual  Review  of  Earth  and  Planetary  Sciences,  vol  40,  pp.  521–49.    Arnould,  Jacques.    (2009).  “Astrobiology,  Sustainability  and  Ethical  Perspectives”.   Sustainability,  vol.  1,  pp  1323  –  1330.    Azua-­‐Bustos,  A.,  Vicuna,  R.,  &  Pierrehumbert,  R.,  2012,  “Early  Mars”,  in  Frontiers  in   AstroBiology,  Chris  Impey,  Jonathan  Lunine,  Joses  Funes  Eds,  Cambridge  University  Press,   Cambridge,  United  Kingdom  and  New  York,  NY,  USA,  157-­‐175    Barnosky,  Anthony  D.,  et  al.  (2011).  “Has  the  Earth’s  sixth  mass  extinction  already  arrived?”   Macmillan  Publishers  Limited,  Nature,  vol.  471,  pp  51-­‐57.    Baum,  Seth  D.  (2010).  “Is  Humanity  Doomed?  Insights  from  Astrobiology.”  Sustainability,   vol.  2,  pp  591-­‐603.    Barreiro,  Marcelo,  Fedorov,  Alexey,  Pacanowski,  Ronald,  Philander  S.  George.    (2007).     “Abrupt  Climate  Changes:  How  Freshening  of  the  Northern  Atlantic  Affects  the   Thermohaline  and  Wind-­‐Driven  Oceanic  Circulations.”    Annual  Review  of  Earth  and   Planetary  Sciences,  vol.  36,  pp.  33–58    Blois,  Jessica  L.,  Hadly,  Elizabeth  A.  (2009).      “Mammalian  Response  to  Cenozoic  Climatic   Change.”    Annual  Review  of  Earth  and  Planetary  Sciences,  vol.  37,  pp.  181–208.    Beltrami,  Edward,  (1986),  “Mathematics  for  Dynamic  Modeling”,  Academic  Press,  San  Diego,   CA.    Brauer  &  Castillo-­‐Chavez,  (2011),  “Mathematical  Models  in  Population  Biology  and   Epidemiology”,  Springer,  New  York   Caldeira,  K.  and  Karting,  J.F.  (1992),  “The  life-­‐span  of  the  biosphere  revisited”.  Nature  360,   721–723.    Carter,  B.,  McCrea,  W.  H.  (1983).    The  Anthropic  Principle  and  its  Implications  for  Biological   Evolution.  Royal  Society  Publishing,  pp  347-­‐363.      Castilo-­‐Rogez,  J.,  &  Lunine,  J.,  2012,  “Small  Habitable  Worlds”in  Frontiers  in  AstroBiology,   Chris  Impey,  Jonathan  Lunine,  Joses  Funes  Eds,  Cambridge  University  Press,  Cambridge,   United  Kingdom  and  New  York,  NY,  USA,  201-­‐228       28   Costanza,  Robert,  et  al.  (2007).    Sustainability  or  Collapse:  What  Can  We  Learn  from   Integrating  the  History  of  Humans  and  the  Rest  of  Nature?  Royal  Swedish  Academy  of   Sciences,  pp  522-­‐527.    Coustenis,  A.,  &  Blanc,  M.,  2012,  “Large  Habitable  Moons”,  in  Frontiers  in  AstroBiology,  Chris   Impey,  Jonathan  Lunine,  Joses  Funes  Eds,  Cambridge  University  Press,  Cambridge,  United   Kingdom  and  New  York,  NY,  USA,  175-­‐201    Decker,  Ethan  H.,  et  al.  (2000).  Energy  and  Material  Flow  Through  the  Urban  Ecosystem.   Annual  Reviews,  pp  685-­‐740.   www.annualreviews.org    Dressing  C.,  &  Charbonneau,  D.,    2013,  The  Occurence  Rate  of  Small  Planets  around  Small   Stars”,  ApJ,  767,  95    Feulner,  Georg.  (2009).  Climate  Modelling  of  Mass–Extinction  Events:A  Review.  Paper  for  the   ESLAB  2008  Special  Issue  of  IJA,  Postdam,  Germany.    Forgan,  Duncan,  Kipping,  David.  (2013).    Dynamical  Effects  on  the  Habitable  Zone  for  Earth-­‐ like  Exomoons.  Cornell  University  Library,  Cornell,  NY.  arXiv:1304.4377.     http://lanl.arxiv.org/abs/1304.4377v1    Goldblatt,  Colin,  Watson,  Andrew  J.  (2012).    The  Runaway  Greenhouse:  implications  for   future  climate  change,  geoengineering  and  planetary  atmosphere.  Cornell  University  Library,   Cornell,  NY.  arXiv:1201.1593.   http://arxiv.org/abs/1201.1593v1    Holling,  C.  S.  (2001).  “Understanding  the  Complexity  of  Economic,  Ecological,  and  Social   Systems.”  Ecosystems,  vol.  4,  pp  390-­‐405.    Howard,  A.W.,  (2013)  Observed  Properties  of  Extrasolar  Planets.  Science  340,  572-­‐576.    Howard,  A.,  et  al  2013,  Planet  Occurrence  with  0.25  AU  of  Solar-­‐type  stars  from  Kepler,  ApJ,   201,  15    Helbing,  D.,  2012,  “Globally  networked  risks  and  how  to  respond”,  Nature  497,  51– 59  Kane,  Stephen  R.,  Hinkel,  Natalie  R.  On  the  Habitable  Zones  of  Circumbinary  Planetary   Systems.  Cornell  University  Library,  Cornell,  NY.,  arXiv:1211.2812.   http://arxiv.org/abs/1211.2812v1    Kates,  Robert  W.  (2010).    Readings  in  Sustainability  Science  and  Technology.  CID  Working   Paper  No.  213,  Harvard  College,  Cambridge,  MA.    Kates,  Robert  W.  (2011).  “What  kind  of  a  science  is  sustainability  science?”  Proceedings  of   the  National  Academy  of  Science  of  the  United  States  of  America.  vol.  108,    no.  49,  pp.  19449– 19450.     29   http://www.pnas.org/content/108/49/19449    Kates,  Robert  W.  (2011).  From  the  Unity  of  Nature  to  Sustainability  Science:  Ideas  and   Practice.  CID  Working  Paper  No.  218,  Harvard  College,  Cambridge,  MA.    Kasting,  J.,  &  Kirschvink,  J.,  2012,  “Evolution  of  a  habitable  planet  Evolution  of  a  habitable   planet”  in  Frontiers  in  AstroBiology,  Chris  ,  Jonathan  Lunine,  Joses  Funes  Eds,  Cambridge   University  Press,  Cambridge,  United  Kingdom  and  New  York,  NY,  USA,  115-­‐132    Kleidon,  Axel.  (2011).  “Life,  hierarchy,  and  the  thermodynamic  machinery  of  planet  Earth.”   Elsevier,  Vol.  7,  Issue  4,  pp  424–460.   http://www.sciencedirect.com/science/article/pii/S1571064510001107    Kleidon,  Axel.  (2012).  How  does  the  Earth  system  generate  and  maintain  thermodynamic   disequilibrium  and  what  does  it  imply  for  the  future  of  the  planet?  Royal  Swedish  Academy  of   Sciences,  pp  1012-­‐1040.    Kopparapu,  Ravi  kumar,  et  al.  (2013).  Habitable  Zones  Around  Main-­‐Sequence  Stars:  New   Estimates.  Cornell  University  Library,  Cornell,  NY.,  arXiv:1301.6674.   http://arxiv.org/abs/1301.6674v2    Krutilla,  Kerry,  Reuveny,  Rafael.  (2006).  “The  systems  dynamics  of  endogenous  population   growth  in  a   renewable  resource-­‐based  growth  model.”  Elsevier,  vol.  56,  pp  256–  267.    Lenton,  T.M  &  Williams,  H.T.P.,  (2013),  “The  Origin  of  Planetary  Scale  Tipping  Points”,   Trends  in  Ecology  &  Evolution,  28,  380-­‐382    Lineweaver,  Charles  H.,  Chopra,  Aditya.  (2012).  “The  Habitability  of  Our  Earth  and  Other   Earths:  Astrophysical,  Geochemical,  Geophysical,  and  Biological  Limits  on  Planet   Habitability.”    Annual  Review  of  Earth  and  Planetary  Sciences,  vol.  40,  pp.  597–623.    Liu,  Jianguo,  et  al.  (2007).  “Complexity  of  Coupled  Human  and  Natural  Systems.”  Science,   vol.  317,  pp.  1513-­‐1516.    Loarie,  Scott  R.,  et  al.  (2009).  “The  velocity  of  climate  change.”  Macmillan  Publishers   Limited,  Nature,  vol.  462,  pp  1052-­‐1055.    May,  Robert  M.  (1977).  “Thresholds  and  breakpoints  in  ecosystems  with  a  multiplicity  of   stable  states.”  Nature  Publishing  Group,  Nature,  vol.  269,  pp.  471-­‐477.    Mayor,  M.  et  al.  (2011),  “The  HARPS  search  for  southern  extra-­‐solar  planets  XXXIV.   Occurrence,  mass  distribution  and  orbital  properties  of  super-­‐Earths  and  Neptune-­‐mass   planets”,  e-­‐print  arXiv:1109.2497    Marshall,  Charles  R.  (2006).    “Explaining  the  Cambrian  “Explosion”  of  Animals.”    Annual   Review  of  Earth  and  Planetary  Sciences,  vol.  34,  pp.  355–84.     30    Mann,  Michael  E.  (2007).  “Climate  Over  the  Past  Two  Millennia.”    Annual  Review  of  Earth   and  Planetary  Sciences,    vol.  35,  pp.  111–36.    McInerney,  Francesca  A.,  Wing,  Scott  L.  (2011).  “The  Paleocene-­‐Eocene  Thermal  Maximum:   A  Perturbation  of  Carbon  Cycle,  Climate,  and  Biosphere  with  Implications  for  the  Future.”   Annual  Review  of  Earth  and  Planetary  Sciences,  vol.  39,  pp  489–516.      Naganuma,  Takeshi.  (2009).  “An  Astrobiological  View  on  Sustainable  Life.”  Sustainability,   vol.  1,  pp.  827-­‐837.    Newman,  M.  E.  J.  (1997).  “A  Model  of  Mass  Extinction.”  Science  Direct,  vol.  189,  pp.  235-­‐252.    Petraitis,  Peter  S.,  Hoffman,  Catharine.  (2010).  “Multiple  stable  states  and  relationship   between  thresholds  in  processes  and  states.”  Marine  Ecology  Progress  Series,  vol.  413,  pp.   189-­‐200.    Pena-­‐Cabrera,  G.V.Y.,  Durand-­‐Manterola,  H.J.(2004),Number  of  Planets  with  life  in  the   galactic  habitable  zone  deduced  by  modified  Drake  Equation,  35th  COSPAR  Scientific   Assembly  35,  1903.    Payne,  Jonathan  L.,  Clapham,  Matthew  E.  (2012).    “End-­‐Permian  Mass  Extinction  in  the   Oceans:  An  Ancient  Analog  for  the  Twenty-­‐First  Century?”  Annual  Review  of  Earth  and   Planetary  Sciences,    vol.  40,  pp.  89–111.    Prantzos,  Nikos.  (2013).  “A  Joint  Analysis  of  the  Drake  equation  and  the  Fermi  Paradox”,    Roe,  Gerard.    (2009).  “Feedbacks,  Timescales,  and  Seeing  Red.”    Annual  Review  of  Earth  and   Planetary  Sciences,  vol.  37,  pp.  93–115.   International  Journal  of  Astrobiology,  12,  246-­‐253    Rockstrom,  J.,  et  al  (2009),  “Planetary  Boundaries:  Exploring  the  Safe  Operating  Space  for   Humanity”,  Ecology  and  Society,  14,  32-­‐55    Seager,  S.,  (2013)  ExoPlanet  Habitability.  Science  340,  577-­‐581.    Seager,  S,  2012,  “Searches  for  Habitable  Planets”  in  Frontiers  in  AstroBiology,  Chris  Impey,   Jonathan  Lunine,  Joses  Funes  Eds,  Cambridge  University  Press,  Cambridge,  United  Kingdom   and  New  York,  NY,  USA    Spiegel,  D  &  Turner,  E.,  (2012),  “Bayesian  analysis  of  the  astrobiological  implications   of  life’s  early  emergence  on  Earth”,  PNAS,  109,  395   Smith,  David  A.  (1977).    “Human  Population  Growth:    Stability  or  Explosion?”  Mathematics   Magazine,  Vol.  50,  pp.  186-­‐197.    Sullivan  &  Baross  2007,  “Planets  and  Life”,  eds.  W.  T.  Sullivan,  III  ,  J.  Baross,  Cambridge,  UK:   Cambridge  University  Press.     31    Sullivan  &  Carney  2007,  “The  History  of  Astrobilogical  Ideas”  in  Planets  and  Life,  eds.  W.  T.   Sullivan,  III  ,  J.  Baross,  Cambridge,  UK:  Cambridge  University  Press,  7-­‐45    Tattersall,  Ian,  Schwartz,  Jeffrey  H.    (2009).  “Evolution  of  the  Genus  Homo.”    Annual  Review   of  Earth  and  Planetary  Sciences,  vol.  37,  pp.  67–92.    Tarter,  J.,  (2007),  Searching  for  Extraterrestrial  Intelligence.  In  Planets  and  Life,  eds.  W.  T.   Sullivan,  III  ,  J.  Baross,  Cambridge,  UK:  Cambridge  University  Press.      “Sustainability  and  the  Ehrlich  equation.”  Population  Matters.  (2011).   http://www.populationmatters.org/wp-­‐content/uploads/ipat.pdf.    Wiley,  Keith  B.  (2011).  The  Fermi  Paradox,  Self-­‐Replicating  Probes,  and  the  Interstellar   Transportation  Bandwidth.  Cornell  University  Library,  Cornell,  NY.,  arXiv:1111.6131.   http://arxiv.org/abs/1111.6131    Watson,  Andrew  J.  (2008).  “Implications  of  an  Anthropic  Model  of  Evolution  for  Emergence   of  Complex  Life  and  Intelligence.”  Astrobiology,  vol.  8,  pp.  1-­‐11.    Winguth,  Arne  Max  Erich.  (2011).  The  Paleocene-­‐Eocene  Thermal  Maximum:  Feedbacks   Between  Climate  Change  and  Biogeochemical  Cycle.  The  University  of  Texas  at  Arlington,   Arlington,  TX.    Ward,  P,  (2007),  “Mass  Extictions”,  In  Planets  and  Life,  eds.  W.  T.  Sullivan,  III  ,  J.  Baross,   Cambridge,  UK:  Cambridge  University  Press.    Wallenhorst,  S.  G.,(1981),  The  Drake  Equation  Reexamined.  The  Quarterly  Journal  of  the   Royal  Astronomical  Society  22,  380.                                           32  
1712.04042
2
1712
2018-01-21T17:24:02
The California-Kepler Survey. IV. Metal-rich Stars Host a Greater Diversity of Planets
[ "astro-ph.EP" ]
Probing the connection between a star's metallicity and the presence and properties of any associated planets offers an observational link between conditions during the epoch of planet formation and mature planetary systems. We explore this connection by analyzing the metallicities of Kepler target stars and the subset of stars found to host transiting planets. After correcting for survey incompleteness, we measure planet occurrence: the number of planets per 100 stars with a given metallicity $M$. Planet occurrence correlates with metallicity for some, but not all, planet sizes and orbital periods. For warm super-Earths having $P = 10-100$ days and $R_P = 1.0-1.7~R_E$, planet occurrence is nearly constant over metallicities spanning $-$0.4 dex to +0.4 dex. We find 20 warm super-Earths per 100 stars, regardless of metallicity. In contrast, the occurrence of warm sub-Neptunes ($R_P = 1.7-4.0~R_E$) doubles over that same metallicity interval, from 20 to 40 planets per 100 stars. We model the distribution of planets as $d f \propto 10^{\beta M} d M$, where $\beta$ characterizes the strength of any metallicity correlation. This correlation steepens with decreasing orbital period and increasing planet size. For warm super-Earths $\beta = -0.3^{+0.2}_{-0.2}$, while for hot Jupiters $\beta = +3.4^{+0.9}_{-0.8}$. High metallicities in protoplanetary disks may increase the mass of the largest rocky cores or the speed at which they are assembled, enhancing the production of planets larger than 1.7 $R_E$. The association between high metallicity and short-period planets may reflect disk density profiles that facilitate the inward migration of solids or higher rates of planet-planet scattering.
astro-ph.EP
astro-ph
Draft version January 23, 2018 Preprint typeset using LATEX style AASTeX6 v. 1.0 THE CALIFORNIA-KEPLER SURVEY. IV. METAL-RICH STARS HOST A GREATER DIVERSITY OF PLANETS Erik A. Petigura1,7,8, Geoffrey W. Marcy2, Joshua N. Winn3, Lauren M. Weiss4,9, Benjamin J. Fulton1, Andrew W. Howard1, Evan Sinukoff5, Howard Isaacson2, Timothy D. Morton4, and John Asher Johnson6 8 1 0 2 n a J 1 2 . ] P E h p - o r t s a [ 2 v 2 4 0 4 0 . 2 1 7 1 : v i X r a 1California Institute of Technology, Pasadena, CA 91125, USA 2Department of Astronomy, University of California, Berkeley, CA 94720, USA 3Princeton University, Princeton, NJ 08544, USA 4University of Montreal, Montreal, QC, H3T 1J4, Canada 5Institute for Astronomy, University of Hawai'i at M¯anoa, Honolulu, HI 96822, USA 6Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA [email protected] 8Hubble Fellow 9Trottier Fellow ABSTRACT Probing the connection between a star's metallicity and the presence and properties of any associated planets offers an observational link between conditions during the epoch of planet formation and ma- ture planetary systems. We explore this connection by analyzing the metallicities of Kepler target stars and the subset of stars found to host transiting planets. After correcting for survey incomplete- ness, we measure planet occurrence: the number of planets per 100 stars with a given metallicity M. Planet occurrence correlates with metallicity for some, but not all, planet sizes and orbital periods. For warm super-Earths having P = 10–100 days and RP = 1.0–1.7 R⊕, planet occurrence is nearly constant over metallicities spanning −0.4 dex to +0.4 dex. We find 20 warm super-Earths per 100 stars, regardless of metallicity. In contrast, the occurrence of warm sub-Neptunes (RP = 1.7–4.0 R⊕) doubles over that same metallicity interval, from 20 to 40 planets per 100 stars. We model the distri- bution of planets as df ∝ 10βM dM, where β characterizes the strength of any metallicity correlation. This correlation steepens with decreasing orbital period and increasing planet size. For warm super- Earths β = −0.3+0.2−0.2, while for hot Jupiters β = +3.4+0.9−0.8. High metallicities in protoplanetary disks may increase the mass of the largest rocky cores or the speed at which they are assembled, enhancing the production of planets larger than 1.7 R⊕. The association between high metallicity and short- period planets may reflect disk density profiles that facilitate the inward migration of solids or higher rates of planet-planet scattering. Keywords: editorials, notices - miscellaneous - catalogs - surveys 1. INTRODUCTION Exploring the connection between planets and their host stars has been a long-standing focus of exoplanet astronomy. Host star metallicity is thought to reflect the metallicity of the protostellar nebula and the pro- toplanetary disk from which planets form. Metal-rich protoplanetary disks are thought to have enhanced sur- face densities of solids. Viewed in the context of core- accretion theory (Lissauer 1995; Pollack et al. 1996), one might expect metal-rich disks to form terrestrial planets and the cores of gas giant planets with greater efficiency than metal-poor disks. If true, metal-rich stars should host greater numbers of gas giant and terrestrial planets. This prediction can be tested by studying the correlation (or lack thereof) between [Fe/H] and planet occurrence. The extent to which stellar metallicity correlates with the presence or absence of planets has been the sub- ject of many previous studies. Gonzalez (1997) observed that the first four extrasolar planets discovered orbited metal-rich stars and concluded that metal-rich stars have metal-rich protoplanetary disks which form plan- ets more efficiently. As the sample of Doppler-detected planets grew into the hundreds, various studies noted a preference for Jovian-mass planets to orbit stars with super-solar metallicities (e.g. Santos et al. 2004; Fischer & Valenti 2005). However, as Doppler surveys pushed into lower regimes of planet mass, various authors noted that the correlation between planet occurrence and host 2 star metallicity appeared to weaken (e.g. Sousa et al. 2008; Ghezzi et al. 2010). The prime Kepler mission (2009–2013) revealed more than 4000 planet candidates with sizes as small as Mer- cury (Barclay et al. 2013). In contrast to previous studies that explored the connection between host star metallicity and planet mass, Kepler studies have fo- cused mainly on the connection between metallicity and planet size. While a handful of Kepler planets have well- measured masses through radial velocities (RVs; e.g., Marcy et al. 2014) or transit-timing variations (TTVs; e.g., Hadden & Lithwick 2017), the vast majority of Ke- pler planets have unknown masses. Current RV instru- ments require bright targets (V < 13 mag), and TTV measurements require tightly packed, multi-planet sys- tems. Both TTV and RV mass measurements are only possible on a small fraction of the total Kepler planet sample. While the Kepler sample provided a large sample of planets for planet-metallicity studies, extensive follow- up spectroscopy was first required to measure host star metallicities. Buchhave et al. (2012) measured the metallicities of 152 stars harboring 226 planets and ob- served that while planets larger than 4 R⊕ orbit metal- rich stars, smaller planets orbit stars with wide-ranging metallicities. Later, using an augmented sample of 405 stars hosting 600 planets, Buchhave et al. (2014) argued for three distinct stellar metallicity distributions, with breakpoints at 1.7 R⊕ and 3.9 R⊕. In contrast, Schlauf- man (2015) found no evidence for different metallicity distributions above and below 1.7 R⊕ and raised several concerns regarding the statistical validity of the Buch- have et al. (2014) analysis. RV mass measurements of transiting planets have re- vealed that planets smaller than 1.7 R⊕ have bulk den- sities consistent with rocky compositions (e.g. Weiss & Marcy 2014; Rogers 2015), at least for the short- period planets (P (cid:46) 20 days) that are amenable to these RV mass measurements. Consequently, the degree to which metallicity correlates with the occurrence of plan- ets smaller than 1.7 R⊕ is of particular interest. On this point, there is disagreement in the literature. Wang & Fischer (2015), using low precision metallicities from the Kepler Input Catalog (KIC; Brown et al. 2011), reported that the rate of planets smaller than 1.7 R⊕ is 1.72+0.18−0.17 times higher for stars with [Fe/H] > 0 dex compared to stars [Fe/H] < 0 dex. In contrast, Buchhave & Latham (2015) found no evidence of a metallicity enhancement among stars hosting planets smaller than 1.7 R⊕. A long-standing limitation of Kepler metallicity stud- ies was that reliable metallicities did not exist for a rep- resentative number of Kepler field stars. In constructing the KIC, Brown et al. (2011) placed a prior on the Kepler field metallicities based on the metallicities of nearby stars, as measured by Nordström et al. (2004), which have a mean of −0.14 dex and dispersion of 0.19 dex. However, it was not clear whether Kepler field stars, which are typically ∼1 kpc from Earth, would follow the metallicity distribution of nearby stars. Such metallicity offsets have been invoked to explain differences in planet occurrence rates between the Kepler field and the solar neighborhood. Howard et al. (2012) measured a hot Jupiter occurrence rate of 0.4 ± 0.1%, roughly 40% that in the solar neighborhood (1.2± 0.4%, Wright et al. 2012), and speculated that "a paucity of metal-rich stars in the Kepler sample is one possible explanation." Today we know that Kepler field stars have a higher mean metallicity than solar neighborhood stars. The Large Sky Area Multi-object Fibre Spectroscopic Tele- scope (LAMOST; Zhao et al. 2012; Cui et al. 2012; Luo et al. 2012) is instrumented with a multiplexed, low- resolution spectrometer (4000 fibers, R = 1800) and can therefore efficiently gather spectra for large samples of Kepler target stars with and without transiting plan- ets. Dong et al. (2014), through analyzing the LAMOST Data Releases 1 and 2, found that the mean metallicity of 12,000 Kepler field stars was −0.04 dex, much closer to solar than to the value of −0.14 dex assumed in the construction of the KIC. Guo et al. (2017) also found a near-solar mean metallicity of −0.04 dex in an analy- sis of 610 Kepler field stars observed by the Hectochelle R = 34, 000 spectrometer at the MMT. Thus, a metal- licity offset cannot explain differences in the hot Jupiter rates. The LAMOST datasets permitted several break- throughs in Kepler planet-metallicity studies, by mea- suring the true metallicity distribution of bright (Kp < 14) Kepler field stars. Mulders et al. (2016) analyzed the LAMOST metallicities and a sample of 665 planet candidates, and found that occurrence rate of hot small planets (P < 10 days, RP < 4 R⊕) is three times higher among super-solar metallicity hosts compared to sub- solar hosts. Dong et al. (2017) also analyzed LAMOST metallicities and a sample of 295 planets and reported a similar trend, also noting that hot Neptune-size planets are typically single. Here, we examine the connection between planets and stellar metallicity using spectroscopy from the California-Kepler Survey (CKS). Given that the CKS produced a homogeneous set of highly precise metal- licities of 1305 stars, we can explore this connection in unprecedented detail. A key advantage of the CKS sam- ple is the highly-precise planet radii (10% precision) and stellar metallicities (0.04 dex precision) compared to pre- vious studies. The CKS sample has high purity (i.e. low false positive rate) due to extensive vetting of false pos- itives. This dataset has already revealed new features in the planet radius distribution, most notably a gap in the size distribution of small (RP = 1–4 R⊕ planets) reported by Fulton et al. (2017). We find that planets larger than Neptune are pref- erentially found around metal-rich stars, while planets smaller than Neptune are found around stars of wide- ranging metallicity. For super-Earth-size planets (RP = 1.0–1.7 R⊕), we observe a positive metallicity correla- tion for P = 1–10 days, but no correlation for P = 10– 100 days. In contrast, rates of sub-Neptune-size planets (RP = 1.7–4.0 R⊕) correlate with metallicity over P = 1–100 days. Planets larger than Neptune are an order of magnitude less common than planets smaller than Nep- tune, and probing the possible metallicity correlations is more challenging. However, we observe strong metallic- ity correlations for both Jovian-size (RP = 8.0–24.0 R⊕) and sub-Saturn-size (RP = 4.0–8.0 R⊕) planets. We describe our planet and stellar samples in Sec- tion 2. Section 3 explores the distribution of planets in the P –RP plane as a function of metallicity, and high- lights areas where metallicity plays a strong effect. In Sections 4–6, we compare host star metallicities to the Kepler field star distribution and compute planet oc- currence as a function of metallicity. We summarize our findings in Section 7, and offer some interpretations of the observed trends. 2. SAMPLE Studies of planet occurrence require both a sample of planets P and a parent stellar sample S from which the planets are drawn. We construct P from the CKS sample (Section 2.1) and apply a series of filters aimed at creating a well-defined sample of high purity with well-measured radii (Section 2.2). We then construct S from the Kepler field stars after applying the same set of filters used to construct P (Section 2.3). Because the metallicities are not known for every star in S, we em- ploy the LAMOST parameters as a proxy (Section 2.4). 2.1. Initial Planet Sample CKS is a large-scale spectroscopic survey of 1305 Ke- pler Objects of Interest (KOIs). The sample selec- tion, spectroscopic observations, and spectroscopic anal- ysis are described in detail in Petigura et al. (2017b, hereafter Paper I). In brief, the sample was initially constructed by selecting all Kepler Objects of Interest (KOIs) brighter than Kp = 14.2 mag.1 A KOI is a Kepler target star which showed periodic photometric dimmings indicative of planet transits. However, not all KOIs have received the necessary follow-up attention needed to confirm the planets. 1 Kepler magnitude; Kp ≈ V − 0.4 mag for G2 stars. 3 Over the course of the project, we included additional targets to cover different planet populations, including multi-candidate hosts, Ultra Short Period candidates, and Habitable Zone candidates. We obtained spectra at the 10 m Keck Telescope using the High Resolu- tion Echelle Spectrometer at a resolution of R = 60, 000 (Vogt et al. 1994). A key feature of the CKS dataset is that stars were observed to a consistent signal-to-noise ratio (S/N) of 45/pixel−1 at the peak of the blaze func- tion near 5500 Å. In Paper I, we extracted the following properties for each star: effective temperature Teff, surface gravity log g, metallicity [Fe/H], and projected rotational ve- locity v sin i using the SpecMatch (Petigura 2015) and SME@XSEDE codes (Paper I). In Johnson et al. (2017, hereafter Paper II), we converted the spectroscopic properties of Teff, log g, and [Fe/H] into stellar mass M(cid:63), radius R(cid:63), and age. This conversion is facilitated by the publicly available isochrones Python package (Morton 2015). Stellar mass, radius, and age are measured to 4%, 11%, and 30%, respectively. We then recompute planetary radii and equilibrium temperatures using our updated CKS parameters and the results from transit fitting per- formed by Mullally et al. (2015). Paper II provides up- dated stellar and planetary properties for 1305 KOIs and 2025 planet candidates, which are the starting point for our planet sample P. 2.2. Filtered Planet Sample We applied a series of filters, described below in bul- lets, to the CKS sample to arrive at our final planet sample P. The filters restrict the range of stellar prop- erties included in our analysis. Since we aim to explore the connection between metallicity and planet size, we also require the planets have well-determined radii. Where possible, we filter based on the DR25 stellar properties table of Mathur et al. (2017). These cuts may be applied homogeneously to the field star sample S. Some filters involve follow-up observations not available for every star in S. In Section 2.3, we quantify the number of field stars that would have been excluded had comparable follow-up been performed. The applied filters are as follows: 1. Stellar brightness. We restrict our sample to the magnitude-limited sub-sample of the CKS sample (i.e. Kp < 14.2 mag). 2. Stellar effective temperature. The spectroscopic tools used in Paper I produce reliable results for Teff = 4700–6500 K. We restrict our analysis to stars having photometric Teff = 4700–6500 K (DR25 stellar properties table). 4 3. Stellar surface gravity. We restrict our analysis to stars having photometric log g = 3.9–5.0 dex (DR25 stellar properties table). 4. Stellar dilution. Dilution from nearby stars can also alter the apparent planetary radii. Furlan et al. (2017) compiled high resolution imaging ob- servations performed by several groups.2 When a nearby star is detected, Furlan et al. (2017) com- puted a radius correction factor (RCF), which ac- counts for dilution assuming the planet transits the brightest star. We elect to not apply this correction factor, but conservatively exclude KOIs where the RCF is larger than 5%. 5. Planet orbital period. We remove planet candi- dates with orbital periods longer than 350 days. This excludes planets that only transit once or twice during Kepler observations, which have a higher false positive rate (Mullally et al. 2015). 6. Planet false positive designation. We exclude can- didates that are identified as false positives accord- ing to Paper I. 7. Planets with grazing transits. Finally, we exclude stars having grazing transits (b > 0.9), which have suspect radii due to covariances with the planet size and stellar limb-darkening during the light curve fitting. Our successive filters, along with a running tally of the number planets which pass them, are listed in Table 1. The filters on dilution and grazing transits each elim- inate a small percentage (8% and 5%) of planets with imprecise radii, even after CKS spectroscopy. Similar filters clarified the bimodal planet radius distribution presented by Fulton et al. (2017), Paper III in the CKS series. In total, P contains 970 planets orbiting 662 stars that pass all filters. Figure 1 shows planets on the [Fe/H]– RP plane as successive cuts are applied. Even though planets across this plane are filtered out, planets larger than Neptune have a higher rate of removal. Properties of the stars hosting these planets are shown in Figure 2. Table 1. Filters Applied to Planet Sample Filter Full sample Kp < 14.2 mag Teff = 4700 − 6500 K log g = 3.9 − 5.0 dex P < 350 days Not a false positive Radius correction factor < 5% Not a grazing transit (b < 0.9) npl,pass npl,pass,run 2025 1359 1328 1212 1191 1105 1017 970 2025 1359 1977 1899 1965 1861 1891 1870 fpl,pass,run n(cid:63),pass,run 1279 954 929 829 817 758 695 662 1.000 0.671 0.977 0.913 0.983 0.928 0.920 0.954 Note-Summary of the filters applied to the CKS catalog to create planet sample P. The column labeled npl,pass is the total number of KOIs that pass a specific filter and npl,pass,run is a running tally of KOIs that pass all filters. For filter i, fpl,pass,run(i) = npl,pass,run(i)/npl,pass,run(i − 1). For example, fpl,pass,run(3) = npl,pass,run(3)/npl,pass,run(2) = 1328/1359 = 0.977. The column labeled n(cid:63),pass,run gives the numbers of unique stars hosting the npl,pass,run KOIs. 2.3. Field Star Sample 2 Alphabetical by author: Adams et al. (2012, 2013); Baranec et al. (2016); Cartier et al. (2015); Dressing et al. (2014); Everett et al. (2015); Gilliland et al. (2015); Horch et al. (2012, 2014); Howell et al. (2011); Law et al. (2014); Lillo-Box et al. (2012, 2014); Wang et al. (2015a,b); Ziegler et al. (2017). In order to compare the properties of stars with and without transiting planets, we need to consider the par- ent population of Kepler field stars S, from which P is drawn. We begin with the Mathur et al. (2017) catalog that lists all 199991 stars observed at some point during the Kepler mission. Where possible, we apply the same set of filters to construct S that were used to construct P. Table 2 summarizes the number of stars that pass 5 Figure 1. Host star metallicity vs. planet size after the application of successive filters. Panel (a): all KOIs in the CKS sample. Panels (b)–(j): blue points show the KOIs that pass successive filters in stellar brightness, effective temperature, surface gravity, planetary orbital period, false positive disposition, radius precision, dilution due to nearby stars, and impact parameter. The gray points show KOIs that do not pass one or more of the filters. cuts on Kp, photometric Teff, and photometric log g. A total of 36959 stars pass all cuts. The distribution of field star properties is shown in Figure 2. The filters used to construct P were not simply cuts on the stellar properties, but also on quality of the CKS stellar radii and the properties of the planet candidates. We cannot apply these same filters to the field stars on a star-by-star basis because these stars do not have CKS stellar radii or detected planets. However, we must as- sess whether the remaining filters from Section 2.2 would have excluded a significant fraction of field stars, assum- ing all stars received similar follow-up attention. Here, we quantify the number of field stars that would have been excluded from S had comparable follow-up been performed. 1. Reliable spectroscopic parameters. Our initial sam- ple of planets orbit stars for which the spectro- scopic analysis described in Section 2 produced reliable results. For a star to be included, v sin i must be less than 20 km s−1. This selection ef- fect excludes some stars that are typically near 6500 K, where v sin i values begin to exceed 20 km s−1. After filtering on Kp, photometric Teff, and photometric log g, 3.1% of stars are excluded be- cause they do not have reliable spectroscopic pa- rameters. 2. Stellar dilution. 7.8% of planet candidates were excluded because the Kepler apertures contained enough flux from neighboring stars, such that the planet radii required a correction factor larger than 5%. The fraction of stars excluded depends on the crowding of Kepler field stars. We make the assumption that this distribution is indepen- dent of KOI status. Had all field stars received a similar level of high-contrast imaging follow-up, 7.8% would have companions sufficiently bright and close to warrant a RCF of > 5%. 3. Planet orbital period. For a KOI to be included in our sample, we required that the orbital period be less than 350 days. Such a cut is strictly a cut on the planet properties and does not preclude any stars from being included in S. 4. Planet false positive disposition. A star must first be identified as a KOI in order to then be desig- nated as a false positive. Therefore, the false pos- itive filter does not exclude significant numbers of field stars from S. 5. Planets with grazing transits. We require that the planets have non-grazing impact parameters. As with the radius filter, it does not affect inclusion in S. However, this cut on impact parameter must be factored into the geometric transit probability, which affects the occurrence calculation described in Section 4. −0.6−0.4−0.20.00.20.4[Fe/H]Full sample (2025)aKp < 14.2 mag (1359)bTeff = 4700−6500 K (1328)clog g = 3.9−5.0 dex (1212)d0.512481632Rp(R⊕)−0.6−0.4−0.20.00.20.4[Fe/H]P < 350 days (1191)e0.512481632Rp(R⊕)Not FP (1105)f0.512481632Rp(R⊕)dilution < 5% (1017)g0.512481632Rp(R⊕)Not grazing (970)h 6 When we account for stars that would have been ex- cluded from our sample due to unreliable spectroscopic parameters or the presence of nearby stars, we find that the planet sample P was drawn from a parent stellar sample S containing 36959×0.922×0.969 = 33020 stars. Table 2. Filters Applied to Stellar Sample Cut n(cid:63),pass 199991 Full sample 81758 Kp < 14.2 mag Teff = 4700 − 6500 K 168885 log g = 3.9 − 5.0 dex 162854 n(cid:63),pass,run 199991 81758 62751 36959 f(cid:63),pass,run 1.000 0.409 0.768 0.589 Note-Summary of the filters applied to S. See Table 1 for column descriptions. 2.4. Metallicity Distribution of Kepler Field Stars Here, we characterize the metallicity distribution of the parent sample S. While the KIC (Brown et al. 2011) and its updates (e.g. Mathur et al. 2017) tabulate metallicities for every Kepler target star, they are inad- equate for characterizing the metallicity distribution of the Kepler field given their low precision. Instead, we use the LAMOST/LASP stellar parameters (Luo et al. 2015) from Data Release 3 (DR3).3 Following De Cat et al. (2015), we crossmatch the LAMOST-DR3 with the KIC by identifying LAMOST spectra taken within 1.2 arcsec of a KIC target star. This crossmatching re- sults in 29997 stars in common having Kp < 14.2 mag. We note a systematic offset between the LAMOST and CKS metallicity scales of ≈0.04 dex when comparing 476 stars in common. We apply a correction, described in Appendix A, but estimate that residual systematic offsets of ≈0.01 dex remain. We apply the same set of filters to the LAMOST stars that we applied to P. These include cuts in Kp, Teff, and log g. After applying these cuts, we are left with 14382 stars with LAMOST parameters. The filters are summarized in Table 3 and the distribution of LAMOST stellar properties is shown in Figure 2. Table 4 summarizes the metallicity distribution of P and S measured through different methods. The mean metallicity of S, as measured by LAMOST, is −0.01 dex, similar to the mean metallicity of the filtered planet sam- ple P, +0.03 dex, as measured by CKS. We note that the mean metallicity of S, according to the DR25 stellar properties table (Mathur et al. 2017), is −0.19 dex. This low value is due to the low metallicity prior used in the photometric modeling. Table 3. Summary of Cuts to LAMOST Sample Cut n(cid:63),pass n(cid:63),pass,run 29997 Full sample 29997 Kp < 14.2 mag 29997 29997 Teff = 4700 − 6500 K 23551 23551 log g = 3.9 − 5.0 dex 18625 14382 f(cid:63),pass,run 1.000 1.000 0.785 0.611 Note-Summary of the cuts applied to the LAMOST DR-2 sample. See Table 1 for column descriptions. 3 http://dr3.lamost.org 7 Table 4. Comparison of Planet Host and Field Star Metallicities P (spec) S (phot) S (spec) (dex) (dex) −0.005 0.207 (dex) −0.187 0.259 0.031 0.185 0.006 −0.069 0.052 0.148 0.001 −0.320 −0.160 −0.020 0.002 −0.116 0.020 0.131 Mean RMS SEM 25% 50% 75% Note-Summary of metallicity distributions for different stellar samples. P (spec) refers to the CKS metallicities of our filtered list of planet hosts (Section 2.2). S (phot) refers to the photometric metallicities of our par- ent stellar sample (Section 2.3). S (spec) refers to the metallicities of S measured us- ing LAMOST data (Section 2.4). S (phot) and S (spec) have significantly different mean metallicities, due the metallicity prior im- posed by Mathur et al. (2017). 3. METALLICITIES OF PLANET HOSTS In this section, we examine the metallicities of planet host stars with respect to planet size and orbital period. We first briefly note the trends seen within the planet sample P, without reference to field star metallicities (Section 3.1). However, features in P –RP –[Fe/H] distri- bution are much more apparent when compared to the metallicities of the parent stellar sample S. We perform this analysis in Section 3.2. By searching for evidence of elevated metallicity in the host stars of certain types of planets, we can identify planet classes that show some positive correlation with metallicity. The advantage of this approach is that it does not require any modeling of the Kepler survey completeness. A more difficult but ultimately more useful approach is to compute planet occurrence as a function of metallicity, which requires modeling survey completeness. This approach is taken in Section 6. 3.1. Properties of Planet Hosts In Figure 3, we show the planet sizes and host star metallicities for the 2025 planet candidates in the CKS sample. Planets smaller than Neptune are found around stars of wide-ranging metallicities, while there is a deficit of planets larger than Neptune around stars with sub- solar metallicity. To investigate variation in average host star metallic- ities with planet sizes, we divided the metallicity mea- √ surements according to planet size. Bins span a factor of 2 in RP for planets smaller than 4 R⊕. Larger plan- ets are placed into bins spanning a factor of 2, owing to lower numbers. Figure 3 shows the mean metallici- ties for these planet radius bins along with the 25% and 75% quantiles. We observe a gradual upward trend in mean host star metallicity from smaller to larger planets. Buchhave et al. (2012, 2014) observed a similar trend in smaller samples of planet hosts. Figure 3 also shows host star metallicity as a function of orbital period. Unlike the RP –[Fe/H], there are no large regions of the P –[Fe/H] plane clearly devoid of planets. After computing the mean metallicity in bins of P spanning 0.25 dex, we observe a slight increase in mean metallicity of about 0.05 dex with decreasing orbital period over P = 1–10 days. Mulders et al. (2016) observed a similar trend in a smaller sample of planet hosts. 3.2. Comparison to Field Stars 8 Figure 2. The distribution of stellar properties for the three samples of stars considered in this work. Panel (a): blue points show Teff and log g of CKS planet hosts that passed all filters P; gray points represent all CKS planet hosts. Panels (b) and (c) show the same quantities as panel (a), but for the Kepler target stars and the LAMOST sample, respectively. Panels (d)–(f): Distributions of host star Kp. Panels (g)–(i): distributions of host star metallicity from different catalogs. The sub-solar mean metallicity of the Kepler target stars (h) is a reflection of the low metallicity prior used in the photometric modeling. Here, we examine the effect of host star metallicity on the 2D distribution of planet size and orbital period. We divided P into four bins of host star metallicity with boundaries at −0.116, +0.020, and +0.131 dex. The boundaries of the bins equally divide the stars in S (see Table 4). Figure 4 shows the distribution of planets in the P –RP plane for the different metallicity bins. If the occurrence of planets were independent of host star metallicity, then the distribution of planets in each plot would be indistinguishable. The panels of Figure 4 show clear differences, indicating that metallicity is associated with certain types of planets, and that the strength of that enhancement depends on both P and RP . To facilitate our investigation into the effect of metal- licity across the P –RP plane, we define several regions. As a matter of convenience, we define a nomenclature for referring to these different regions. While the bound- aries are matters of taste, we choose physically moti- vated boundaries, when possible. We consider four do- mains of planet size, defined below: 1. Jupiters. RP = 8–24 R⊕. The lower limit is motivated by the fact that planets larger than 8 R⊕ tend to have masses ranging from 100 to 10,000 M⊕, while planets smaller than 8 R⊕ tend to have lower masses ranging from 6 to 60 M⊕ (see, e.g., Petigura et al. 2017a, Figure 9). At MP ≈ 100 M⊕ electron degeneracy pressure be- gins contribute significantly to a planet's pressure 4000500060007000Teff (K)3.03.54.04.55.0logg (dex)CKSa4000500060007000Teff (K)3.03.54.04.55.0logg (dex)Fieldb4000500060007000Teff (K)3.03.54.04.55.0logg (dex)LAMOSTc10121416Kepmag020406080100120Number of StarsCKSd10121416Kepmag0250050007500100001250015000Number of StarsFielde10121416Kepmag01000200030004000Number of StarsLAMOSTf−1.0−0.50.00.5[Fe/H]050100150200250300Number of StarsCKSg−1.0−0.50.00.5[Fe/H]0100002000030000Number of StarsFieldh−1.0−0.50.00.5[Fe/H]0100020003000400050006000Number of StarsLAMOSTi support (Zapolsky & Salpeter 1969). Thus 8 R⊕ approximates the dividing line between planets with different pressure support and interior struc- tures. The upper radius limit of 24 R⊕ is based on the known sizes of hot Jupiters. The most highly irradiated Jovians discovered to date can exceed 2 Rjup (e.g., WASP-79b, Smalley et al. 2012).4,5 2. Sub-Saturns. RP = 4–8 R⊕. Sub-Saturns are typically typically less massive the larger Jupiters. They are roughly 10× more rare than the smaller sub-Neptunes, suggesting a different formation pathway. A radius of 4 R⊕ approximates a break- point in the planet size distribution where planet occurrence rises rapidly with decreasing size (see Fulton et al. 2017). 3. Sub-Neptunes. RP = 1.7–4.0 R⊕. The lower ra- dius limit corresponds to a likely transition radius between rocky planets and planets that have en- velopes that are an appreciable fraction of the total planet size. Among the observations supporting such a transition are measurements of planet bulk density from transits and radial velocities (Marcy et al. 2014; Weiss & Marcy 2014; Rogers 2015) and the observation of a gap in the radius distribution of planets by Fulton et al. (2017). 4. Super-Earths. RP = 1.0–1.7 R⊕. Planets that are smaller than sub-Neptunes. Few planets in this size range have well-measured masses, but those that do are often consistent with rocky composi- tions. We also consider three domains of orbital period: 1. Hot Planets. P = 1–10 days. The upper limit of 10 days corresponds to a breakpoint in distri- bution of planet occurrence with orbital period (Howard et al. 2012; Fressin et al. 2013). Be- low P ≈ 10 days, planet occurrence per log P is observed to decline, while planet occurrence is roughly constant for longer periods. 2. Warm Planets. P = 10–100 days. An intermediate range of orbital periods. While warm planets are intrinsically more common than hot planets, they represent about half of the total sample due to falling completeness and transit probability with increasing orbital period. 3. Cool Planets. P = 100 − 350 days. The longest period planets included in our survey. There are 4 Exoplanet Archive (Akeson et al. 2013) 5 https://exoplanetarchive.ipac.caltech.edu/ 9 very few planets in our sample with such long pe- riods due to falling completeness and decreasing transit probability. Here, we note some qualitative features in Figure 4. Stars of all metallicity bins host warm super-Earths and warm sub-Neptunes. Cool Jupiters are intrinsically rare at all metallicities, but they are present in all four metal- licity bins. We observe a steady increase of hot Jupiters with increasing metallicity. Sub-Saturns of all orbital periods are almost completely absent in the lower two metallicity bins, which represent half of the parent sam- ple S; they are almost exclusively found around high metallicity stars. While there are some examples of hot super-Earths in each metallicity bin, their numbers in- crease with increasing metallicity. Finally, there is al- most a complete absence of hot sub-Neptunes in the low- est metallicity bins. Hot sub-Neptunes are more com- mon with increasing metallicity. In order to quantitatively assess the extent to which metallicity enhances the production of different types of planets, we compare the distribution of planet host metallicities to that of the field star population. In Ta- ble 5, we list the mean metallicities of the various planet subclasses as well as the standard error of the mean (SEM), which we compare to the mean field star metal- licity, as measured from LAMOST spectra. We assess the significance of the difference between field star and planet host star metallicities using the stu- dent t-test, which evaluates the difference between the means of the two samples in units of SEM, the t-statistic. The t-test also returns a p-value which is the probabil- ity that field stars and planet host stars are drawn from distributions with the same mean value. We may observe statistically significant differences in mean metallicity for two reasons: (1) intrinsic differ- ences between planet host and field star metallicities or (2) residual offsets in the CKS and LAMOST metallic- ity scales. While we calibrated LAMOST metallicities to the CKS scale, we estimate that offsets of 0.01 dex may remain (see Section 2.4 and Appendix A). To account for the possibility of such residual offsets, we perform three t-tests for each sample where we shift the LAMOST metallicities by −0.01 dex, 0.00 dex, and +0.01 dex. Each of these different tests returns a different p-value. We use the largest (most conservative) p-value to assess differences in mean metallicities. If the largest p-value is less than 0.01, we deem the metallicity difference to be significant. Stars hosting Jupiter-size planets have enhanced metallicities, (cid:104)[Fe/H](cid:105) = +0.12 ± 0.04 dex. The hot Jupiters hosts, with mean metallicity of (cid:104)[Fe/H](cid:105) = +0.19 ± 0.04 dex, are significantly enhanced relative to field stars (p-value < 8 × 10−4). While the mean metal- 10 Figure 3. Panel (a) shows the sizes and host star metallicities for the 970 planets in the filtered CKS planet catalog P. We observe a clear deficit of planets larger than Neptune with sub-Solar metallicities. The Fulton et al. (2017) radius gap may be seen from RP = 1.5–2.0 R⊕. We show the mean host star metallicity for various size ranges of planets with the red lines. The vertical bars show the standard error of the mean. The purple lines show the 25% and 75% quantiles. Mean metallicity is roughly constant from 0.7 R⊕ to 2.0 R⊕, rises from 2.0 R⊕ to 4.0 R⊕, and is roughly constant from 4 R⊕ to 16 R⊕. Panel (b): same as (a) except showing orbital period on the x-axis. We observe a small 0.05 dex increase in mean metallicity with decreasing orbital period over P = 1–10 days. licities for the warm and cool Jupiter hosts are also en- hanced, the small numbers of such planets prevent a high significance detection of a metallicity enhancement. Of all the planet size classes studied, the sub-Saturn hosts have the highest mean metallicity, (cid:104)[Fe/H](cid:105) = +0.16 ± 0.02 dex. Among the sub-Saturns, the hot sub-Saturns have the highest mean host star metallic- ity, (cid:104)[Fe/H](cid:105) = +0.26 ± 0.04 dex. We find that the hot and warm sub-Saturns hosts were significantly enhanced compared to field stars, while small numbers of detected cool sub-Saturns prevented a detailed comparison. As a whole, the sub-Neptunes have a mean metallicity of (cid:104)[Fe/H](cid:105) = +0.05 ± 0.01 dex, close to the field star value. However, when split according to orbital period, we find that the hot sub-Neptunes have an enhanced mean metallicity, (cid:104)[Fe/H](cid:105) = +0.11 ± 0.02 dex. The large mean metallicity, combined with the large num- ber of such planets (N = 108), results in a very sig- nificant detection of a metallicity enhancement (p-value < 4 × 10−9). Due to their high detectability and high intrinsic occurrence, the warm sub-Neptunes have the largest total number of any P –RP subclass, N = 282. The large number of planets means that even small offsets in mean metallicities can be significant. While the mean metallicity of warm sub-Neptunes, (cid:104)[Fe/H](cid:105) = +0.04±0.01 dex), was not as high as that of the hot sub- Neptunes, it was significantly elevated relative to field stars (p-value < 2 × 10−3). Finally, the super-Earth-size planets have the lowest mean metallicity of +0.01 ± 0.01 dex, which is consis- tent with field star metallicity. However, as with the sub-Neptunes, the hot super-Earth hosts exhibit en- hanced metallicity of +0.05 ± 0.01 dex. The difference in mean metallicity (in dex) is not as large for the sub- Neptunes, so while the offset is still significant (p-value < 2 × 10−4), it is not as significant as for the hot sub- Neptunes. Despite the large number of warm super- Earths, we cannot detect a significant offset in mean metallicity. One may wonder whether the mean metallicities of stars hosting warm super-Earths/sub-Neptunes could possibly be different from the mean metallicity of the Kepler field (−0.01 dex), given the high intrinsic occur- rence of these planets. While numerous previous works have shown that warm super-Earths and sub-Neptunes are intrinsically common (see, e.g., Petigura et al. 2013), neither class of planets is found around 100% of stars. For example, in Section 5 we show that there are ≈17 warm super-Earths per 100 stars. Therefore, the mean metallicity of warm super-Earth hosts could be signif- icantly different than that of field stars. We consider the following limiting case: Suppose that the process that produces warm super-Earths has a step function dependence on metallicity. Among Kepler targets, 17% 0.50.71234571020Planet size (Earth-radii)−0.6−0.4−0.20.00.20.4[Fe/H]aMedian Uncert. 0.3131030100300Orbital Period (days)bMedian Uncert. 11 Figure 4. Orbital periods and radii of planets orbiting host stars belonging to different metallicity bins. Each metallicity bin captures an equal fraction (25%) of the parent stellar sample S (see Table 4). In panel (a), blue points represent planets orbiting host stars with [Fe/H] < −0.116 dex, the lowest metallicity bin. The gray points show the full planet sample P for context. At top right, we show fp, the fraction of planets belonging to this metallicity bin. Panels (b)–(d): same as (a) except for different metallicity bins. If planets formed with equal efficiency regardless of host star metallicity, each bin would have fp = 25% and the distribution of blue points would be indistinguishable from bin to bin. Planets around the lowest metallicity stars (a) are clearly confined to a restricted range of P –RP space compared to planets around the highest metallicity hosts (d), notably in the lower right envelope of longer periods and smaller sizes. 0.512481632Planet Size (Earth-radii) Median Uncert.[Fe/H] < −0.116fp=20%a[Fe/H] = (−0.116,+0.020)fp=22%b0.3131030100300Orbital Period (days)0.512481632Planet Size (Earth-radii)[Fe/H] = (+0.020,+0.131)fp=30%c0.3131030100300Orbital Period (days)[Fe/H] > +0.131fp=28%d 12 of stars have [Fe/H] > 0.16 dex. A universe where every star with [Fe/H] > 0.16 dex produced one warm super- Earth, and every star with [Fe/H] < 0.16 dex produced zero, would be consistent with the measured occurrence. By consulting the distribution of Kepler field star metal- licities (see Section 2.4), we find that the average metal- licity of all stars with [Fe/H] > 0.16 dex is 0.25 dex. In this limiting case, the average metallicity of warm super- Earth hosts would be (cid:104)[Fe/H](cid:105) = 0.25 dex, much higher than the observed value of −0.04±0.02 dex. That warm super-Earth and sub-Neptune hosts have mean metallic- ities nearly the same as that of field stars demonstrates qualitatively that their occurrence is not a steep function of metallicity. We show this quantitatively in Section 6. In summary, close-in (P < 10 days) planets of all sizes have enhanced metallicity host stars. Warm super- Earth-size planets at intermediate distances (P = 10– 100 days) orbit stars with a metallicity distribution sim- ilar to that of field stars. Larger planets in this pe- riod range have enhanced metallicities, and the sub- Neptunes and sub-Saturns have a significant metallic- ity enhancement. Finally, the number of detected cool planets (P = 100–350 days) of all sizes is too small to permit the detection of different host star metallicities. Table 5. Comparison of Planet Host and Field Star Metallicities Name Hot Jupiters Warm Jupiters Cool Jupiters Hot Sub-Saturns Warm Sub-Saturns Cool Sub-Saturns Hot Sub-Neptunes Warm Sub-Neptunes Cool Sub-Neptunes Hot Super-Earths Warm Super-Earths Cool Super-Earths All Jupiters All Sub-Saturns All Sub-Neptunes All Super-Earths P day 1–10 10–100 100–350 1–10 10–100 100–350 1–10 10–100 100–350 1–10 10–100 100–350 1–350 1–350 1–350 1–350 RP R⊕ 8.0–24.0 8.0–24.0 8.0–24.0 4.0–8.0 4.0–8.0 4.0–8.0 1.7–4.0 1.7–4.0 1.7–4.0 1.0–1.7 1.0–1.7 1.0–1.7 8.0–24.0 4.0–8.0 1.7–4.0 1.0–1.7 n(cid:63) (cid:104)[Fe/H](cid:105) dex t-stat p-value Siga 14 +0.19 ± 0.04 4 +0.06 ± 0.02 13 +0.06 ± 0.05 7 +0.26 ± 0.04 19 +0.14 ± 0.02 7 +0.11 ± 0.06 108 +0.11 ± 0.02 282 +0.04 ± 0.01 29 −0.05 ± 0.03 181 +0.05 ± 0.01 132 −0.04 ± 0.02 4.5 < 8 × 10−4 3.4 < 6 × 10−2 1.4 < 3 × 10−1 6.5 < 7 × 10−4 5.6 < 6 × 10−5 1.9 < 1 × 10−1 7.0 < 4 × 10−9 4.0 < 2 × 10−3 −1.3 < 3 × 10−1 4.7 < 2 × 10−4 −2.2 < 1 × 10−1 4 −0.36 ± 0.03 −11.4 < 2 × 10−3 4.2 < 6 × 10−4 31 +0.12 ± 0.03 7.2 < 1 × 10−7 33 +0.16 ± 0.02 419 +0.05 ± 0.01 6.2 < 7 × 10−7 1.3 < 8 × 10−1 317 +0.01 ± 0.01 Y N N Y Y N Y Y N Y N Y Y Y Y N Note-We inspected the metallicity distribution of various groups of planets defined by their sizes and orbital periods. For each group, we computed the mean metallicity (cid:104)[Fe/H](cid:105) and the standard error of the mean. We compare these metallicities to the metallicities of Kepler field stars, as measured by LAMOST, using the student t-test. This test returns a t-statistic, which quantifies the statistical difference in mean metallicities, and a p-value, the probability that the two samples were drawn from distributions with the same mean. aIs the difference between planet host and field star metallicities significant (i.e., is p-value < 10−2)? 4. PLANET OCCURRENCE: METHODOLOGY In Section 3, we examined the number distribution of planets belonging to host stars of differing metallici- ties. We now address the underlying prevalence of plan- ets. In this section, we our methodology for computing planet occurrence and fitting parameterized descriptions of planet occurrence to the observed data. Section 5 presents planet occurrence as a function of orbital pe- riod and planet size. Section 6 presents planet occur- rence as a function of orbital period, planet size, and stellar metallicity. 4.1. Definitions Adopting the notation of Youdin (2011), the proba- bility that a star with properties zzz has a planet with (cid:90) X (cid:90) (cid:88) j X properties xxx that lie in a volume of dxxx of phase space is df = ∂f (xxx, zzz) ∂xxx dxxx. (1) In this paper, zzz is metallicity M = [Fe/H]6 and xxx is some combination of log P and log RP .7 As a matter of convenience we express differential distribution as ∂f (xxx, zzz) ∂xxx = Cg(xxx, zzz), (2) where C is a normalization constant and g is a shape function that may depend on stellar and/or planetary properties. The number of planets per star (NPPS) having stellar properties zzz over a specified volume of planet properties X is f (zzz) = Cg(xxx, zzz)dxxx (3) The NPPS having stellar properties within a range of stellar properties Z is f = 1 n(cid:63) Cg(xxx, zzzj)dxxx (4) where j labels stars where zj ∈ Z. Note that our definition of planet occurrence is NPPS. As a matter of convenience, we often express this in units of planets per 100 stars. Some papers (e.g. Fischer & Valenti 2005) define planet occurrence as the frac- tion of stars with planets (FSWP). For a transit survey like Kepler, computing FSWP is challenging because one must compute the number of stars without plan- ets. However, a transit survey cannot distinguish be- tween stars without planets and stars without transiting planets. Converting between NPPS and FSWP requires detailed modeling of planet multiplicity and mutual in- clinations and is beyond the scope of this work. For more discussion on these points, see Youdin (2011) and references therein. 4.2. Occurrence within a Cell In this paper, we compute and display planet occur- rence over rectilinear "cells" constructed from various combinations of log P , log RP , and M. We express the size of these cells as ∆ log P , ∆ log RP , ∆M, measured in dex. The occurrence within a cell is fcell, which de- pends on the number of independent trials ntrial that could have yielded a detected planet. In the limit where every star in S has one transiting planet that is also de- tectable by the Kepler pipeline, ntrial = n(cid:63). In practice, n(cid:63)(cid:88) 13 (6) ntrial must account for (1) non-transiting orbital incli- nations and (2) lack of detectability due to insufficient S/N. For a given planet size and orbital period, ntrial = ptr,i(P ) pdet,i(P, RP ) (5) i=1 where i labels each star in S, ptr,i is the transit probabil- ity, and pdet,i is the probability that a transiting planet would be detectable by the Kepler pipeline. We repre- sent this more compactly as ntrial = n(cid:63) (cid:104)ptr,i pdet,i(cid:105) , where (cid:104)·(cid:105) denotes the arithmetic mean. Following Bowler et al. (2015), if we assume that planet occurrence is log-uniform within a cell, ntrial for a cell of size ∆xxx = ∆ log P × ∆ log RP is (cid:104)ptrpdet(cid:105) dxxx. ntrial = (cid:90) (7) n(cid:63) ∆xxx For a given cell with ntrial trials, npl detected planets, and nnd = ntrial − npl non-detections, the likelihood of fcell is given by the binomial distribution P (fcellnpl, ntrial) = P (nplfcell, ntrial) cell (1 − fcell)nnd = Cf npl where C is the following normalization constant:8 C = (ntrial + 1)Γ(ntrial + 1) Γ(npl + 1)Γ(nnd + 1) . (8) (9) (10) If no transits are detected in a given cell, we may place an upper bound on the occurrence rate by calculating the maximum value of fcell that yields zero planet de- tections (npl = 0) with 90% probability. This is done by numerically finding fcell that satisfies P (fnpl, 0)df = 90%. (11) 0 When analyzing planet occurrence as a function of metallicity, we must consider the fraction of stars in S that falls within a specified range of metallicity. There- fore, to compute fcell for a cell bounded by [RP,1, RP,2], [P1, P2], and [M1, M2] we simply multiply ntrial from Equation 7 by the fraction of LAMOST stars with metallicities between M1 and M2. This treatment as- sumes that planet detectability does not depend on metallicity. We justify this assumption in Appendix B. 4.3. Completeness To compute occurrence, we must estimate (cid:104)ptrpdet(cid:105). The probability that a randomly inclined planet at a dis- (cid:90) fcell 6 We use M interchangeably with [Fe/H] for compactness. 7 In this paper, log only refers to log10. Natural logs are always expressed as ln. 8 Here, the factorials have been replaced with gamma functions via Γ(x + 1) = x! and (n + 1) is a normalization constant. 14 tance of a will transit with b < 0.9 is ptr = 0.9R(cid:63)/a, as- suming circular orbits. We compute ptr for each star us- ing the tabulated values of R(cid:63) and M(cid:63) from the Q1–Q17 (DR25) stellar properties table (Mathur et al. 2017), along with Kepler's Third Law. The probability of detecting a transiting planet with size RP and period P depends on number of factors, the most significant of which is the S/N. We follow the methodology of Fulton et al. (2017) and compute an expected S/N using the following formula: (cid:19)2(cid:114) (cid:18) RP R(cid:63) Tobs P (cid:18) 1 (cid:19) σ(T14) S/N = (12) Here, Tobs is the time the star was observed by Ke- pler and σ(T14) represents the photometric variability on transit timescales. The DR25 stellar properties ta- ble lists Tobs, R(cid:63), and photometric noise9 on 3, 6, and 12 hr timescales. We compute σ(T14) at intermediate values of T14 through piecewise linear interpolation (or extrapolation) in log σ–log T14 space. Characterizing pdet as a function of S/N has been ad- dressed several previous works (e.g. Fressin et al. 2013; Petigura et al. 2013; Christiansen et al. 2015). This is a challenging problem, which depends on how a com- plex, multi-stage transit pipeline performs in the face of correlated, non-stationary, and non-Gaussian photo- metric noise. We adopted the pdet (S/N) of Fulton et al. (2017), who used the transit injections of Christiansen et al. (2015) to derive the following relationship: pdet(s) = Γ(k) tk−1e−tdt (13) 0 where s is the S/N, k = 17.56, l = 1.00, and θ = 0.49. We illustrate (cid:104)pdet(cid:105) and (cid:104)ptrpdet(cid:105) as a function of P and RP in Figure 5. 4.4. Fitting the Occurrence Distribution We often wish to characterize the occurrence distri- bution with parametric models. These models serve to quantify shapes and trends in the population of planets. We extend the maximum-likelihood method of Howard et al. (2012) to fit Cg(xxx, zzz; θθθ) to the observed planet population, where θθθ represents the vector of shape pa- rameters in our model. The occurrence within each cell fcell gives an estimate (cid:90) s−l θ of the differential occurrence rate via fcell ∆xxx Cg(xxx, zzz; θθθ) = (14) The log-likelihood of a given model with parameters {C, θθθ} given the observed rate in a single cell labeled ncell(cid:88) by i is ln Li = npl,i ln Cg∆xxx + nnd,i ln(1 − Cg∆xxx). (15) Each cell is an independent constraint on Cg(xxx, zzz; θθθ), so a fit to multiple cells requires maximizing the combined log-likelihood over all cells: ln L = Li. (16) i=1 When fitting parameterized models of occurrence in Sections 5 and 6, we use very small bins having ∆ log P = 0.05 dex and ∆ log M = 0.05 dex. As a result, many bins have one or zero planet detections. The maximum- likelihood method is stable for small bin sizes because Equation 15 incorporates non-detections. We experi- mented with even finer bins, but did not observe sig- nificant changes in the results. After finding the max- likelihood model, we explore the credible range of models using Markov Chain Monte Carlo (MCMC) sampling.10 5. PLANET OCCURRENCE: PERIOD AND RADIUS We first present planet occurrence on the P –RP plane, averaging over stellar metallicity. Following Howard et al. (2012), we divided the domain of orbital period and planet size into numerous cells spanning ∆ log P × ∆ log RP = 0.25 dex × 0.15 dex. For each cell, we computed fcell according to the prescription from Sec- tion 4.2. We display fcell as a checkerboard in Figure 6, where each cell in is shaded and annotated according to the occurrence of planets within the cell. We list the cell-by-cell rates, uncertainties, and upper limits in Ta- ble C1 in the Appendix. We also show occurrence as a function of orbital pe- riod for various size classes in Figure 7, computed over bins spanning ∆ log P = 0.25 dex. Errorbars and up- per limits are computed according to the methodology presented in Section 4.2. For super-Earths and sub- Neptunes, inspection of the binned occurrence rates re- veals that df /d log P increases with P until a transi- tion period P0, above which occurrence is nearly uni- form in log P . This "knee" points toward an important orbital distance in the formation or subsequent migra- tion of planets. We characterized the period distribution with the following parameterization from Howard et al. (2012): = CP β(cid:16) 1 − e−(P/P0)γ(cid:17) . (17) df d log P Far from P0, this function reverts to a power law with 9 Combined Differential Photometric Precision (CDPP, Jenkins et al. 2010) 10 We used the affine invariant sampler of Goodman & Weare (2010) as implemented in Python by Foreman-Mackey et al. (2013). 15 Figure 5. Average detectability of planets as a function of planet size and orbital period. Panel (a): the probability that a transiting planet would be detected averaged overall stars in our sample, (cid:104)pdet(cid:105). Panel (b): the product of the transit and detection probability averaged overall stars in the sample (cid:104)pdetptr(cid:105). The effective number of stars from which the planet detections are drawn is simply ntrial = n(cid:63) (cid:104)pdetptr(cid:105). the following indices: P α P β df d log P ∝ if P (cid:28) P0, where α = γ + β if P (cid:29) P0 (18) super-Earths We fit this distribution using the methodology from Sec- tion 4.2 and list the associated parameters in Table 6. Our best-fit model and 1σ range of credible models are shown in Figure 7. As explained in Section 4.4, the max-likelihood fitting uses much finer bins than those displayed in Figure 7. The binned rates shown in Fig- ure 7 serve to guide the eye and are not used in the fitting. For is 6.5+1.6−1.2 days. At shorter orbital periods, occurrence rises rapidly with increasing P , with df ∝ P αd log P , where α = 2.4+0.4−0.3. At longer orbital periods, df ∝ P βd log P , where β = −0.3+0.2−0.2. The transition period for sub- Neptunes is farther out at P0 = 11.9+1.7−1.5 days, but the power law indices are similar, with α = 2.3+0.2−0.2 and β = −0.1+0.1−0.1. The distributions of sub-Saturns and Jupiters are not well-described by the power law cut- off model. Their occurrence gradually increases over P = 1–300 days. transition period P0 the Figures 6 and 7 are convenient tools for calculating planet occurrence over various domains of period and radius. However, several features in the planet popu- lation are hard to see in these plots due to the coarse bin sizes and the arbitrary location of the bin bound- aries. Figure 8 shows a finer view of planet occurrence. We computed occurrence in bins of period and radius spanning 0.25 dex and 0.10 dex respectively. We then shifted the bins in small steps of ∆ log P and ∆ log RP to smoothly trace out planet occurrence in the P –RP plane. The shading in Figure 8 gives a bird's eye view of the prevalence of various types of planets. Panels (a) and (b) of Figure 8 show occurrence on a linear scale to highlight the most abundant types of planets (super-Earths and sub-Neptunes). We note the rapid increase in occurrence for RP (cid:46) 4 R⊕, which has been noted by Howard et al. (2012) and numerous subse- quent works. Recently, Fulton et al. (2017) re-examined the radius distribution of small planets using the CKS catalog. Fulton et al. (2017) also noted a gap in the radius distribution at RP = 1.7 R⊕. This gap was pre- dicted by various groups who modeled the erosion ero- sion of envelopes (e.g. Lopez & Fortney 2013; Owen & Wu 2013; Jin et al. 2014; Chen & Rogers 2016). We also resolve the gap that separates the super-Earth and sup-Neptune populations. Panels (c) and (d) of Figure 8 show occurrence on a logarithmic scale to highlight domains of low planet occurrence. Although hot Jupiters are rare, they con- stitute a distinct island in the P –RP plane, surrounded by a sea of still lower occurrence. Hot planets of in- termediate sizes are very rare. This triangular "Hot Planet Desert" has been noted by Mazeh et al. (2016) and Dong et al. (2017), and is thought to be due to photo-evaporative envelope stripping. We also observe an factor of ≈5–10 increase in occur- rence of Jupiters from P = 100–300 days. Cumming et al. (2008) analyzed the planets detected by radial ve- locities from the Keck Planet Search and observed that giant planet occurrence increases by a factor ≈5 from P = 100–300 days. This rise in the occurrence of Jovian 131030100300Orbital Period (days)0.5124816Planet Size (Earth-radii)a131030100300Orbital Period (days)0.5124816Planet Size (Earth-radii)b0.00.20.40.60.81.0⟨pdet⟩0.0000.0010.0100.0200.0500.100⟨ptrpdet⟩ 16 planets between 100–300 days, observed in both the Ke- pler stellar population and among nearby stars, may be associated with an ice-line at ∼1 au. The joint distribution of planets in the P –RP plane shown in Figure 8 is an important quantitative descrip- tion of the population of planets. For example, this dis- tribution is useful for studies predicting planet yields in future surveys (e.g. Sullivan et al. 2015). To facilitate such studies, we sample this distribution and provide the periods and radii for a representative population of planets in a sample of 100,000 Sun-like stars in Table C2. This synthetic population also offers an convenient way to compute integrated planet occurrence (but not un- certainties) over arbitrary bins of P and RP . Table 6. Best-fit Parameters for Planet Period Distributions Size Class Super-Earth RP 1.0–1.7 Sub-Neptune 1.7–4.0 [Fe/H] all < 0 > 0 all < 0 > 0 β log10 C −0.32+0.32−0.25 −0.3+0.2−0.2 −0.11+0.51−0.39 −0.4+0.3−0.4 −0.35+0.32−0.23 −0.4+0.2−0.3 −0.28+0.18−0.18 −0.1+0.1−0.1 −0.75+0.27−0.22 +0.1+0.1−0.2 +0.09+0.29−0.24 −0.3+0.1−0.2 P0 γ 6.5+1.6−1.2 8.8+2.3−1.8 5.1+1.4−0.8 11.9+1.7−1.5 10.6+2.5−1.8 13.2+2.8−2.1 2.7+0.3−0.2 3.0+0.4−0.3 2.9+0.4−0.3 2.4+0.2−0.2 2.7+0.4−0.4 2.5+0.2−0.2 Note-Best-fit parameters associated with the occurrence model defined in Equa- tion 17. 6. PLANET OCCURRENCE: PERIOD, RADIUS, AND STELLAR METALLICITY Here, we analyze planet occurrence as a function of period, radius, and stellar metallicity. As a first step, we divided the super-Earth and sub-Neptune populations into two groups, depending on whether their host stars had sub- or super-solar metallicities. We then repeated the analysis from Section 5, modeling the occurrence distribution of both groups according to Equation 17. Figure 9 shows the occurrence of super-Earths and sub-Neptunes as a function of orbital period for sub- and super-solar metallicity hosts. The occurrence rates for P > 10 days are generally consistent for the two metallicity bins. For P < 10 days, the occur- rence rates of super-Earths/sub-Neptunes is ≈2–3 times higher for super-solar metallicity hosts compared to sub- solar metallicity hosts. To quantify the metallicity correlation for our different planet classes, we modeled occurrence using the follow- ing parametric function: df d log P dM = CP α10βM . (19) This model extends the following exponential model, df dM = C10βM , (20) which was used in Fischer & Valenti (2005) and in sev- eral subsequent works. Given that M = log(nFe/nH) − log(nFe/nH)(cid:12), Equation 20 is equivalent to a power law relationship between planet occurrence and the number density of iron atoms in a star's photosphere relative to hydrogen, (cid:20) nFe/nH (cid:21)β . (21) df dM = C (nFe/nH)(cid:12) We fit the distribution defined in Equation 19 using the methodology from Section 4.4 and list the associated model parameters in Table 7. For the hot super-Earths we found β = +0.6+0.2−0.2, indi- cating a significant positive metallicity correlation. That β ∼ 1 for hot super-Earths means that their occurrence is nearly proportional to the number of iron atoms in a star's photosphere, relative to hydrogen. Through our MCMC modeling, we found that the metallicity index β was only weakly correlated with the period index α. In contrast, α and the normalization constant C were highly covariant. The metallicity correlation steepens for larger plan- ets. For hot sub-Neptunes, β = +1.6+0.3−0.3. For hot sub-Saturns and hot Jupiters, our planet sample P con- tains only 7 and 14 planets, respectively. Small sample size leads to larger uncertainties on β, and for hot sub- Saturns and Jupiters, β = +5.5+1.6−1.5 and β = +3.4+0.9−0.8, respectively. Despite these larger uncertainties, it is clear that hot sub-Saturns and Jupiters have signifi- cantly steeper metallicity dependencies than hot super- Earths and sub-Neptunes. Our occurrence rate model depends on both P and 17 Figure 6. The domain of orbital period and planet size, which has been sub-divided into numerous sub-domains, "cells." Each cell is annotated with the average number of planets per 100 stars having properties within each cell. This quantity is also reflected through the shading of each cell. The number of planets has been corrected for the probability of transiting and for detection completeness. M, and thus cannot be displayed on a 2D plot. For dis- play purposes, we performed an integration from P = 1– 10 days. Figure 10 shows the number of planets belong- ing to different size classes per 100 stars having P = 1– 10 days for various metallicity intervals spanning ∆M = 0.2 dex. As in Figure 7, the binned rates serve to guide the eye and are not used in the fitting. We observe different metallicity dependencies for more distant planets. As shown in Figure 10, warm super- Earths are consistent with no metallicity dependence β = −0.3+0.2−0.2. Warm sub-Neptunes show a posi- tive metallicity correlation, but their power law index β = +0.5+0.2−0.2 is significantly shallower than that of the hot sub-Neptunes. The warm sub-Saturns have a posi- tive metallicity correlation of β = +2.1+0.7−0.7, again more shallow than that of the hot sub-Saturns. Finally, our sample contains too few warm Jupiters (4) to search for a metallicity correlation. In summary, some but not all planet subclasses are as- sociated with stellar metallicity. Warm super-Earths ex- hibit no metallicity dependence. The planet-metallicity correlation steepens with increasing size and decreas- ing orbital period, with the hot Jupiters and hot sub- Saturns showing the steepest metallicity dependencies. We discuss some physical interpretation for these trends in the following section. 131030100300Orbital Period (days)0.512481632Planet Size (Earth-radii)0.010.030.10.31310Planets per 100 Stars per Bin 18 Figure 7. Planet occurrence as a function of orbital period for various size classes. For example, green points show the number of super-Earths per 100 stars per 0.25 dex interval in period. Downward arrows represent upper limits (90%). For the super-Earths and sub-Neptunes, we fit the occurrence using the power law and exponential cutoff model described in Section 5 (Equation 17). The solid lines and bands show the best-fitting model and 1σ range of credible models, respectively. For super-Earths and sub- Neptunes at P < 10 days, planet occurrence increases like df /d log P ∝ P α, where α ≈ 2.4+0.4−0.3 and α ≈ 2.3+0.2−0.2, respectively. At longer orbital periods, occurrence is nearly uniform in log P . A transition period P0 characterizes where the distributions changes slope, which occurs at P0 = 6.5+1.6−1.2 days and P0 = 11.9+1.7−1.5 days, respectively. The distributions of sub-Saturns and Jupiters are not well-described by this model. Their occurrence gradually increases over P = 1–300 days. Table 7. Best-fit Parameters for Planet-Metallicity Distribu- tions α β log10 C −2.22+0.09−0.09 +1.7+0.1−0.1 +0.6+0.2−0.2 Hot Super-Earth Warm Super-Earth +0.05+0.22−0.23 −0.6+0.2−0.2 −0.3+0.2−0.2 −2.95+0.15−0.16 +2.2+0.2−0.2 +1.6+0.3−0.3 Hot Sub-Neptune Warm Sub-Neptune −0.85+0.14−0.13 +0.2+0.1−0.1 +0.5+0.2−0.2 −4.93+0.65−0.76 +2.2+0.8−0.7 +5.5+1.6−1.5 Hot Sub-Saturn −2.59+0.53−0.53 +0.5+0.3−0.4 +2.1+0.7−0.7 Warm Sub-Saturn −3.32+0.29−0.31 +0.9+0.4−0.4 +3.4+0.9−0.8 Hot Jupiter ··· Warm Jupiter ··· ··· Note-Best-fit parameters associated the metallicity occurrence model defined in Equation 19. joint period- 7. SUMMARY AND DISCUSSION 1 year. While numerous previous works have addressed The Kepler survey has revealed many valuable in- sights regarding the prevalence of planets with P (cid:46) 131030100300Period (days)0.010.030.10.3131030Planets per 100 Stars per 0.25 dex P IntervalSuper-EarthsSub-NeptunesSub-SaturnsJupiters 19 Figure 8. Panel (a) shows the planet sample P and planet occurrence in the P –RP plane. The shading at each (P ,RP ) point represents the number of planets per 100 stars within an interval centered at (P ,RP ) that spans 0.25 dex in log P and 0.10 dex in log RP . The size of the interval is indicated at top right. For example, the darkest contour indicates that there are ≈4 planets per 100 Sun-like stars having periods within 0.125 dex of 40 days and radii within 0.05 dex of 2.5 R⊕. Panel (b): same as (a), but without detected planets for clarity. Warm sub-Neptunes are the most abundant class of planets, and warm super-Earths are another common class of planets. The super-Earths and sub-Neptunes are separated by a diagonal gap of low planet occurrence, described by Fulton et al. (2017). Panels (c)–(d): same as (a)–(b), but with logarithmic shading to highlight domains of low occurrence. The hot Jupiter population is an "island," distinct from the rest of the planet population. There is a "desert" of hot planets having intermediate sizes. Kepler planet occurrence in the P –RP plane,11 we have leveraged the high-precision, high purity CKS catalog to produce a clearer picture of planet occurrence as a function of orbital period and size (Section 5; Figures 6– 8). This complements the work of Fulton et al. (2017) that focused on super-Earth and sub-Neptune occur- 11 A non-exhaustive list includes: Youdin (2011); Catanzarite & Shao (2011); Traub (2012); Howard et al. (2012); Fressin et al. (2013); Dong & Zhu (2013); Petigura et al. (2013); Foreman- Mackey et al. (2014); Burke et al. (2015); Dressing & Charbonneau (2015); Mulders et al. (2015). rence rates with P = 1–100 days. Our occurrence rates may be easily incorporated into yield calculations for future surveys sensitive to planets with P (cid:46) 1 yr. We found a hot Jupiter occurrence rate of 0.57+0.14−0.12%, which we compare to results from RV surveys in Section 7.3. While hot Jupiters are rare, we find that they occupy a distinct island in P –RP space, surrounded by a sea of still lower occurrence. Using the CKS catalog combined with LAMOST metallicities for a large sample of Kepler field stars, we have computed planet occurrence as a function of host star metallicity (Section 6; Figures 9 and 10). We ob- 1310301003000.512481632Planet Size (Earth-radii)Low CompletenessP−RP IntervalΔlogP = 0.25 dexΔlogRP = 0.10 dexa1310301003000.512481632Low CompletenessWarm Sub-NeptunesWarm Super-EarthsRadius GapCool Jupitersb131030100300Orbital Period (days)0.512481632Planet Size (Earth-radii)Low Completenessc131030100300Orbital Period (days)0.512481632Low CompletenessHot JupitersHot Planet Desertd012345Planets per 100 Stars per P−RP interval0.010.030.10.31310Planets per 100 Stars per P−RP interval 20 Figure 9. Panel (a) is analogous to Figure 7, except showing the period distribution of super-Earths at different host star metallicities. For example, red points show the number of super-Earths per 100 stars with super-solar metallicity per 0.25 dex period interval. Hot super-Earths are more common around metal-rich stars. Panel (b) same as (a) except for sub-Neptunes. Hot sub-Neptunes are also more common around metal-rich stars. serve a clear association between stellar metallicity and the prevalence of certain types of planets. Figure 11 summarizes our findings: stellar metallicity is associ- ated with the occurrence of hot planets (P < 10 days) and with the occurrence of large planets (RP > 1.7 R⊕, P < 100 days). While the occurrence of some types of planets cor- relate strongly with metallicity, others exhibit only a moderate or negligible correlation. The interpretation of these metallicity trends is not yet clear. There are many ways in which differences in metallicity might alter the processes of planet formation and evolution: through differences in the surface density of solids in the proto- planetary disk, the growth rate and abundance of solid cores, the availability of gas, the efficiency of gas ac- cretion onto protoplanets, or the rate of disk migration. There may also be others. We offer some interpretations of these trends in Sections 7.1–7.3. Finally, we consider future studies that could extend our understanding of the connection between planets and host star metallic- ity in Section 7.4. 7.1. Metallicity and the Prevalence of Large Planets We found that metallicity is associated with the preva- lence of large planets. This trend is consistent with the planet-metallicity correlation for Jovian-mass plan- ets found by RV surveys (Santos et al. 2004; Fischer & Valenti 2005) and with previous analyses of the metallic- ities of Kepler planet hosts (e.g. Buchhave et al. 2012). We observed that for smaller planets, the metallicity correlation weakens, which is consistent with trends ob- served in previous RV studies (e.g. Sousa et al. 2008; Ghezzi et al. 2010) and metallicity studies of Kepler planet hosts (e.g. Buchhave et al. 2012, 2014). Our analysis builds upon these previous studies by provid- ing planet occurrence as a function of metallicity based on the high-precision, high purity CKS catalog of planet radii and metallicities. There are roughly 15 warm super-Earths per 100 stars over the metallicities ranging from −0.4 dex to +0.4 dex. If we assume that stellar metallicity traces disk solid surface densities from 0.10–0.40 au (which correspond P ≈ 10–100 day orbits), then warm super-Earths can form with moderate efficiency, even in disks with low solid surface densities. Moreover, a factor of 6 increase in solids does not enhance the likelihood of a star to host a warm super-Earth. RV and TTV mass measurements of super-Earths typ- ically find masses of 1–10 M⊕ (Marcy et al. 2014, see Figure 7 of Sinukoff et al. 2017 for an updated census). The bulk densities of these planets are usually consistent with combinations of rock and iron, but are inconsistent with even small envelope fractions of Menv/MP ∼ 1%, which would significantly reduce the observed bulk den- sities. Even though sub-Neptunes are significantly larger than super-Earths, they share several important char- acteristics. Sub-Neptunes are typically 5–15 M⊕, which 131030100300Period (days)0.030.10.3131030Planets per 100 Stars per 0.25 dex P Interval[Fe/H] < 0[Fe/H] > 0Super-Earthsa131030100300Period (days)0.030.10.3131030Planets per 100 Stars per 0.25 dex P Interval[Fe/H] < 0[Fe/H] > 0Sub-Neptunesb 21 Figure 10. Panel (a): Points show the number of planets per 100 stars for bins of host star metallicity spanning ∆M = 0.2 dex having P = 1–10 days. Triangles represent upper limits (90%). The colors correspond to different planet size classes. We modeled the observed occurrence rates with an exponential model, df ∝ P α10βM dM, where β characterizes the strength of a metallicity correlation (see Section 6, Equation (19)). The solid lines and bands show the best-fitting model and the 1σ credible range of models, respectively. Metallicity correlates with the occurrence of super-Earths (β = +0.6+0.2−0.2), sub-Neptunes (β = +1.6+0.3−0.3), sub-Saturns (β = +5.5+1.6−1.5) and Jupiters (β = +3.4+0.9−0.8). The strength of the correlation increases with planet size. Panel (b): same as (a) except for planets with P = 10–100 days. Warm super-Earths are not correlated with metallicity (β = −0.3+0.2−0.2). We observe positive correlations for larger planets. For warm sub-Neptunes (β = +0.5+0.2−0.2); for warm sub-Saturns (β = +2.1+0.7−0.7). Comparing the two panels, the metallicity correlation is stronger for P < 10 days. overlaps with the super-Earth mass range. Sub-Neptune bulk densities require H/He envelopes that are 1–10% of the planet's mass, which significantly enlarge the planets without significantly altering their masses. Therefore, the typical sub-Neptune contains a comparable amount of solid material as a typical super-Earth. Also, their overall occurrence rates are roughly comparable at 5– 10% per 0.25 dex period interval (Figure 9). We see an important difference between super-Earths and sub-Neptunes when we compare their dependencies on stellar metallicity. There are comparable numbers of warm super-Earths and warm sub-Neptunes around stars having [Fe/H] = [−0.4,−0.2] dex, about 20 per 100 stars. Unlike the warm super-Earths, the occurrence of warm sub-Neptunes increases with metallicity. There are about 40 warm sub-Neptunes per 100 stars having [Fe/H] = [+0.2,+0.4] dex. Perhaps the enriched disks of metal-rich stars form super-Earths more efficiently, but they easily acquire the few percent envelopes and are converted into sub- Neptunes. This could occur if super-Earths form more quickly in metal-rich disks, allowing for more time to accrete 1–10% gaseous envelopes. Such a timing argument was proposed by Daw- son et al. (2015), to explain an apparent absence of warm super-Earths around the highest metallicity stars ([Fe/H] > 0.18 dex) in the Buchhave et al. (2014) metal- −0.4−0.20.00.20.4[Fe/H]0.013000.030.10.3131030100Planets per 100 StarsSuper-EarthsSub-NeptunesSub-SaturnsJupitersaP = 1−10 days−0.4−0.20.00.20.4[Fe/H]0.013000.030.10.3131030100Planets per 100 StarsSuper-EarthsSub-NeptunesSub-SaturnsJupitersbP = 10−100 days 22 Figure 11. Summary of the planet-metallicity correlation across the P –RP plane. For each planet subclass studied, we display the average host star metallicity (cid:104)[Fe/H](cid:105) and the β index, which quantifies the strength of the planet-metallicity correlation. licity catalog.12 While only 12/132 (9%) of our warm super-Earths have [Fe/H] > 0.18 dex, this is consistent with the fact that only 12% of field stars have [Fe/H] > 0.18 dex. As shown in Figures 4 and 10, we do not observe such an absence of warm super-Earths, but in- stead find nearly equal occurrence of warm super-Earths around stars of wide-ranging metallicities. This un- derscores the importance of having accurate knowledge of the distribution of field star metallicities in planet- metallicity studies. Nevertheless, accelerated formation of cores in metal-rich disks may explain the rise in warm sub-Neptune occurrence with metallicity. An alternative to the timing argument, discussed above, is that high disk metallicities may somehow in- crease the rate of envelope accretion. However, the gas accretion models of Pollack et al. (1996) and (more re- cently) Lee & Chiang (2015) predict the opposite ef- fect: dusty envelopes should accrete more slowly, due to higher opacities and longer cooling timescales. As we consider larger warm planets, the metallicity correlation becomes stronger. Sub-Saturns have a β in- dex of +2.1+0.7−0.7. If disk metallicity assists with the ac- cretion of gas through accelerated core formation, it may also explain the sub-Saturns. However, abrupt strength- 12 The Dawson et al. (2015) definition of warm super-Earths was slightly different than ours: RP < 1.5 R⊕, P < 15 days. ening of the metallicity correlation above 4 R⊕ and the roughly order of magnitude decrease in the frequency of sub-Saturns compared to sub-Neptunes suggests the for- mation pathways of sub-Saturns and sub-Neptunes may be quite different. Around 20 sub-Saturns have well-measured masses from either RVs or TTVs (see Petigura et al. 2017a for a recent compilation). For sub-Saturns, we observe an order of magnitude scatter in the observed masses at a given size, indicating a diversity in core and envelope masses. While some sub-Saturns have ≈5 M⊕ cores, similar to the super-Earths and sub-Neptunes, many have cores of ≈50 M⊕. In addition, the most massive sub-Saturns tend to be found around the most metal- rich hosts (Petigura et al. 2017a). The existence of ≈50 M⊕ cores in planets with 20% envelope fractions challenges the classic core-accretion models of Pollack et al. (1996) that predict that cores larger than 10 M⊕ should undergo runaway accretion. Perhaps these massive sub-Saturn cores are the result of the late-stage mergers (or series of mergers) of 10 M⊕ cores. This formation scenario requires one or more closely spaced 10 M⊕ planets. As we have shown, the probability for a star to produce a 10 M⊕ core increases with metallicity. Therefore, the probability for a star to produce two 10 M⊕ cores likely increases with a steeper power law index. Thus, the production of sub-Saturns 110100Orbital Period (days)11.74.08.024.0Planet size (Earth-radii)Hot Super-Earths ⟨[Fe/H]⟩ = +0.05±0.01β=+0.6+0.2−0.2Hot Sub-Neptunes ⟨[Fe/H]⟩ = +0.11±0.02β=+1.6+0.3−0.3Hot Sub-Saturns ⟨[Fe/H]⟩ = +0.26±0.04β=+5.5+1.6−1.5Hot Jupiters ⟨[Fe/H]⟩ = +0.19±0.04β=+3.4+0.9−0.8Warm Super-Earths ⟨[Fe/H]⟩ = −0.04±0.02β=−0.3+0.2−0.2Warm Sub-Neptunes ⟨[Fe/H]⟩ = +0.04±0.01β=+0.5+0.2−0.2Warm Sub-Saturns ⟨[Fe/H]⟩ = +0.14±0.02β=+2.1+0.7−0.7Warm Jupiters ⟨[Fe/H]⟩ = +0.06±0.02β = Unconstrained by collisions may explain the mass-metallicity depen- dence and the steeper relationship between metallicity and planet occurrence. This could also explain the mass- metallicity correlation. This theory also predicts that if late-stage mergers play a large role in the formation of sub-Saturns, they should produce relic eccentricities that are observable at later times. Our planet sample includes only four warm Jupiters, which is insufficient to search for trends. Fischer & Valenti (2005) analyzed 1040 FGK-type stars from the Keck, Lick, and Anglo-Australian Telescope planet search programs and found a planet-metallicity correla- tion with a β = 2 index. However, this sample included planets with P = 1–4000 days, with the bulk of the sample having P > 300 days longer than the periods considered here. 7.2. Metallicity and the Prevalence of Short-period Planets We found that metallicity is associated with the pres- ence of short-period planets having P < 10 days. This is consistent with trends observed by Mulders et al. (2016) and Dong et al. (2017), who studied the host star metal- licities for samples of 665 and 295 planets, respectively. Both studies used spectroscopically constrained metal- licities from LAMOST. Our analysis incorporates the high-precision CKS metallicities for a larger sample of 970 planets. While close-in planets, of all sizes, are rare compared to more distant planets, their frequency increases as a function of stellar metallicity. In Section 5, when con- sidering just the period distribution of super-Earths and sub-Neptunes, we found that for P < 10 days, the occur- rence of super-Earths and sub-Neptunes declines with decreasing period. The slope of this falloff is approxi- mated by df ∝ P αd log P with α values of 2.4+0.4−0.3 and 2.3+0.2−0.2, respectively. In Section 6, when considering the joint P –[Fe/H] oc- currence distribution, we found that both hot super- Earths and sub-Neptunes are correlated with host star metallicity. This is especially noteworthy for the hot super-Earths, given that the warm super-Earths exhibit no such correlation. For P < 10 days, the frequency of hot super-Earths grows from 4% for stars having [Fe/H] = [−0.4,−0.2] dex to 10% for stars having [Fe/H] = [+0.2,+0.4] dex (see Figure 10). We see an even steeper increase in the hot sub-Neptune rates from 1% to 8% for the same metallicity intervals. In Sections 7.2.1 and 7.2.2, we consider two possible formation pathways: (1) in situ models, where planets form near their final lo- cation, and (2) high eccentricity migration, where the semi-major axes of fully formed planets are altered by major scattering events. In Section 7.2.3, we consider some observational tests that may distinguish between these scenarios. 7.2.1. In Situ Formation 23 The prevalence of compact multi-planet systems of super-Earths/sub-Neptunes discovered by Kepler helped inspire in situ planet formation models (see, e.g., Hansen & Murray 2013 and Chiang & Laughlin 2013). Chiang & Laughlin (2013) introduced the min- imum mass extrasolar nebula (MMESN) as a means of estimating the surface density profile Σ(a) of the proto- planetary disk. They took a sample of Kepler planets, estimated the mass in solids, and smeared those solids out over an area 2πa2. The MMESN assumes no large- scale migration of solid building blocks or planets. In the context of the MMESN framework, the falloff in super- Earth and sub-Neptune occurrence at P < 10 days re- flects a declining surface density profile. Lee & Chi- ang (2017) offered a physical explanation for declining density profiles. In their models, disks are truncated by interactions with the protostar magnetosphere at the co-rotation radius. The existence of hot sub-Neptunes presents a chal- lenge to strictly in situ models, because sub-Neptunes with P < 10 days are vulnerable to photo-evaporation. Owen & Wu (2017) performed a population synthesis starting with 3–10 M⊕ cores and a period distribu- tion of df ∝ P 1.9d log P , which is similar to the power law index presented in this work. They then gave the simulated planets a range of envelope fractions from 1 to 30% and tracked the radii of these planets as they were subjected to photo-evaporation. The Owen & Wu (2017) simulations produced very few sub-Neptunes with P < 10 days. In the Lee & Chiang (2017) model, hot super- Earth/sub-Neptune occurrence rates are set by stellar rotation on the pre-main sequence. It is not clear if stellar metallicity is correlated with pre-main-sequence rotation rates, and Lee & Chiang (2017) did not pre- dict a metallicity dependence for hot super-Earths/sub- Neptunes. Metallicity may still influence the vertical and radial profiles of the inner disk, even if there is no correla- tion with stellar rotation. The inner regions of high metallicity disks may have higher densities of free elec- trons, which could result in stronger coupling to stel- lar magnetic fields and change the accretion rate and density profile of the disk. The formation of hot super- Earths/sub-Neptunes may also involve significant radial transport of solids, either through the inward drift of "pebbles" that are no longer aerodynamically coupled to the gas disk (see, e.g., Lambrechts & Lega 2017, and references therein). The efficiency of this process is also governed by the radial and vertical density profiles. Future theoretical modeling of protoplanetary disks 24 is needed to investigate whether in situ models can re- produce the metallicity dependence observed here. In situ models tend to produce dynamically cool planetary systems, which can be corroborated by present-day ec- centricity and obliquity measurements 7.2.2. High Eccentricity Migration High eccentricity migration may explain the hot super-Earth/sub-Neptune period and metallicity distri- butions. Some of the hot super-Earths/sub-Neptunes may be the result of scattering events between multiple planets. Such encounters preserve the total orbital en- ergy of both planets, but the semi-major axes of both planets may be significantly altered. Dawson & Murray- Clay (2013) proposed a similar mechanism to explain the fact that high eccentricity giant planets with a = 0.1– 1.0 au, almost exclusively orbit stars with super-solar metallicities. A formation pathway involving scattering may help to explain the existence of hot sub-Neptunes, which are at risk of photo-evaporative stripping (see Section 7.2.1). Scattering events may occur late enough such that a sub-Neptune is not subjected to the majority its star's XUV output (t (cid:38) 100 Myr). For hot sub-Neptunes, the metallicity correlation is stronger (β = +1.6+0.3−0.3) than for the warm sub-Neptunes (β = +0.5+0.2−0.2). This steeper relationship may support a scattering interpretation given that the probability of scattering depends on the probability of a disk pro- ducing two closely spaced planets of comparable mass. If high eccentricity migration is active, the present-day systems planetary systems should be dynamically hot, provided that star–planet interactions are not strong enough to circularize or realign orbits. Again, this may be corroborated with eccentricity and obliquity mea- surements. 7.2.3. Observational Tests Additional clues to the formation of hot super- Earths/sub-Neptunes may lie in their orbital eccentrici- ties and degree of alignment with stellar spin axes (obliq- uities). If planets initially form in the plane of their disks, but are then scattered by other planets, we pre- dict relic eccentricities and obliquities. If these plan- ets form in relative isolation, and are simply a product of the local disk density profile, their eccentricities and obliquities should be low. Current state-of-the-art RV facilities have provided mass measurements of several dozen hot super-Earths and sub-Neptunes, but are not precise enough to mea- sure orbital eccentricities well. Obliquity measurements via the Rossiter–McLaughlin technique have been made for only a handful of sub-Jovian-size planets (e.g. HAT- P-11, Winn et al. 2010; Hirano et al. 2011). As the pre- cision and observing efficiency of RV facilities improve, we should be able to place tighter constraints on the orbits of these hot planets. 7.3. Hot Jupiters Hot Jupiters represent an extreme of planet formation, and many theories have been proposed to explain their formation. A non-exhaustive list of theories includes star–planet Kozai (e.g. Wu & Murray 2003), planet- planet Kozai (e.g. Naoz et al. 2011), smooth disk-driven Type-I migration (e.g. Ida & Lin 2008), high eccentric- ity migration, and in situ formation (e.g. Batygin et al. 2016). Hot Jupiters are also of interest because they of- fer a point of comparison between the Kepler population of planets and those around nearby stars. We find a hot Jupiter rate of 0.57+0.14−0.12%,13 which is about 1.5σ larger than 0.4 ± 0.1% reported by Howard et al. (2012). This difference cannot be due to differences the treatment of completeness; for hot Jupiters, pdet is nearly unity. The larger occurrence rate presented here is likely due to the improved radius measurements in the CKS sam- ple. Howard et al. (2012) included 13 hot planets with RP = 5.6–8.0 R⊕, slightly too small to be classified as a hot Jupiter by their criteria. With our improved CKS planet radii, we have found that these large, hot sub- Saturns are exceptionally rare. Our planet sample P includes just 3 such planets, while Howard et al. (2012) included 13, even though their stellar parent is less than twice as large as ours nS (58041 vs. 33020). Some of the hot sub-Saturns in Howard et al. (2012) were misidenti- fied hot Jupiters. A major challenge in comparing the hot Jupiter rates in the Kepler and the solar neighborhood is their low intrinsic occurrence. Even though our planet sample P was drawn from a parent stellar sample S containing 33020 stars, P contained just 14 detections due to low intrinsic occurrence and the requirement of transiting geometries. Even blind RV surveys which are not lim- ited to transiting geometries must observe hundreds of stars to detect a few hot Jupiters. Another challenge is that RV surveys are typically not magnitude-limited samples like S. RV surveys often prioritize stars that are single, slowly rotating, and have stellar properties that fall within a preselected range of interest (e.g. M-stars or evolved stars). Wright et al. (2012) attempted to retroactively remove these selection effect from the Cal- ifornia Planet Search database, to produce a magnitude- limited sample of stars that could be directly compared 13 The rate of 0.57+0.14−0.12% is slightly higher than a simple sum of the most likely rates from each hot Jupiter bin from Figure 6 due to the asymmetric uncertainties from low counts per bin. to the Kepler stellar sample. Wright et al. (2012) re- ported 10 hot Jupiters from a sample of 836 stars or a rate of 1.2 ± 0.4%, twice the rate presented here. How- ever, the large fractional uncertainty in the Wright et al. (2012) measurement means that these two results differ by only 1.5σ. In Section 6, we modeled the hot Jupiter occurrence as df ∝ P α10βM d log P dM and found a β index of +3.4+0.9−0.8. This strong association between hot Jupiters and metallicity has also been observed by RV stars. Fis- cher & Valenti (2005) studied the occurrence of Jupiter- mass planets in the CPS sample and modeled their oc- currence according to df ∝ 10βM and β index of 2.0. Johnson et al. (2010) performed a similar analysis, con- sidered an additional mass dependence df ∝ M α (cid:63) 10βM and found a β index of 1.2±0.2. It is worth recalling that both studies included long period planets (P > 1 yr), which constituted the bulk of the sample. Guo et al. (2017) repeated the Johnson et al. (2010) analysis, but restricted the study to hot Jupiters and found β = 2.1 ± 0.7. While this result differs with our measure- ment of β by about 2σ, it is clear that hot Jupiters in the solar neighborhood and in the Kepler field are both strongly associated with metallicity. In the previous sections, we have noted the general tendency for the metallically correlation to steepen with decreasing orbital period and increasing planet size, and have hypothesized some processes that could ac- count for these trends such as accelerated core/envelope growth, higher disk densities, and planet-planet scatter- ing. Given this general trend, it is perhaps not surprising that hot Jupiters would have the strongest metallicity correlation. However, given that the hot Jupiters oc- cupy a distinct island on the P –RP plane, it is possible that their formation pathway is radically different than that of the other planet classes considered here. 7.4. Future Survies Our target sample contains very few stars (16) with metallicities below −0.4 dex. As a result, we are insen- sitive to planet occurrence at low metallicities. Probing planet occurrence at low metallicities would shed light on some of the questions raised by this analysis, such as, What is the minimum metallicity needed to form a warm super-Earth? Since metal-poor stars are rare, a blind survey like Kepler must target a large number of stars in order to capture sufficient numbers of low metallicity stars. There is some room for improvement with the existing Kepler sample. Our sample was drawn from ≈38,000 of the ≈150,000 stars observed by Kepler due to our mag- nitude limit. Characterizing the metallicities of Kepler planet hosts and field stars fainter than Kp = 14.2 mag would enlarge the sample of metal-poor stars. How- 25 ever, given that these new stars would be between Kp = 14–16 mag, their amenability to the detection of small planets is poor. Upcoming missions like TESS (Ricker et al. 2015) and PLATO (Rauer et al. 2014) will survey more stars than Kepler and thus cast a significantly larger net for rare metal-poor stars. Gaia will also identify metal-poor stars across the sky. A targeted RV survey of low metal- licity stars would also probe planet occurrence at low metallicities. Such a survey would not be restricted to transiting planets, and would need to survey fewer stars to gather statistically significant planet samples. Finally, one could search for planets in globular clus- ters. Gilliland et al. (2000) conducted an important early transit search with HST, which targeted 47 Tu- canae ([Fe/H] = −0.78 dex). The expected yield was 17 hot Jupiters, but the survey yielded no detections, which was attributed to low metallicities. Recently, Ma- suda & Winn (2017) re-assessed the significance of the null result, using updated knowledge of the typical size, periods, and host star properties of hot Jupiters. Their revised yield calculation was 2± 1, and thus the null re- sult does not require a metallicity effect. However, based on the steep dependence of hot Jupiter occurrence with metallicity, yields would likely be low. Future globular cluster surveys with higher precision or longer baselines would help constrain planet population at −1 dex. How- ever, high stellar densities of globulars may complicate direct comparisons to field stars. 8. CONCLUSION We have measured planet occurrence as a func- tion of period, radius, and stellar metallicity within a magnitude-limited sample of Kepler target stars. We have leveraged precise planet radii and stellar metal- licities from the CKS catalog of Kepler planet-hosting stars, as well as the metallicity distribution of Kepler field stars measured by LAMOST. In summary, for P < 100 days the default planetary system contains either no planets detectable by Kepler (e.g. the solar system) or a system of one or more super-Earths/sub-Neptunes with P = 10–100 days. Stars with high metallicity are associated with some mechanism(s) that also allows for "misplaced" planets: sub-Saturns/Jupiters with P < 100 days and super-Earths/sub-Neptunes with P < 10 days. We have noted the general features of planet occurrence as a function of P , RP , and [Fe/H] and have offered some speculation regarding their physical ori- gins. Detailed population synthesis models that treat the disk profile, growth of cores, migration, dynamical instabilities, and photo-evaporation must produce these features. The CKS project was conceived, planned, and initi- of Sciences. 26 ated by A.W.H., G.W.M., J.A.J., H.T.I., and T.D.M. A.W.H., G.W.M., J.A.J. acquired Keck telescope time to conduct the magnitude-limited survey. We thank the many observers who contributed to the measurements reported here. We thank the referee, Lars Buchhave, for his detailed and thoughtful comments on the techniques and inter- pretations. We also thank Konstantin Batygin, Brendan Bowler, Ian Crossfield, and Eve Lee for enlightening con- versations that improved the final manuscript. Kepler was competitively selected as the tenth NASA Discovery mission. Funding for this mission is provided by the NASA Science Mission Directorate. We thank the Kepler Science Office, the Science Operations Center, the Threshold Crossing Event Review Team (TCERT), and the Follow-up Observations Program (FOP) Work- ing Group for their work on all steps in the planet dis- covery process, ranging from selecting target stars and pointing the Kepler telescope to developing and run- ning the photometric pipeline to curating and refining the catalogs of Kepler planets. The Guoshoujing Telescope (the Large Sky Area Multi-Object Fiber Spectroscopic Telescope, LAMOST) is a National Major Scientific Project built by the Chi- nese Academy of Sciences. Funding for the project has been provided by the National Development and Reform Commission. LAMOST is operated and managed by the National Astronomical Observatories, Chinese Academy E.A.P. acknowledges support from Hubble Fellow- ship grant HST-HF2-51365.001-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. L.M.W. acknowledges support from Gloria and Ken Levy and from the Trottier Family. T.D.M. acknowledges NASA grant NNX14AE11G. This work made use of NASA's Astrophysics Data System Bibliographic Services. Finally, the authors wish to recognize and acknowl- edge the very significant cultural role and reverence that the summit of Maunakea has always had within the in- digenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain. Software: All code used in this paper is available at https://github.com/California-Planet-Search/ the following pub- cksmet/. We made use of astropy (Astropy licly available Python modules: (Huber Collaboration et al. 2017), (Newville et al. 2014), mat- plotlib (Hunter 2007), numpy/scipy (van der Walt et al. 2011), pandas (McKinney 2010), and pinky (https://github.com/pgromano/pinky). isoclassify et lmfit 2013), al. REFERENCES Adams, E. R., Ciardi, D. R., Dupree, A. K., et al. 2012, AJ, 144, Christiansen, J. L., Clarke, B. D., Burke, C. J., et al. 2015, ApJ, 42 810, 95 Adams, E. R., Dupree, A. K., Kulesa, C., & McCarthy, D. 2013, Cui, X.-Q., Zhao, Y.-H., Chu, Y.-Q., et al. 2012, Research in AJ, 146, 9 Akeson, R. L., Chen, X., Ciardi, D., et al. 2013, PASP, 125, 989 Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, A&A, 558, A33 Baranec, C., Ziegler, C., Law, N. M., et al. 2016, AJ, 152, 18 Barclay, T., Rowe, J. F., Lissauer, J. J., et al. 2013, Nature, 494, 452 Astronomy and Astrophysics, 12, 1197 Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP, 120, 531 Dawson, R. I., Chiang, E., & Lee, E. J. 2015, MNRAS, 453, 1471 Dawson, R. I., & Murray-Clay, R. A. 2013, ApJL, 767, L24 De Cat, P., Fu, J. N., Ren, A. B., et al. 2015, ApJS, 220, 19 Dong, S., Xie, J.-W., Zhou, J.-L., Zheng, Z., & Luo, A. 2017, Batygin, K., Bodenheimer, P. H., & Laughlin, G. P. 2016, ApJ, ArXiv e-prints, arXiv:1706.07807 829, 114 Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Tamura, M. 2015, ApJS, 216, 7 Dong, S., & Zhu, Z. 2013, ApJ, 778, 53 Dong, S., Zheng, Z., Zhu, Z., et al. 2014, ApJL, 789, L3 Dressing, C. D., Adams, E. R., Dupree, A. K., Kulesa, C., & Brown, T. M., Latham, D. W., Everett, M. E., & Esquerdo, McCarthy, D. 2014, AJ, 148, 78 G. A. 2011, AJ, 142, 112 Buchhave, L. A., & Latham, D. W. 2015, ApJ, 808, 187 Buchhave, L. A., Latham, D. W., Johansen, A., et al. 2012, Nature, 486, 375 Dressing, C. D., & Charbonneau, D. 2015, ApJ, 807, 45 Everett, M. E., Barclay, T., Ciardi, D. R., et al. 2015, AJ, 149, 55 Fischer, D. A., & Valenti, J. 2005, ApJ, 622, 1102 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. Buchhave, L. A., Bizzarro, M., Latham, D. W., et al. 2014, 2013, PASP, 125, 306 Nature, 509, 593 Foreman-Mackey, D., Hogg, D. W., & Morton, T. D. 2014, ApJ, Burke, C. J., Christiansen, J. L., Mullally, F., et al. 2015, ApJ, 795, 64 809, 8 Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, Cartier, K. M. S., Gilliland, R. L., Wright, J. T., & Ciardi, D. R. 81 2015, ApJ, 804, 97 Catanzarite, J., & Shao, M. 2011, ApJ, 738, 151 Chen, H., & Rogers, L. A. 2016, ApJ, 831, 180 Chiang, E., & Laughlin, G. 2013, MNRAS, 431, 3444 Fulton, B. J., Petigura, E. A., Howard, A. W., et al. 2017, AJ, 154, 109 Furlan, E., Ciardi, D. R., Everett, M. E., et al. 2017, AJ, 153, 71 Ghezzi, L., Cunha, K., Smith, V. V., et al. 2010, ApJ, 720, 1290 Gilliland, R. L., Cartier, K. M. S., Adams, E. R., et al. 2015, AJ, 149, 24 Gilliland, R. L., Brown, T. M., Guhathakurta, P., et al. 2000, ApJL, 545, L47 Gonzalez, G. 1997, MNRAS, 285, 403 Goodman, J., & Weare, J. 2010, Communications in Applied Mathematics and Computational Science, 5, 65 Guo, X., Johnson, J. A., Mann, A. W., et al. 2017, ApJ, 838, 25 Hadden, S., & Lithwick, Y. 2017, AJ, 154, 5 Hansen, B., & Murray, N. 2013, ArXiv e-prints, arXiv:1301.7431 Hirano, T., Narita, N., Shporer, A., et al. 2011, PASJ, 63, 531 Horch, E. P., Howell, S. B., Everett, M. E., & Ciardi, D. R. 2012, AJ, 144, 165 -. 2014, ApJ, 795, 60 Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS, 201, 15 Howell, S. B., Everett, M. E., Sherry, W., Horch, E., & Ciardi, D. R. 2011, AJ, 142, 19 Huber, D., Zinn, J., Bojsen-Hansen, M., et al. 2017, ApJ, 844, 102 Hunter, J. D. 2007, Computing In Science & Engineering, 9, 90 Ida, S., & Lin, D. N. C. 2008, ApJ, 673, 487 Jenkins, J. M., Caldwell, D. A., Chandrasekaran, H., et al. 2010, ApJL, 713, L87 Jin, S., Mordasini, C., Parmentier, V., et al. 2014, ApJ, 795, 65 Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R. 2010, PASP, 122, 905 Johnson, J. A., Petigura, E. A., Fulton, B. J., et al. 2017, AJ, 154, 108 Lambrechts, M., & Lega, E. 2017, A&A, 606, A146 Law, N. M., Morton, T., Baranec, C., et al. 2014, ApJ, 791, 35 Lee, E. J., & Chiang, E. 2015, ApJ, 811, 41 -. 2017, ApJ, 842, 40 Lillo-Box, J., Barrado, D., & Bouy, H. 2012, A&A, 546, A10 -. 2014, A&A, 566, A103 Lissauer, J. J. 1995, Icarus, 114, 217 Lopez, E. D., & Fortney, J. J. 2013, ApJ, 776, 2 Luo, A.-L., Zhang, H.-T., Zhao, Y.-H., et al. 2012, Research in Astronomy and Astrophysics, 12, 1243 Luo, A.-L., Zhao, Y.-H., Zhao, G., et al. 2015, Research in Astronomy and Astrophysics, 15, 1095 Marcy, G. W., Isaacson, H., Howard, A. W., et al. 2014, ApJS, 210, 20 Masuda, K., & Winn, J. N. 2017, AJ, 153, 187 Mathur, S., Huber, D., Batalha, N. M., et al. 2017, ApJS, 229, 30 Mazeh, T., Holczer, T., & Faigler, S. 2016, A&A, 589, A75 McKinney, W. 2010, in Proceedings of the 9th Python in Science Conference, ed. S. van der Walt & J. Millman, 51 – 56 Morton, T. D. 2015, isochrones: Stellar model grid package, Astrophysics Source Code Library, , , ascl:1503.010 Mulders, G. D., Pascucci, I., & Apai, D. 2015, ApJ, 798, 112 Mulders, G. D., Pascucci, I., Apai, D., Frasca, A., & Molenda-Żakowicz, J. 2016, AJ, 152, 187 Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015, ApJS, 217, 31 27 Naoz, S., Farr, W. M., Lithwick, Y., Rasio, F. A., & Teyssandier, J. 2011, Nature, 473, 187 Newville, M., Stensitzki, T., Allen, D. B., & Ingargiola, A. 2014, LMFIT: Non-Linear Least-Square Minimization and Curve-Fitting for Python, , , doi:10.5281/zenodo.11813 Nordström, B., Mayor, M., Andersen, J., et al. 2004, A&A, 418, Owen, J. E., & Wu, Y. 2013, ApJ, 775, 105 -. 2017, ApJ, 847, 29 Petigura, E. A. 2015, PhD thesis, University of California, Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Proceedings of the National Academy of Science, 110, 19273 Petigura, E. A., Sinukoff, E., Lopez, E. D., et al. 2017a, AJ, 153, Petigura, E. A., Howard, A. W., Marcy, G. W., et al. 2017b, AJ, Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 989 Berkeley 142 154, 107 124, 62 Press, W. H. 2002, Numerical recipes in C++ : the art of scientific computing, ed. Press, W. H. Rauer, H., Catala, C., Aerts, C., et al. 2014, Experimental Astronomy, 38, 249 Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2015, Journal of Astronomical Telescopes, Instruments, and Systems, 1, 014003 Rogers, L. A. 2015, ApJ, 801, 41 Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153 Schlaufman, K. C. 2015, ApJL, 799, L26 Sinukoff, E., Howard, A. W., Petigura, E. A., et al. 2017, AJ, Smalley, B., Anderson, D. R., Collier-Cameron, A., et al. 2012, Sousa, S. G., Santos, N. C., Mayor, M., et al. 2008, A&A, 487, Sullivan, P. W., Winn, J. N., Berta-Thompson, Z. K., et al. 2015, Traub, W. A. 2012, ApJ, 745, 20 van der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, Computing in Science Engineering, 13, 22 Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, 2198, 362 Wang, J., & Fischer, D. A. 2015, AJ, 149, 14 Wang, J., Fischer, D. A., Horch, E. P., & Xie, J.-W. 2015a, ApJ, 153, 271 A&A, 547, A61 373 ApJ, 809, 77 Wang, J., Fischer, D. A., Xie, J.-W., & Ciardi, D. R. 2015b, Weiss, L. M., & Marcy, G. W. 2014, ApJL, 783, L6 Winn, J. N., Johnson, J. A., Howard, A. W., et al. 2010, ApJL, Wright, J. T., Marcy, G. W., Howard, A. W., et al. 2012, ApJ, 806, 248 ApJ, 813, 130 723, L223 753, 160 Wu, Y., & Murray, N. 2003, ApJ, 589, 605 Youdin, A. N. 2011, ApJ, 742, 38 Zapolsky, H. S., & Salpeter, E. E. 1969, ApJ, 158, 809 Zhao, G., Zhao, Y.-H., Chu, Y.-Q., Jing, Y.-P., & Deng, L.-C. 2012, Research in Astronomy and Astrophysics, 12, 723 Ziegler, C., Law, N. M., Morton, T., et al. 2017, AJ, 153, 66 APPENDIX A. CALIBRATION OF LAMOST METALLICITIES This paper relies on CKS metallicities for the planet-hosting stars and LAMOST metallicities to characterize the metallicity distribution of Kepler field stars. Differences in metallicity scales often result from different spectral datasets or spectroscopic pipelines. The CKS and LAMOST surveys used different spectra, line-lists, and model-atmospheres. 28 Here, we assess the agreement between the CKS and LAMOST metallicity scales, correct for a small offset, and report the uncertainty in our calibration. There are 476 stars observed by both CKS and LAMOST. Figure A1 shows the difference between LAMOST and CKS metallicities (∆[Fe/H]) as a function of CKS metallicity, Kp, and the S/N of the LAMOST spectra. On average, the LAMOST metallicities are 0.04 dex lower and we observe a small metallicity-dependent systematic trend. The dispersion of ∆[Fe/H] increases for fainter stars or decreasing LAMOST S/N (the CKS spectra have homogeneous S/N). This indicates that the precision of the LAMOST metallicities declines with decreasing S/N. We derived a correction ∆ that calibrates the LAMOST metallicities onto the CKS scale via LAMOSTcal = LAMOSTraw + ∆. The correction is a linear function of the following form: (cid:18) [Fe/H] (cid:19) 0.1 dex ∆[Fe/H] = c0 + c1 . (A1) The coefficients are chosen such that they minimize the RMS difference between the calibrated LAMOST and CKS parameters. Before fitting, we removed 17 stars where the CKS and LAMOST metallicities differed by more than 0.2 dex. These outliers are likely due to rare failure modes of the LAMOST pipeline.14 We estimated the uncertainties on the best-fit parameters using bootstrap resampling (with replacement), as de- scribed in Press (2002). The best-fit coefficients are c0 = 0.037± 0.002 and c1 = −0.015± 0.001. We show a one-to-one comparison of the CKS and LAMOST metallicities (without the 17 outliers) in Figure A2. After placing the LAMOST metallicities on the CKS scale, there is no residual mean offset (by construction) and a RMS dispersion of 0.05 dex. Our calibration amounts to a correction of +0.110 ± 0.007 dex at [Fe/H] = −0.5 dex and −0.036 ± 0.007 dex at [Fe/H] = +0.5 dex. The uncertainty associated with our correction at −0.5 dex and +0.5 dex sets an upper bound on the residual systematic offset between the LAMOST and CKS metallicity scales relevant to this paper, which treats planet hosts having metallicities between −0.5 and +0.5 dex. We estimate that the CKS and the calibrated LAMOST metallicities are consistent to 0.01 dex. To verify that the main results of this paper are not sensitive to zero-point offsets between the CKS and LAMOST metallicities scales, we performed a parallel analysis after perturbing our calibrated LAMOST metallicities by 0.01 dex. We compared the slopes of the metallicity distributions for different planet classes, which is quantified by the β index in Equation 19. Adding 0.01 dex to the LAMOST metallicities increases ntrial in the high metallicity bins, which results in lower measured occurrence at high metallicities. For low metallicity bins, ntrial decreases, resulting larger measured occurrence. The net effect is that the slope of the metallicity distribution becomes less steep (i.e., smaller β indices). (Subtracting 0.01 dex from the calibrated LAMOST metallicities has the opposite effect and results in larger beta-indices.) Figure A3 is analogous to Figure 11, but we created it after shifting the LAMOST metallicities by +0.01 dex. Comparing the β indices, these two figures summarize the effects of a possible offset in metallicity scales. For all planet classes, β decreases, as expected. For example, for warm super-Earths, β decreases from −0.3 ± 0.2 to −0.4 ± 0.2; for warm sub-Neptunes, β decreases from +0.5 ± 0.2 to +0.4 ± 0.2. For all planet classes, the change in β is smaller than our adopted 1σ uncertainties from our maximum-likelihood fitting, which only incorporates counting statistics. Therefore, we conclude that the uncertainties on the β indices due to our LAMOST-CKS calibration are smaller than those due to counting statistics. Thus, the conclusions of this paper are not sensitive to errors in our LAMOST-CKS metallicity calibration. B. METALLICITY AND PLANET DETECTABILITY In Section 3, we compared the properties of planets belonging to host stars of different metallicities, and in Section 6, we computed planet occurrence as a function of host star metallicity. For both calculations, we made the assumption that planet detectability is not a strong function of host star metallicity. Here, we check the validity of such an assumption. Metallicity could correlate with planet detectability through a correlation with stellar size or noise properties. As a direct measure of the possible dependence of planet detectability with metallicity, we computed the single transit signal-to-noise ratio (S/N1) of a 3-hour transit of putative 1 R⊕ planet transiting each star in the LAMOST sample. 14 To verify that our calibration does not depend sensitively on our choice to remove outliers, we left them in but solved for the coefficients that minimized the sum of the absolute differences, a metric which is resistant to outliers. The coefficients agreed to 1.5σ or better. 29 Figure A1. Difference between CKS and LAMOST [Fe/H] for 476 stars in common as a function of CKS metallicity (a), Kp (b), and the reported S/N of the LAMOST spectra (c). We observe a small metallicity-dependent systematic difference, which we calibrate out in Appendix A. Figure A2. Comparison of CKS and LAMOST metallicities. (Left) Uncalibrated LAMOST metallicities as a function of CKS metallicities for stars in common. The green dashed line represents equality. On average, the LAMOST metallicities are 0.05 dex lower than the CKS values. (Right) comparison of CKS and LAMOST metallicities after removing a linear systematic trend. S/N1 was computed according to (cid:19)2(cid:18) (cid:18) RP R(cid:63) 1 CDPP3 (cid:19) , S/N1 = (B2) where RP = 1 R⊕ and R(cid:63) was computed using LAMOST Teff, log g, and [Fe/H] measurements and the publicly available isoclassify package (Huber et al. 2017).15 As Figure B4 shows, we do not observe a significant dependence of planet detectability with stellar metallicity. 15 https://github.com/danxhuber/isoclassify −0.4−0.20.00.20.4[Fe/H] (CKS)−0.4−0.20.00.20.4Δ [Fe/H] (LAMOST-CKS)a101214Kpb3001003010LAMOST (SNR)c−0.4−0.20.00.20.4−0.4−0.20.00.20.4[Fe/H] (dex) [LAMOST]Mean(Δ) = -0.039 dexRMS(Δ) = 0.057 dex−0.4−0.20.00.20.4[Fe/H] (dex) [CKS]−0.150.000.15LAMOST - CKS−0.4−0.20.00.20.4−0.4−0.20.00.20.4[Fe/H] (dex) [LAMOST]Mean(Δ) = 0.000 dexRMS(Δ) = 0.049 dex−0.4−0.20.00.20.4[Fe/H] (dex) [CKS]−0.150.000.15LAMOST - CKS 30 Figure A3. Same as Figure 11, but after adding 0.01 dex to the LAMOST metallicities to simulate the effect of residual offsets between the CKS and LAMOST metallicity scales. For all planet classes, the change in β is smaller than the adopted uncertainty on β, which incorporates counting statistics alone. The uncertainties on β are therefore dominated by counting statistics, rather than uncertainties in our CKS-LAMOST metallicity calibration. Figure B4. Planet detectability vs. metallicity. The y-axis shows the signal-to-noise ratio of a putative 1 R⊕ planet, which has a transit duration of three hours. We do not observe a significant trend in planet detectability with metallicity. 1101002UEiWaO PHUiod (days)11.74.08.024.0POanHW sizH ((aUWh-Uadii)HoW SuSHU-(aUWhs ⟨[FH/+]⟩ +0.05±0.01β=+0.5+0.2−0.2HoW SuE-1HSWunHs ⟨[FH/+]⟩ +0.11±0.02β=+1.5+0.3−0.3HoW SuE-SaWuUns ⟨[FH/+]⟩ +0.26±0.04β=+5.2+1.6−1.4HoW -uSiWHUs ⟨[FH/+]⟩ +0.19±0.04β=+3.3+0.9−0.9WaUP SuSHU-(aUWhs ⟨[FH/+]⟩ −0.04±0.02β=−0.4+0.2−0.2WaUP SuE-1HSWunHs ⟨[FH/+]⟩ +0.04±0.01β=+0.4+0.2−0.2WaUP SuE-SaWuUns ⟨[FH/+]⟩ +0.14±0.02β=+2.0+0.7−0.7WaUP -uSiWHUs ⟨[FH/+]⟩ +0.06±0.02β 8nconsWUainHd−0.4−0.3−0.2−0.10.00.10.20.30.4[Fe/H] (dex)0.10.31310Single Transit SNR 31 C. PLANET OCCURRENCE: PERIOD AND RADIUS Here we provide two tables to supplement Section 5 which treats planet occurrence as a function of period and radius. Table C1 lists the occurrence measurements displayed in Figure 6 along with upper limits. Table C2 is a sampling of the occurrence distribution shown in Figure 8. We sample the occurrence the where the following conditions are met: 1. P = 1–300 days (i.e., the range of periods shown in Figure 8). 2. RP = 0.5–32 R⊕ (i.e., the range of planet sizes shown in Figure 8). 3. log (RP /R⊕) > 0.15 log (P/1 day), which ensures that pipeline completeness, pdet > 0.25. The integrated occurrence within this domain is 110.7 planets per 100 stars. We simulated the periods and radii of a population of 110733 planets in a sample of 100000 Sun-like stars by drawing 110733 (P ,RP ) pairs according to their measured occurrence rates using the Python package pinky. These samples are listed in Table C2. Table C1. Planet Occurrence RP,1 RP,2 1.41 1.00 1.41 1.00 1.41 1.00 1.00 1.41 1.41 1.00 1.41 1.00 1.41 1.00 1.00 1.41 1.41 1.00 1.00 1.41 P1 1.00 1.78 3.16 5.62 10.00 17.78 31.62 56.23 100.00 177.83 P2 1.78 3.16 5.62 10.00 17.78 31.62 56.23 100.00 177.83 316.23 npl 7 15 40 48 51 23 10 4 0 0 pdet 0.95 0.93 0.89 0.84 0.76 0.66 0.53 0.39 0.27 0.16 ntrial 6551.6 4369.1 2874.8 1847.5 1144.3 673.5 370.5 187.4 86.4 36.3 fcell 0.12+0.05−0.04 0.36+0.10−0.08 1.41+0.23−0.21 2.63+0.39−0.36 4.50+0.64−0.59 3.50+0.76−0.65 2.87+0.95−0.78 2.49+1.26−0.97 < 2.61 ··· Note-Planet occurrence computed over various intervals in the P –RP plane spanning 0.25 dex and 0.15 dex, respectively. Each bin boundary is given by [RP,1,RP,2] and [P1,P2] respectively. We list the number of planets per bin npl, the pipeline detectability pdet, and the number of effective trials ntrial. The number of planets per 100 stars per bin is given by fcell, which is also shown graphically in Figure 6. We report 90% upper limits on fcell when there are no planets in a bin, and we do not report fcell when pdet < 0.25. Table C1 is available in its entirety in machine-readable format. A portion is shown here for guidance regarding its form and content. Table C2. Simulated Planet Properties Planet 0 1 2 3 P days 45.194 94.327 56.235 44.367 RP R⊕ 1.185 0.964 1.100 2.254 Table C2 continued 32 Table C2 (continued) Planet 4 5 6 7 8 9 P days 26.778 67.644 28.391 236.817 RP R⊕ 0.827 6.516 1.366 7.830 336.385 11.736 22.677 2.457 Note-Simulated periods and radii of 110733 plan- ets in a population of 100000 stars based on the measured occurrence rates from Kepler (Sec- tion 5). This table may be used to compute yield simulations for future surveys or integrated oc- currence values over arbitrary bins of P and RP . Table C2 is available in its entirety in machine- readable format. A portion is shown here for guid- ance regarding its form and content.
1612.02425
2
1612
2017-04-06T13:27:37
Detection of the atmosphere of the 1.6 Earth mass exoplanet GJ 1132b
[ "astro-ph.EP", "astro-ph.SR" ]
Detecting the atmospheres of low-mass low-temperature exoplanets is a high-priority goal on the path to ultimately detect biosignatures in the atmospheres of habitable exoplanets. High-precision HST observations of several super-Earths with equilibrium temperatures below 1000K have to date all resulted in featureless transmission spectra, which have been suggested to be due to high-altitude clouds. We report the detection of an atmospheric feature in the atmosphere of a 1.6 Mearth transiting exoplanet, GJ 1132b, with an equilibrium temperature of ~600K and orbiting a nearby M dwarf. We present observations of nine transits of the planet obtained simultaneously in the griz and JHK passbands. We find an average radius of 1.43 +/- 0.16 Rearth for the planet, averaged over all the passbands, and a radius of 0.255 +/- 0.023 Rsun for the star, both of which are significantly greater than previously found. The planet radius can be decomposed into a "surface radius" at ~1.375 Rearth overlaid by atmospheric features which increase the observed radius in the z and K bands. The z-band radius is 4sigma higher than the continuum, suggesting a strong detection of an atmosphere. We deploy a suite of tests to verify the reliability of the transmission spectrum, which are greatly helped by the existence of repeat observations. The large z-band transit depth indicates strong opacity from H2O and/or CH4 or a hitherto unconsidered opacity. A surface radius of 1.375 +/- 0.16 Rearth allows for a wide range of interior compositions ranging from a nearly Earth-like rocky interior, with ~70% silicate and ~30% Fe, to a substantially H2O-rich water world.
astro-ph.EP
astro-ph
DRAFT VERSION APRIL 7, 2017 Preprint typeset using LATEX style AASTeX6 v. 1.0 DETECTION OF THE ATMOSPHERE OF THE 1.6 EARTH MASS EXOPLANET GJ 1132 B JOHN SOUTHWORTH 1, LUIGI MANCINI 2,3,4, NIKKU MADHUSUDHAN 5, PAUL MOLLI `ERE 2, SIMONA CICERI 6, THOMAS HENNING 2 1 Astrophysics Group, Keele University, Staffordshire, ST5 5BG, UK 2 Max Planck Institute for Astronomy, K onigstuhl 17, 69117 Heidelberg, Germany 3 Dipartimento di Fisica, Universit`a di Roma Tor Vergata, Via della Ricerca Scientifica 1, 00133 – Roma, Italy 4 INAF – Osservatorio Astrofisico di Torino, via Osservatorio 20, 10025, Pino Torinese, Italy 5 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 6 Department of Astronomy, Stockholm University, SE-106 91 Stockholm, Sweden ABSTRACT Detecting the atmospheres of low-mass low-temperature exoplanets is a high-priority goal on the path to ul- timately detect biosignatures in the atmospheres of habitable exoplanets. High-precision HST observations of several super-Earths with equilibrium temperatures below 1000 K have to date all resulted in featureless transmission spectra, which have been suggested to be due to high-altitude clouds. We report the detection of an atmospheric feature in the atmosphere of a 1.6 M⊕ transiting exoplanet, GJ 1132 b, with an equilibrium temperature of ∼600 K and orbiting a nearby M dwarf. We present observations of nine transits of the planet obtained simultaneously in the griz and J HK passbands. We find an average radius of 1.43 ± 0.16 R⊕ for the planet, averaged over all the passbands, and a radius of 0.255 ± 0.023 R⊙ for the star, both of which are significantly greater than previously found. The planet radius can be decomposed into a "surface radius" at ∼1.375 R⊕ overlaid by atmospheric features which increase the observed radius in the z and K bands. The z-band radius is 4σ higher than the continuum, suggesting a strong detection of an atmosphere. We deploy a suite of tests to verify the reliability of the transmission spectrum, which are greatly helped by the existence of repeat observations. The large z-band transit depth indicates strong opacity from H2O and/or CH4 or a hitherto unconsidered opacity. A surface radius of 1.375 ± 0.16 R⊕ allows for a wide range of interior compositions ranging from a nearly Earth-like rocky interior, with ∼70% silicate and ∼30% Fe, to a substantially H2O-rich water world. Keywords: planetary systems - stars: fundamental parameters - stars: individual: GJ 1132 1. INTRODUCTION large numbers of planets M dwarfs are bounteous throughout our Galaxy, and (Cassan et al. 2012; host Dressing & Charbonneau 2015). The myriad of planets orbiting M dwarfs is dominated by small and low-mass rocky bodies with masses and radii comparable to Earth's: (Morton & Swift 2014; Gaidos et al. 2016), and many occur in multiple systems (Muirhead et al. 2015). Planets around M dwarfs are of particular interest because the dimness of the host stars means their habitable zones (e.g. Kopparapu et al. 2013) are located at short orbital periods, which are much more accessible to observational study (Gillon et al. 2016). However, planets within the habitable zone of M dwarfs face additional challenges such as tidal locking and large incident fluxes of high-energy photons (Lammer et al. 2003; Cunha et al. 2015; Shields et al. 2016). From an analysis of the M dwarfs observed by the Kepler satellite, Gaidos et al. (2016) found that there were on aver- age 2.2 planets per star and that the radius distribution peaked at 1.2 R⊕. These objects are expected to have tenuous at- mospheres, with those of radius below approximately 1.5– 1.6 R⊕ being almost entirely rocky (Weiss & Marcy 2014; Rogers 2015). This raises a problem for the overarching goal of constraining the chemical compositions and atmo- spheric characteristics of rocky planets, because the obser- vational signatures of their atmospheres are below the levels detectable with current facilities. GJ 1132 is a nearby very-low-mass star which was re- cently found to host a transiting and potentially rocky planet (Berta-Thompson et al. 2015, hereafter BT15) of mass 1.6 M⊕ and radius 1.2 R⊕. Its proximity to the Sun (12.04± 0.24 pc; Jao et al. 2005) means that it is comparatively bright and therefore well-suited to analyses aimed at constraining the properties of both the star and the planet. Schaefer et al. (2016) presented simulations of the interior and atmosphere of GJ 1132 b, finding that the atmosphere is likely tenuous and dominated by O2. Whilst significantly hotter than Earth, the planet is one of the coolest transiting planets of known mass and is therefore of great interest for comparative plan- etology. BT15 found that GJ 1132 A is an old (5 Gyr or 2 more) and slowly-rotating (rotation period 125 d) M4.5 V star, making it representative of a large population of planet host stars expected to be found in the (relatively) near future (Cloutier et al. 2017). Dittmann et al. (2016, hereafter D16) presented extensive photometry of the GJ 1132 system, comprising 21 transits ob- served with the MEarth telescopes and a 100 hr light curve from the Spitzer satellite. D16 presented revised properties for the system, and searched for transit timing variations or additional transits which would be caused by the presence of a third body in the system. In this work we present extensive simultaneous optical and near-infrared photometry of nine transits of the planet GJ 1132 b in front of the star GJ 1132 A. We use these data to redetermine the physical properties of the system, improve the fidelity of its orbital ephemeris, and construct a transmis- sion spectrum of the planet. We then interpret the transmis- sion spectrum using suites of theoretical spectra from two model atmosphere codes. We clearly detect the planetary at- mosphere but the data in hand are not able to resolve ambi- guities in the relative contributions of different molecules to the atmospheric opacity. At 1.6 M⊕ GJ 1132 b is by a substantial factor the lowest- mass planet with an atmosphere which has been observation- ally detected. The two other low-mass planets with claimed detections are 55 Cnc e (Winn et al. 2011), with a mass of 8.08±0.31 M⊕ (Demory et al. 2016), for which Tsiaras et al. (2016) found atmospheric features that could most easily be explained by HCN opacity, and GJ 3470 b (Bonfils et al. 2012), with a mass of 13.7 ± 1.6 M⊕ (Biddle et al. 2014), which shows an enhanced radius in the blue attributable to Rayleigh absorption (Nascimbeni et al. 2013; Biddle et al. 2014; Dragomir et al. 2015). Three other low-mass planets have been subjected to transmission spectroscopy which has failed to reveal at- mospheric signatures, most likely due to the presence of clouds or of a high-metallicity atmosphere with a large mean molecular mass. They are GJ 1214 b (Charbonneau et al. 2009), with a mass of 6.26±0.91 M⊕ (Anglada-Escud´e et al. 2013), for which no atmospheric features have been detected (Bean et al. 2010, 2011; de Mooij et al. 2012; Kreidberg et al. 2014b), HD 97658 b (Dragomir et al. 2013), with a mass of 7.55+0.83 −0.79 M⊕ (Van Grootel et al. 2014), for which there was also a non-detection of the atmosphere (Knutson et al. 2014b), and GJ 436 b (Gillon et al. 2007), with a mass of 25.4±2.1 M⊕ (Lanotte et al. 2014), for which Knutson et al. (2014a) found no atmospheric features. 2. OBSERVATIONS AND DATA REDUCTION Extensive observations of GJ 1132 were obtained in service mode using the GROND multi-band imager (Greiner et al. 2008) mounted on the MPG 2.2 m telescope at ESO La Silla, Chile. This instrument acquires images simul- taneously in four optical and three near-infrared passbands. Table 1. Dates and numbers of the observations presented in this work. Observing night Number of observations 2016/01/14 2016/02/14 2016/02/24 2016/02/27 2016/03/26 2016/03/29 2016/04/03 2016/04/08 2016/05/17 g 54 59 108 116 104 94 62 94 62 r 63 59 104 121 104 96 63 93 64 i 72 44 108 124 101 95 63 94 68 z 69 J H K 252 224 123 372 364 369 168 358 97 60 75 65 385 353 251 280 251 277 167 273 A total of nine transits were observed, using the telescope- defocussing approach (Alonso et al. 2008; Southworth et al. 2009) with point spread functions (PSFs) of typically 30 pix- els in radius. A tenth transit observation suffered from a large systematic effect during transit, due to either weather or in- strumental effects, so was not included in our analysis. The optical bands have response functions similar to those of the SDSS g, r, i and z filters (Fukugita et al. 1996). They are equipped with 2k×2k CCDs which see (approximately) the same 5.4′×5.4′ field, with a plate scale of 0.158′′ pixel−1, and are each read out through two amplifiers. The near- infrared bands have 1k×1k Rockwell HAWAII-1 detectors with a field of view of 10′×10′, which fully encompasses the optical field, at a plate scale of 0.6′′ pixel−1 Several phenomena affected the quality of our optical ob- servations. Firstly, GJ 1132 is observationally difficult for the GROND imager due to its low Teff and concomitant large variations in flux level through the optical wavelength region. The requirement for a common exposure time and focus level meant a compromise had to be made between obtaining ad- equate count rates in the g band whilst avoiding overexpo- sure in the z band. Three of the z-band observing sequences suffered from saturation effects which precluded the extrac- tion of reliable photometry from these images. Secondly, an issue was encountered in the r and z bands attributable to electronic noise in the CCD controllers. This caused approxi- mately 0.3% of the pixels on each image to be assigned count rates near the bias level, with the identities of the affected pixels varying at random with each image. When these coin- cided with the point spread functions (PSFs) of the target or comparison stars they caused a non-astrophysical drop in the number of counts detected from the star, increasing the scat- ter on the photometry. A non-repeating effect such as this cannot be calibrated out, but a partial mitigation of the effect was achieved by detecting each pixel with anomalously low count rates and replacing its value with the mean level from the adjacent eight pixels. Thirdly, the small field of view of GROND meant that the r, i and z bands suffered from a 3 Figure 1. The optical light curves of GJ 1132 obtained using GROND, arranged in rows according to passband and in columns according to date. The filter and date are encoded in the names of the datafile printed at the base of each panel. Each plot covers a total of 0.14 d centred on the time given on the x-axis. lack of decent comparison stars. Photometry was therefore performed against comparison stars which were significantly fainter than GJ 1132 itself. A brief observing log of the nine included transits is given in Table 1. The optical data were reduced using the DEFOT pipeline (Southworth et al. 2009, 2014), which performs aperture photometry using the APER routine from DAOPHOT (Stetson 1987) contained in the NASA ASTROLIB library1. The aper- ture positions and sizes were specified manually in order to yield data with the lowest statistical and systematic er- rors. The science images were not calibrated using bias or flat-field frames as these tended to have little effect on the final light curves beyond a slight increase in the scat- ter of the datapoints. Differential magnitudes for each light curve were formed versus an optimal ensemble of compar- ison stars, whilst simultaneously fitting for and removing a quadratic trend of magnitude with time caused primarily by airmass variations. The timestamps were converted into the BJD(TDB) system using routines from Eastman et al. (2010). As a final step, we rescaled the errorbars for each light curve to obtain a reduced χ2 of χ 2 ν = 1.0 versus a fitted model (see below). The light curves are shown in Fig. 1 and will be made available at the CDS2. The infrared data were reduced using standard IDL rou- tines following the methods outlined by Chen et al. (2014). In brief, we constructed a master dark frame and a master flat field by median-combining the individual dark and sky flat field images obtained before the science observations. From each science image we subtracted the master dark frame and divided by the normalised master flat field. No correction for non-linearity was applied to the data because the counts were below the level at which linearity becomes important for the GROND infrared detectors. We obtained the light curves us- ing aperture photometry routines to extract the flux of the target and several comparison stars. We also tried to correct for the odd-even readout pattern present along the x-axis, but found no improvement in the light curve so decided not to include the correction for this effect when obtaining our final 1 http://idlastro.gsfc.nasa.gov/ 2 http://cdsweb.u-strasbg.fr 4 light curves. We were able to obtain useful results for five transits in each band (Table 1); the remaining datasets suf- fered from correlated noise features which were larger than the transit depth. 3. LIGHT CURVE MODELLING The optical GROND data were combined into four sets, one for each of the g, r, i and z passbands, and then each was modelled using the JKTEBOP code (Southworth 2013). We fitted for the orbital inclination (i), time of mid-transit (T0), and the sum and ratio of the fractional radii (rA+rb and k = rb ) where the fractional radii are those of the star and rA planet in units of the orbital semimajor axis (rA,b = RA a ). The orbital period was fixed to the known value (see Sec- tion 4) and the orbit was assumed to be circular. A quadratic function versus time was applied to the magnitude values for each observed transit in order to propagate the uncertainty in this from the light curve generation process. b , Limb darkening was incorporated using each of five laws (see Southworth 2008) and with the nonlinear coefficient fixed. Fits were obtained with the linear coefficient ei- ther fixed or fitted in order to see how this changed the re- sults. We adopted limb darkening coefficients calculated us- ing the PHOENIX model atmospheres by Claret (2004), and linearly interpolated the coefficients to the stellar temperature of Teff = 3270 ± 140 K (BT15). An additional complication arises due to the presence of the faint nearby star USNO B1.0 0428-0265237 (Monet et al. 2003), which was separated from GJ 1132 by approximately 6.5′′ at the times of our observations. A small fraction of the flux from this star leaked into the aperture used to mea- sure the counts of GJ 1132. From a simple PSF model we measured the amount of contamination as 1.2 ± 0.3% in the g-band, 0.5 ± 0.2% in r, 0.3 ± 0.1% in i and 0.2 ± 0.1% in z, where the errorbars are very conservative. This effect was included as 'third light' in the JKTEBOP model following the approach in Southworth (2010). The best fits are plotted in Fig. 2 in the case of fixed limb darkening coefficients. We also modelled the two best light curves of GJ 1132 from BT15, which are the g- and i-band data from the PISCO imager. Their appearance in Fig. 2 differs from that in BT15, leading us to investigate the discrepancy. We checked for correlation with the instrumental parameters supplied with the flux measurements (airmass, x-position, y-position, PSF width and sky background level), finding linear Pearson cor- relation coefficients less than 0.13 in all cases. We therefore conclude that the difference in appearance is only because BT15 binned their data into 1.5 min intervals before plotting it. Shortly before our own manuscript was submitted, D16 presented an analysis of GJ 1132 based on observations of 21 transits with the MEarth telescopes and 100 hr with the Spitzer satellite using the 4.5 µm channel of the IRAC im- ager. Whilst the original data are not available to us, we have been able to obtain binned light curves from both facilities by digitizing fig. 2 in D16. We used the scatter of the data around the best-fit model to assess the photometric precisions of the datasets. For the Spitzer data we used limb darkening coefficients from Claret et al. (2012) and for the MEarth data we adopted the mean of the coefficients for the i and z bands used above. For the Spitzer light curve we found that fit- ting for any limb darkening coefficients returned a nett limb brightening, due to the scatter of the data during transit. The MEarth light curve was able to support the fitting of one limb darkening coefficient. Results for both datasets are included in Table 2. Table 2 gives the best-fitting photometric parameters for each light curve, together with the weighted mean value for each parameter. Error estimates for the photometric param- eters were obtained using Monte Carlo (Southworth et al. 2004) and residual-permutation simulations (Southworth 2008), and the larger value retained in each case. These were then inflated to account for any variation between the solutions adopting the five different limb darkening laws by adding in quadrature the largest difference in values of each parameter from solutions with the various limb darkening laws. Finally, the weighted mean values agree well for all pa- rameters, indicating that the errorbars are robust. In all cases we have chosen to adopt the solutions with all limb darkening coefficients fixed. We find that the alternative solutions with one fitted coefficient agree to well within the uncertainties, with the exception of the Spitzer data as discussed above. There is a significant disagreement between the photomet- ric parameters determined by ourselves and those by BT15 and D16. An inspection of Table 2 shows that there is a cor- relation between the parameters rA and i, whereby higher- inclination solutions give a smaller fractional radius for the star. BT15 and D16 obtain solutions with higher i than our own, resulting in lower rA values of 0.0625 ± 0.0043 and 0.0605+0.0027 −0.0022, respectively. This can be explained in the case of BT15 because rA is a proxy for the density of the star (Sozzetti et al. 2007; Southworth 2017), so this quantity was effectively fixed in the light curve solutions by the impo- sition of external constraints on both the mass and radius of the host star. Such a constraint was not imposed in the current work or D16, so the discrepancy between these two works remains. We have checked and ruled out the possibility that it could be caused by the treatment of limb darkening or third light, by quantifying the effect of varying our own treatment of these phenomena within reasonable limits. We also tried numeri- cally integrating the fitted model (Southworth 2011) to match the cadence of the Spitzer and MEarth data but found this caused little change in the solutions. We also tried various initial values for the fitted parameters, including the values found by D16, but the fits to data converged on the same solutions as given in Table 2. Solution determinacy is there- fore not an issue, and the cause of the discrepancy remains 5 Figure 2. Optical light curves of GJ 1132 from this work (GROND), BT15 (PISCO) and D16 (Spitzer and MEarth), compared to the JKTEBOP best fits.The GROND data have been binned by a factor of five and plotted versus orbital phase in order to make the plot clearer; all analysis in this work was based on the original data. The Spitzer and MEarth datasets are digitized versions of data heavily binned before plotting. The residuals of the fits are plotted at the base of the figure, offset from unity. Labels give the source and passband for each dataset. The polynomial baseline functions have been removed from the data before plotting. unidentified. It is most likely due to differences in the overall analysis procedure. 4. ORBITAL EPHEMERIS Many of our light curves have a relatively high scatter com- pared to the transit depth. We therefore combined all four light curves of each individual transit into a single dataset and fitted them with JKTEBOP whilst fixing the sum of the radii (rA + rb) to the best fit from Table 2. This yielded nine measured times of mid-transit, which are given in Table 3. We augmented these timing measurements with the refer- ence time of mid-transit given by BT15 from a global analy- sis of their photometric data (2457184.55786± 0.00032) and with the 22 timings given by D16 from orbital cycle 91 on- wards. The timings were then fit with a straight line to yield the linear ephemeris: T0 = BJD(TDB) 2457184.55759(30) + 1.6289287(18)×E where the fit has χ 2 ν = 1.13 and the uncertainties (given in brackets and relative to the preceding digit) have been in- creased by √1.13 to account for this. The scatter around the best fit gives no indication of deviations from a linear ephemeris, so we do not find any evidence for transit timing variations. 6 Table 2. Parameters of the fit to the GROND and PISCO light curves of GJ 1132 from the JKTEBOP analysis. Source rA + rb k i (◦) rA rb GROND g-band 0.102 ± 0.028 0.0535 ± 0.0060 85.4 ± 2.0 0.097 ± 0.026 0.0052 ± 0.0019 GROND r-band 0.066 ± 0.019 0.0517 ± 0.0040 88.5 ± 2.1 0.063 ± 0.018 0.0032 ± 0.0012 GROND i-band 0.070 ± 0.014 0.0514 ± 0.0023 87.9 ± 1.6 0.067 ± 0.013 0.0034 ± 0.0008 GROND z-band 0.106 ± 0.028 0.0611 ± 0.0061 85.3 ± 1.9 0.100 ± 0.026 0.0061 ± 0.0021 PISCO g-band 0.117 ± 0.039 0.0633 ± 0.0086 84.5 ± 2.6 0.110 ± 0.036 0.0070 ± 0.0029 PISCO i-band 0.086 ± 0.030 0.0541 ± 0.0050 86.6 ± 2.4 0.082 ± 0.028 0.0044 ± 0.0019 Spitzer 4.5 µm 0.090 ± 0.014 0.0496 ± 0.0010 86.3 ± 1.0 0.086 ± 0.013 0.0043 ± 0.0007 MEarth 0.064 ± 0.024 0.0492 ± 0.0026 88.5 ± 1.8 0.061 ± 0.024 0.0030 ± 0.0013 Final results 0.0814 ± 0.0072 0.05041 ± 0.00086 86.58 ± 0.63 0.0775 ± 0.0068 0.00397 ± 0.00042 Table 3. Times of mid-transit from the data presented in this work. Orbital cycle Transit time (BJD/TDB) 0.0 2457184.55786 ± 0.00032 134.0 2457402.83351 ± 0.00026 153.0 2457433.78361 ± 0.00024 159.0 2457443.55662 ± 0.00039 161.0 2457446.81540 ± 0.00028 178.0 2457474.50661 ± 0.00030 180.0 2457477.76559 ± 0.00043 183.0 2457482.65244 ± 0.00046 186.0 2457487.53908 ± 0.00031 210.0 2457526.63273 ± 0.00030 5. PHYSICAL PROPERTIES OF GJ 1132 The physical properties of the GJ 1132 system were es- tablished by BT15 using the following steps. Firstly, the measured trigonometric parallax (BT15, Jao et al. 2005) and 2MASS J HK magnitudes were used to determine the K- band absolute magnitude of the star, from which its mass was found via the calibration by Delfosse et al. (2000). Secondly, an empirical calibration of mass versus density (Hartman et al. 2015) was used to find the density and thus radius of the star. Thirdly, the transit light curves were fitted with the stellar density constrained to the value found in the second step, which is equivalent to constraining rA. We are in a position to modify this approach to rely less on general empirical calibrations and more on data ob- tained for GJ 1132 itself. We adopted the stellar mass of 0.181 ± 0.019 M⊙ from BT15, and used this along with the photometric parameters measured in Section 3 to determine the physical properties of the system using standard equa- tions (e.g. Hilditch 2001). This process in effect used rA as a proxy for the density of the star (Seager & Mall´en-Ornelas 2003), which combined with its mass yields its radius. The properties of the planet could then be determined relative to those of the star using k and the velocity amplitude of the star measured by BT15, 2.76 ± 0.92 m s−1. The uncertain- ties were propagated by a Monte Carlo approach. The measured physical properties of the GJ 1132 system are summarised in Table 4. It can be seen that we find a rather lower density for the star – measured directly from the light curves whereas BT15 obtained their value from a calibration of stellar properties – which results in increased radii for both star and planet. The measured density of GJ 1132 b has de- creased by a factor of two (1.8σ), which has a significant impact on the likely properties of this body. 6. TRANSMISSION SPECTRUM OF GJ 1132 B The main aim of the current work is to explore if con- straints on the atmospheric composition of GJ 1132 b can be obtained using multi-band photometry. We therefore sought to determine the radius of the planet in each of the passbands for which we possess a light curve. It is important to remove sources of uncertainty common to all passbands, so that the significance of any relative variations between passbands can be assessed. Following the approach of Southworth et al. (2012), we modelled the nine available light curves (GROND griz, GROND J HK and PISCO gi) with rA, i and the orbital period fixed to the final values determined above. We in- cluded as fitted parameters rb, the reference time of mid- transit, and the quadratic function of magnitude versus time for each transit. We adopted quadratic limb darkening with the linear coefficient fitted and the nonlinear coefficient fixed to the values used above. As the rA and i parameters are fixed, it is not only more tractable but also more important to fit for limb darkening in order to propagate its uncertainty and avoid biases arising from the use of theoretical values. Third light was included for the optical bands and constrained as in Section 3. For the J HK bands we neglected third light, as it is negligible at infrared wavelengths, but iteratively clipped points lying more than 3σ from the best fit in order to re- move scattered data which was biasing the fitting process. The near-IR light curves and best fits are shown in Fig. 3. We did not include the data from D16 in this analysis because imperfections in the digitization process could significantly affect our results. In addition, the wavelength resolution of the MEarth data is poor and the Spitzer data cannot support Table 4. Derived physical properties of GJ 1132 from the current work. The values found by BT15 and D16 are included for comparison. 7 Quantity Stellar mass Stellar radius MA RA Symbol Unit This work BT15 D16 Stellar surface gravity log gA Stellar density Planet mass Planet radius Planet surface gravity ρA Mb Rb gb ρb Planet density Equilibrium temperature T ′ eq Orbital semimajor axis a M⊙ R⊙ cgs ρ⊙ M⊕ R⊕ m s−2 g cm−3 K au 0.181 ± 0.019 0.181 ± 0.019 0.255 ± 0.023 0.207 ± 0.016 0.2105+0.0102 −0.0085 4.881 ± 0.074 10.9+3.4 −2.4 1.63 ± 0.54 1.43 ± 0.16 ⋆ 7.8+3.4 −2.8 3.1+1.7 −1.2 644 ± 38 0.01533 ± 0.00053 21.0 ± 4.3 1.62 ± 0.55 19.4+2.6 −2.5 1.16 ± 0.11 1.130 ± 0.056 11.7 ± 4.3 6.0 ± 2.5 579 ± 15 6.2 ± 2.0 ⋆ This value of Rb is averaged over all observed passbands. However, while interpreting the spectrum we consider the z-band radius at 1.57 ± 0.05 R⊕ to be contributed by the planetary atmosphere whereas the remaining datapoints are consistent with a continuum, the "surface" of the planet, at a bulk radius of 1.375 R⊕. Table 5. Values of rb for each of the six light curves. The errorbars in this table exclude all common sources of uncertainty so should only be used to interpret relative differences in rb. To convert rb to Rb we multiplied by 23455.0, the orbital semimajor axis in units of R⊕. The central wavelengths and full widths at half maximum transmission are given for the GROND filters. Passband Central Band full rb wavelength (nm) width (nm) GROND g GROND r GROND i GROND z PISCO g PISCO i GROND J GROND H GROND K 477 623 763 913 1230 1645 2165 138 138 154 137 410 420 570 0.00382 ± 0.00011 0.00402 ± 0.00009 0.00386 ± 0.00006 0.00446 ± 0.00015 0.00438 ± 0.00010 0.00396 ± 0.00007 0.00354 ± 0.00045 0.00324 ± 0.00044 0.00473 ± 0.00058 The results of this process are given in Table 5 along with de- tails of the passbands used, and shown in Fig. 4. For the pur- poses of visualising our results we assumed that the PISCO passbands correspond to those from Fukugita et al. (1996). It can be seen that the g, r and i passbands agree well, except for a discrepancy between our own g-band results and those from the PISCO g-band data from BT15. We suggest that our rb value is more reliable as it is based on observations of nine transits whereas the PISCO data cover only one transit. It is also apparent that the radius of the planet in z is significantly larger than that in the other passbands, the discrepancy being a maximum of 4.1σ versus our i-band value of rb. 6.1. Testing the robustness of the results Figure 3. Near-IR light curves of GJ 1132 compared to the JKTEBOP best fits. The data have been binned by a factor of five and plotted versus orbital phase in order to make the plot clearer. Labels give the source and passband for each dataset. The polynomial baseline functions have been removed from the data before plotting. The residuals of the fits are plotted at the base of the figure, offset from unity. the fitting of limb darkening. For each light curve we determined the best-fitting rb, and obtained its uncertainty via 1000 Monte Carlo simulations. The optical transmission spectrum is the most important result of the current work, so we have deployed a suite of tests to probe the reliability of the radius values and errorbars 8 Figure 4. Transmission spectrum of GJ 1132 b: the measured plane- tary radius (Rb) as a function of the central wavelength of the pass- bands used. The passband names are given at the top of the plot. The horizontal lines indicate the FWHM of the passband used and the vertical lines show the relative errorbars in the Rb measurements. Filled circles show results from GROND data and open circles those from PISCO data. Figure 5. Results of testing the measurements of the planetary radius (Rb) in the GROND griz passbands. The vertical lines show the relative errorbars in the Rb measurements. This figure shows the different measurements of Rb obtained whilst performing several tests on the robustness of the results. The results from the various tests are offset in wavelength for clarity. found for each passband. These have been applied to the optical bands only, as it is here that the measurements are most precise and extensive, and because the optical bands are the most affected by issues such as limb darkening and contaminating light. 6.1.1. Sensitivity to individual transit observations A cursory inspection of Fig. 1 reveals that the transit depth in each passband may vary between different transits. The most obvious indicators are an unusually shallow transit in g on 2016/03/26 and an apparently very deep transit in z on Figure 6. Measured planetary radius for each individual transit in the GROND griz bands, with relative errorbars. The horizontal line shows the overall radius measurements obtained in Section 5. 2016/05/17. This possibility can be probed by determining a set of solutions in each passband, each one based on the full dataset but for one transit omitted. We performed this analysis for the GROND data and plot the results in Fig. 5. The PISCO data were not considered: they are unsuitable for this analysis because they cover only one transit. We found that the planetary radius measurements were not significantly affected by the omission of individual transit datasets: all results exist within the 1σ errorbars found for our default solution. We conclude that our results are ro- bust against the omission of individual light curves. This is a particularly powerful test in the current case, because of the large number of datasets obtained in each band, and makes us confident in the reliability of both the measured values and errorbars of the optical transmission spectrum. We also modelled every optical transit individually and show the resulting Rb values in Fig. 6. The z-band values of Rb show a clear increase relative to those for other bands, peaking at a 4.4σ difference between i and z. Even if we reject the highest radius measurement, the last one in the z band, we obtain a difference in Rb between the i and z bands significant at the 3.1σ level. The measurements exhibited in Fig. 6 are tabulated in Table 6. 6.1.2. Treatment of limb darkening Our default approach to obtaining the transmission spec- trum was to model limb darkening using the quadratic law, with the linear coefficient included as a parameter of the fit and the nonlinear coefficient fixed to a value interpolated from theoretical predictions by Claret (2004). This causes a dependence on theoretical model atmospheres, which is an important consideration for a planet host star as cool as GJ 1132 A. We therefore calculated the planetary radius mea- surements in the GROND and PISCO bands using an alter- nate approach: fixing both limb darkening coefficients. Table 6. Measured planetary radius for each individual transit in the GROND griz bands, with relative errorbars, as plotted in Fig. 6. 9 Observing night Measured planetary radius ( R⊕) g r i z 2016/01/14 1.209 ± 0.154 1.393 ± 0.125 1.420 ± 0.062 1.459 ± 0.105 2016/02/14 1.475 ± 0.091 1.302 ± 0.110 1.422 ± 0.051 2016/02/24 1.567 ± 0.151 1.302 ± 0.096 1.151 ± 0.091 2016/02/27 1.318 ± 0.122 1.417 ± 0.086 1.376 ± 0.063 1.643 ± 0.133 2016/03/26 1.221 ± 0.120 1.441 ± 0.083 1.255 ± 0.064 2016/03/29 1.457 ± 0.228 1.389 ± 0.122 1.468 ± 0.101 1.715 ± 0.141 2016/04/03 1.515 ± 0.146 1.447 ± 0.175 1.293 ± 0.089 1.731 ± 0.231 2016/04/08 1.570 ± 0.121 1.519 ± 0.133 1.309 ± 0.063 1.445 ± 0.159 2016/05/17 1.255 ± 0.183 1.563 ± 0.100 1.492 ± 0.087 1.998 ± 0.150 Figure 7. Planetary radius measurements from individual transits. The upper panel shows results for fitting the g, r and i-band data together. The lower panel includes also the z-band data, which are available for six of the nine transits. The errorbars include only those contributions specific to individual transits, and do not include sources of uncertainty common to all transits. The dotted line shows the overall measured value of Rb = 1.43 R⊕ found in Section 5. Fig. 5 shows the effect on the results for the GROND bands; the PISCO bands show a similar effect. We find that fixing both limb darkening coefficients causes the solution to move to higher planetary radius. Whilst the effect is simi- lar for all four bands, it affects r and i more than g and z. It therefore modifies the transmission spectrum at the 1–2σ level, in particular the slope seen through the optical wave- length region. As the g and z bands are relatively unaffected, limb darkening cannot explain the anomalously large radius we find in the z band. However, we conclude that observed transmission spectra can be significantly affected by the treat- ment of limb darkening and urge future studies to check for this possibility in all cases. 6.1.3. Effect of contaminating light Our default solution includes a third-light contribution from a nearby star. This phenomenon can have a signif- icant effect on the transmission spectrum when the spec- tral energy distribution of the target and contaminant differ (Southworth & Evans 2016), as is the case here. We tested the importance of this effect by calculating solutions for the GROND data whilst neglecting the third-light contribution. Once again, the results are represented in Fig. 5 and show that the contamination from the nearby star has a negligible effect on the results. The effect is 0.12σ for the g band and smaller for the other bands. The PISCO data will be similarly (un)affected. We conclude that our transmission spectrum measurements are robust against an amount of contaminat- ing light significantly in excess of that currently known for GJ 1132. 6.1.4. Temporal variability We modelled the light curves from all passbands but for each transit individually, in order to check for the presence of temporal variability in the planet. To deal with the com- plication of having z-band data available for only a subset of the transits, we obtained results for the g, r and i bands to- gether, and also for all four optical bands when possible. Due to the scatter in the data we fixed the limb darkening coeffi- cients for this analysis, using the linear coefficients 0.39 for gri and 0.31 for griz, and the quadratic coefficients 0.45 for gri and 0.51 for griz. 10 Our results are shown in Fig. 7. There is no overall trend in measured planet radius during our observations. 6.1.5. Stellar activity Low-mass stars frequently show dark starspots, which are capable of modifying the transit depth. Starspots occulted by the planet will make the transit shallower, and unoc- culted starspots will make it deeper (e.g. Ballerini et al. 2012; Tregloan-Reed et al. 2013; Oshagh et al. 2013). GJ 1132 A is a relatively old star so will not show strong activity, but the fact that a rotation period has been observed (125 d, BT15) means that it does show some starspots. However, there is evidence that many M dwarfs show a large number of small spots (Jackson & Jeffries 2012, 2013), which would greatly reduce the effect of spot activity because the occulted part of the stellar surface would be very similar to the unocculted ar- eas. This is consistent with the lack of starspot features seen in transits of planets in front of M dwarfs (e.g. Mancini et al. 2014; Awiphan et al. 2016). The very small radius of GJ 1132 b to that of GJ 1132 A means that unocculted starspots would be the dominant con- tributor to transit depth variations for this system. Because starspots are cooler than the rest of the stellar surface, un- occulted spots would have a wavelength-dependent effect on the transit depth. They would cause transits to become grad- ually deeper as one observed at bluer wavelengths. This does not provide an explanation for the current case, where the z- band radius is significantly larger than the gri-band points. For completeness, we note that unocculted plage could have the opposite effect (see also Oshagh et al. 2014), but this has never been observed. It would also require the plage to have a very clumpy distribution in order to not be occulted dur- ing the transits we observed, which is not the situation seen for the Sun, and would not cause an abrupt change in transit depth as found in the z-band. Finally, we note that our observations of GJ 1132 were distributed over 124 d, and therefore fortuitously sample the 125 d rotation period of GJ 1132 A very well. Moreover, ob- servations in the different passbands were obtained simulta- neously so temporal variations are unable to affect the rela- tive transit depth measurements. We therefore conclude that our observations are not significantly affected by spot activity in the host star. Having checked for and ruled out issues due to individual transits, contaminating light, planetary and stel- lar variability, we conclude that our transmission spectrum is robust. 7. CONSTRAINTS ON THE INTERIOR AND ATMOSPHERIC COMPOSITION OF THE PLANET Our multi-band photometric observations allow us to place joint constraints on the interior and atmospheric composition of GJ 1132 b. As discussed above our observations include measurements of transit depth, and hence the planetary ra- dius, in seven photometric bands. As shown in Fig. 4, six of these seven measurements are consistent with a "surface radius" of 1.375 R⊕ to within the ∼2σ uncertainties. By "surface radius" we mean the smallest radius which is ob- servationally accessible. However, the measurement in the z-band at 0.9 µm differs from a flat spectrum by over 4σ, thereby making it highly statistically significant, and sug- gests a potential contribution from the planetary atmosphere. Additionally, the z-band also overlaps with a strong H2O ab- sorption band whereas the remaining six bands probe win- dows in the H2O opacity. Therefore, it is possible that the latter six bands are all measuring the radius of the planetary "surface" while the z-band is probing a potentially H2O-rich atmosphere contributing to a higher measured radius. The importance of using radii measured in different bands to rep- resent the interior versus atmosphere has been suggested pre- viously by Madhusudhan & Redfield (2015), an effect which we are likely witnessing in the present case. Therefore, in what follows, we use a combination of interior and atmo- spheric models to jointly interpret the data. The observed mass and radius of GJ 1132 b allow us to in- vestigate the possible interior composition of the planet us- ing internal structure models. The planet mass as shown in Table 4 is 1.63 ± 0.54 M⊕. For the radius, we use the baseline value of 1.375 ± 0.16 R⊕ discussed above to repre- sent the bulk 'grey' radius of the planet. Fig. 8 shows model mass-radius curves of homogeneous super-Earths of differ- ent compositions spanning Fe, silicates, and H2O. The mod- els are described in Madhusudhan et al. (2012). We find that the mass and radius are consistent with two broad composi- tional regimes. Firstly, an exactly Earth-like composition, with 33% iron, 67% silicates and no volatile layer, is in- consistent with the data within the 1σ uncertainties. But, a composition with higher silicate-to-iron fraction, includ- ing a pure silicate planet, is ostensibly consistent with the data, albeit marginally. On the other hand, the data are also consistent with a large range of H2O mass fractions be- tween 0% and 100% in our models. In principle, consid- eration of temperature-dependent internal structure models would lead to larger model radii for the same composition (Thomas & Madhusudhan 2016) and therefore could lower the upper limit on the water mass fraction. Nevertheless, the mass and radius of GJ 1132 b allow for a degenerate set of solutions ranging between a purely silicate bare-rock planet and an ocean planet with a substantial H2O envelope. The degeneracy between the two scenarios could potentially be resolved using spectroscopic observations of the planetary at- mosphere, as discussed below. Considering the transmission spectrum of the planet, we report a tentative inference of H2O in the planetary atmo- sphere which provides initial signs of a water-rich world. Fig. 9 shows the data and model transmission spectra of GJ 1132 b with different H2O mixing ratios in a H2-rich at- mosphere; other compositions are explored in the following section. The data are inconsistent with a flat spectrum by 11 Figure 8. Mass-radius plot showing the properties of GJ 1132 b as well as values for other planets taken from literature sources and compiled in TEPCat (Southworth 2011). All planets in this mass range are included if their mass is measured to 3σ significance (i.e. the upper and lower errorbars on their mass measurements are both less than one third of the mass value). The names of individual planets are noted, and "K" has been used as shorthand for "Kepler" to aid the clarity of the plot. The coloured curves show the mass-radius relation for planets composed of pure water (blue), enstatite (green) and iron (brown). The position of the Earth is shown with a ⊕ symbol. over 4σ. As discussed above, the key constraint on the at- mospheric composition of the planet is governed by the z- band measurement which shows a substantially higher tran- sit depth relative to the baseline. We model the atmospheric transmission spectrum of the planet using the exoplanetary atmospheric modelling method of Madhusudhan & Seager (2009) and Madhusudhan (2012). Given the limited num- ber of datapoints we did not embark on a full retrieval exer- cise, but instead systematically investigated a grid of model compositions. In this section, we consider models compris- ing only gaseous H2 and H2O at different mixing ratios to explore the potential contribution of H2O to the z-band mea- surement. We nominally fixed the temperature structure to be isothermal at 600 K, representative of the equilibrium tem- perature of the planet; nominal variations in the temperature don't change our conclusions significantly. We find that the best model fits to the data, particularly the z-band point, are obtained for H2O volume mixing ratios of 1–10%, as shown in Fig. 9. On the other hand, the remaining data which are consistent with a flat spectrum are fit relatively easily for a wide range of models, as the corresponding photometric bands largely probe windows in opacity. The amplitude of the observed z-band feature is phys- ically plausible for a range of compositions. The differ- ence between the z-band radius and the continuum provides an estimate of the thickness of the observable atmosphere (Hatm) to be 0.22 ± 0.06 R⊕ or ∼1400 ± 400 km. Con- sidering an isothermal atmosphere at the equilibrium tem- perature (644 K) the atmospheric scale height (Hsc) for a H2-dominated atmosphere (mean molecular mass µ = 2) is ∼300 km. The value of Hsc for 10% H2O (µ = 3.6) and 20% H2O (µ = 5.2) is 167 km and 115 km, respectively, as- suming H2 occupies the remaining fraction. Therefore, the atmospheric height can be explained by ∼5 scale heights of an H2-dominated atmosphere given a strong absorber in the z-band. Similarly, Hatm can also be explained by .10 scale heights of a 10% H2O atmosphere. Conversely, considering ∼8–10 scale heights expected for a saturated spectral feature (Madhusudhan & Redfield 2015) at the same temperature, the mean molecular mass of the atmosphere is constrained to be µ ∼ 2.8–5.5. Such a µ is possible in a H2-rich at- mosphere with ∼10–20% H2O or corresponding fractions of 12 Figure 9. Comparison between the observed transmission spectrum of GJ 1132 b and theoretical spectra for a range of H2O volume mixing ratios (coloured lines) in a H2-dominated atmosphere. The red points show results from our GROND observations and the coloured circles indicate the band-integrated values of the theoretical spectra. The dashed black line shows the baseline radius of the planet. other heavy molecules, again assuming that a strong absorber in the z-band is present in the atmosphere. Future observa- tions with HST and JWST in the near-infrared could further constrain the presence and composition of such an absorber. The combined interior and atmosphere models fitting the data are consistent with a water-rich envelope in GJ 1132 b. And, given that the data are inconsistent with a flat spectrum the atmosphere is unlikely to be of very high mean molec- ular mass, e.g. 100% H2O or CO2, or one with thick high altitude clouds. However, since the inference is based pri- marily on one photometric datapoint (albeit obtained from light curves of six transits), future observations are critical to validate our current finding. We predict that a transmis- sion spectrum of GJ 1132 b observed with HST/WFC3 in the 1.1–1.8 µm spectral range should be able to detect the wa- ter absorption in its atmosphere. On the contrary, if such a HST spectrum does not detect an H2O feature then it might suggest the presence of another molecule in the atmosphere which we have not accounted for in our present models, or the possibility of time-variable events in the planetary atmo- sphere. It is also necessary that additional observations of the planetary transit in the z-band be conducted to bolster the current finding. 7.1. Transmission spectra calculations using petitCODE Using the newest version of the petitCODE (Molli`ere et al. 2015, 2016), we performed a second, independent, explo- ration of the atmospheric properties of the planet. petitCODE self-consistently calculates the radiative-convective equilib- rium structures of irradiated or self-luminous exoplanet at- mospheres. Molecular and atomic line opacities, cloud opac- ities, and H2–H2 and H2–He collision induced absorption (CIA) are taken into account. The code also includes molec- ular Rayleigh and cloud particle scattering. For these calculations we used the planetary parameters as defined in Table 4. We assumed a stellar effective tem- perature of 3270 K from BT15. We calculated our standard suite of models as defined in Molli`ere et al. (2016) for this planet, consisting of one fiducial model without clouds, two models at half and twice the solar C/O ratio, respectively, and nine different cloud model approaches as defined in ta- ble 2 of Molli`ere et al. (2016). For all models (three clear and nine cloudy) we calculated atmospheric structures and spectra at four different scaled-solar chemical compositions, corresponding to atmospheric metallicities of (cid:2) Fe H (cid:3) = 0, 1, 2 and 3. We considered Na2S and KCl clouds only, because higher- temperature cloud species can most likely not be mixed up to the higher layers of the atmosphere (Charnay et al. 2015; Parmentier et al. 2016). Cloud models 1 to 4 repre- sent models of varying cloud thickness using the model by Ackerman & Marley (2001) and adopting fsed values of 0.01 to 3, where fsed is the ratio of the particle-mass-averaged settling speed and the atmospheric mixing velocity. Cloud models 5 to 9 represent extended clouds with mono-dispersed size distributions. The particle size is fixed at 0.08 µm, which leads to Rayleigh scattering of the clouds in the optical. The cloud mass fraction is equal to the mass fraction derived 13 Figure 10. Comparison between the transmission spectrum of GJ 1132 b observed using GROND, and theoretical spectra from petitCODE. Black points show the measured values and coloured points the band-integrated theoretical values of the planetary radius. The top panel shows the four investigated values of (cid:2) Fe H (cid:3) = 2 and different choices of C/O ratio or treatment of cloud. H (cid:3), and the bottom panel shows models with (cid:2) Fe from equilibrium chemistry, but not larger than a maximum value Xmax, which decreases from model 5 to 7. Xmax can thus be thought of as a simple parametrisation of the settling strength. Cloud models 8 and 9 adopt the same parameter choice as cloud model 6, but additionally include iron clouds (model 8) or a spherical, homogeneous cloud particle shape (model 9). The cloud models therefore cover the parame- ter space of larger-particle clouds (models 1 to 4), leading to flatter transmission spectra in the optical, to small-particle clouds (models 5 to 9) which lead to Rayleigh scattering. Model 8 will not exhibit Rayleigh scattering if iron can con- dense within the atmosphere, due to the strong optical ab- sorption of iron. Each of the model spectra was compared to the observed transmission spectrum and the level of agreement deter- mined. This was done by identifying the lowest χ2 value between the observed transmission spectra and passband- integrated model values from petitCODE, whilst allowing for an overall shift in radius because the spectra were calcu- lated assuming a planetary base radius of 1.43 R⊕ at 10 bar pressure. Faced with a multitude of choices over which data to consider, we defaulted to using the GROND griz points with one limb darkening coefficient fitted and the GROND J HK points with fixed limb darkening coefficients. This is because there is a disagreement between the planet radius in the GROND and PISCO g bands, and the former was ob- tained from nine transits versus the single transit for the latter. We prefer results obtained with fitted limb darkening coeffi- cients because of limitations in the theoretical understanding of the atmospheres of stars as cool as GJ 1132 A. Alterna- tive choices of datapoints will be discussed below. Selected models are shown in Fig. 10. We found that the best fit to the GROND grizJ HK trans- mission spectrum is obtained for the petitCODE model with (cid:2) Fe H (cid:3) = 2 and a twice-solar or solar C/O ratio (χ2 = 18.6 and 19.1, respectively, for seven datapoints) although accept- 14 able agreement also occurs for the models with (cid:2) Fe H (cid:3) = 1 and (cid:2) Fe H (cid:3) = 3 (χ2 from 19.9 to 20.7 depending on C/O ra- tio). The models with (cid:2) Fe H (cid:3) = 3 have muted spectral features, because the high mean molecular mass leads to a low atmo- spheric scale height. They match the H-band radius better, at the expense of a poorer agreement with the z-band radius. A featureless spectrum (i.e. a straight line) is a worse fit with χ2 = 22.1. The χ2 values are all quite large, and reflect the difficulty of finding a good agreement between the ob- served and theoretical spectra. Including cloud in the models serves to increase the χ2 because the flattened spectral fea- tures match the observed transmission spectrum less well. If we restrict the analysis to the GROND griz datapoints, as these are the most reliable measurements, the models with (cid:2) Fe H (cid:3) = 1 and (cid:2) Fe H (cid:3) = 2 are the best match. The greatest agreements are for (cid:2) Fe H (cid:3) = 2 and a twice-solar C/O ratio (χ2 = 8.9 for four datapoints), followed by (cid:2) Fe H (cid:3) = 1 and a solar (χ2 = 10.0) or twice-solar (χ2 = 10.5) C/O ratio. The straight-line fit is clearly disfavoured with χ2 = 16.9. The slight preference for the twice-solar C/O ratio is because a higher C/O ratio enhances the amount of CH4, an effect which outweighs the loss of H2O and so leads to a larger radius in the z band. Even these models underpredict the detected variation of planet radius with wavelength, which also means a cloudy or featureless spectrum is strongly dis- favoured. We therefore conclude that our results are best explained by a clear atmosphere with (cid:2) Fe H (cid:3) = 2 and strong z-band opacity contributed by CH4 or H2O. Turning now to the transmission spectrum from the GROND data with all limb darkening coefficients fixed, we find that the best models have (cid:2) Fe H (cid:3) = 2, with χ2 = 19.4 (seven dataponts) for a solar C/O ratio, 19.5 for twice-solar C/O and 19.8 for half-solar C/O. (cid:2) Fe H (cid:3) = 3 with twice-solar C/O is nearly as good (χ2 = 20.2), as is (cid:2) Fe H (cid:3) = 1 with half- solar C/O (χ2 = 20.8). A straight-line fit has χ2 = 22.1. The quality of our data and limitations in the treatment of limb darkening means that we are able to infer that the at- mosphere of GJ 1132 b is best-represented by models with (cid:2) Fe H (cid:3) = 2, but the C/O ratio is not usefully constrained. 8. SUMMARY AND CONCLUSIONS GJ 1132 is a benchmark nearby system containing a low- mass planet transiting a late-M dwarf. We have presented extensive photometry of the system comprising light curves of nine transits observed simultaneously in the griz optical and J HK near-IR passbands. We have analysed these and literature data to determine the physical properties of the sys- tem. We find that the planet is larger than previously thought, 1.43 ± 0.16 R⊕ versus 1.16 ± 0.11 R⊕, from a methodolog- ical approach which relies more on observations of the sys- tem and less on empirical calibrations of the properties of low-mass stars. The planet's measured mass and radius are consistent within 1σ with theoretical predictions for a planet composed of silicates or water; a 100% iron composition gives a radius too small by ∼2σ. Our repeat observations allowed us to check for variability in the measured planet radius, and two of the transits do in- deed yield radii which are modestly discrepant with measure- ments from other transits. This could indicate excess scatter among the results, starspots on the stellar surface, unidenti- fied systematic effects in our data, or the presence of variabil- ity in the planet's atmosphere (e.g. Armstrong et al. 2016). We have constructed an optical-infrared transmission spec- trum of GJ 1132 b by modelling all light curves with a con- sistent geometry. We find an increased planet radius in the z band, to a significance level of 4σ, indicative of atmospheric opacity due to water, methane or another unidentified source. Detailed investigation of the resulting errorbars was enabled by the observation of nine transits. We find that our results are robust against the rejection of individual transits or the in- clusion of contaminating light from a nearby star. The treat- ment of limb darkening is more concerning, as it affects the results in the r and i bands at the level of 2.7σ and 3.5σ, respectively. We urge fellow researchers to consider this is- sue in similar analyses, especially for very cool stars where theoretical limb darkening coefficients are less reliable. The transmission spectrum was modelled using the atmo- spheric models of Madhusudhan & Seager (2009), with the finding that H2O likely causes the enlarged z-band radius of the planet. The best fits to the observations are found for H2O volume mixing ratios of 1–10%, implying a water-rich at- mospheric composition which would cause observable spec- tral features in a 1.1–1.8 µm transmission spectrum obtained using HST/WFC3. From simulations of the atmosphere of GJ 1132 b, Schaefer et al. (2016) found that the presence of H2O implied either an H2 envelope or low UV flux from the host star early in the lifetime of the system, and the ongoing presence of a magma ocean on the planet's surface. We also calculated theoretical spectra using the petitCODE (Molli`ere et al. 2015), which yield similar results except for the finding that the large z-band radius is explicable by an enhanced abundance of CH4. A high metallicity of (cid:2) Fe H (cid:3) = 2 is preferred, depending on the datapoints con- sidered, which is in line with the mass–metallicity correla- tion seen for more massive planets (Kreidberg et al. 2014a; Mordasini et al. 2016). A straight line is a much poorer fit to the transmission spectrum, confirming that we have detected the atmosphere of a 1.6 M⊕ planet. We advocate extensive further observations to refine and extend our understanding of the GJ 1132 system. High- precision optical light curves from large telescopes would be able to confirm or disprove the larger radius of the planet in the z-band, and shed light on the discrepancy seen in the g-band. Intermediate-band photometry at 900 nm or bluer than 500 nm would enable finer distinctions to be made be- tween competing model spectra and a clearer understanding of the chemical composition of the planetary atmosphere. The planet's mean density measurement is also hindered by the weak detection of the velocity motion of the host star, an issue which could be ameliorated with further radial velocity measurements using large telescopes. Finally, infrared transit photometry and spectroscopy should allow the detection of a range of molecules via the absorption features they imprint on the spectrum of the planet's atmosphere as backlit by its host star. Our results show that a 1.6 M⊕ planet with an equilib- rium temperature of 650 K is capable of retaining an exten- sive atmosphere. The atmosphere contains multiple molec- ular species and has likely persisted for many Gyr since the 15 formation of the system. ACKNOWLEDGEMENTS JS acknowledges financial support from the Leverhulme Trust in the form of a Philip Leverhulme Prize. We thank BT15 for making their photometric data available. The fol- lowing internet-based resources were used in research for this paper: the ESO Digitized Sky Survey; the NASA As- trophysics Data System; the SIMBAD database and VizieR catalogue access tool operated at CDS, Strasbourg, France; and the arχiv scientific paper preprint service operated by Cornell University. REFERENCES Ackerman, A. S., Marley, M. S., 2001, ApJ, 556, 872 Alonso, R., Barbieri, M., Rabus, M., Deeg, H. J., Belmonte, J. A., Almenara, J. M., 2008, A&A, 487, L5 Jao, W.-C., Henry, T. J., Subasavage, J. P., Brown, M. A., Ianna, P. A., Bartlett, J. L., Costa, E., M´endez, R. A., 2005, AJ, 129, 1954 Knutson, H. A., Benneke, B., Deming, D., Homeier, D., 2014a, Nature, Anglada-Escud´e, G., Rojas-Ayala, B., Boss, A. P., Weinberger, A. J., Lloyd, 505, 66 J. P., 2013, A&A, 551, A48 Armstrong, D. J., de Mooij, E., Barstow, J., Osborn, H. P., Blake, J., Saniee, N. F., 2016, Nature Astronomy, 1, 0004 Awiphan, S., et al., 2016, MNRAS, 463, 2574 Ballerini, P., Micela, G., Lanza, A. F., Pagano, I., 2012, A&A, 539, A140 Bean, J. L., Miller-Ricci Kempton, E., Homeier, D., 2010, Nature, 468, 669 Bean, J. L., et al., 2011, ApJ, 743, 92 Berta-Thompson, Z. K., et al., 2015, Nature, 527, 204 Biddle, L. I., et al., 2014, MNRAS, 443, 1810 Bonfils, X., et al., 2012, A&A, 546, A27 Cassan, A., et al., 2012, Nature, 481, 167 Charbonneau, D., et al., 2009, Nature, 462, 891 Charnay, B., Meadows, V., Leconte, J., 2015, ApJ, 813, 15 Chen, G., van Boekel, R., Madhusudhan, N., Wang, H., Nikolov, N., Knutson, H. A., et al., 2014b, ApJ, 794, 155 Kopparapu, R. K., et al., 2013, ApJ, 765, 131 Kreidberg, L., et al., 2014a, ApJ, 793, L27 Kreidberg, L., et al., 2014b, Nature, 505, 69 Lammer, H., Selsis, F., Ribas, I., Guinan, E. F., Bauer, S. J., Weiss, W. W., 2003, ApJ, 598, L121 Lanotte, A. A., et al., 2014, A&A, 572, A73 Madhusudhan, N., 2012, ApJ, 758, 36 Madhusudhan, N., Redfield, S., 2015, International Journal of Astrobiology, 14, 177 Madhusudhan, N., Seager, S., 2009, ApJ, 707, 24 Madhusudhan, N., Lee, K. K. M., Mousis, O., 2012, ApJ, 759, L40 Mancini, L., et al., 2014, A&A, 562, A126 Molli`ere, P., van Boekel, R., Dullemond, C., Henning, T., Mordasini, C., Seemann, U., Henning, T., 2014, A&A, 564, A6 2015, ApJ, 813, 47 Claret, A., 2004, A&A, 428, 1001 Claret, A., Hauschildt, P. H., Witte, S., 2012, A&a, 546, A14 Cloutier, R., Doyon, R., Menou, K., Delfosse, X., Dumusque, X., Artigau, ´E., 2017, AJ, 153, 9 Molli`ere, P., van Boekel, R., Bouwman, J., Henning, T., Lagage, P.-O., Min, M., 2016, A&A, in press, arXiv:1611.08608 Monet, D. G., et al., 2003, AJ, 125, 984 Mordasini, C., van Boekel, R., Molli`ere, P., Henning, T., Benneke, B., Cunha, D., Correia, A. C. M., Laskar, J., 2015, International Journal of 2016, ApJ, 832, 41 Astrobiology, 14, 233 de Mooij, E. J. W., et al., 2012, A&A, 538, A46 Delfosse, X., Forveille, T., S´egransan, D., Beuzit, J.-L., Udry, S., Perrier, Morton, T. D., Swift, J., 2014, ApJ, 791, 10 Muirhead, P. S., et al., 2015, ApJ, 801, 18 Nascimbeni, V., Piotto, G., Pagano, I., Scandariato, G., Sani, E., Fumana, C., Mayor, M., 2000, A&A, 364, 217 M., 2013, A&A, 559, A32 Demory, B.-O., et al., 2016, Nature, 532, 207 Dittmann, J. A., Irwin, J. M., Charbonneau, D., Berta-Thompson, Z. K., Newton, E. R., 2016, ApJ, submitted, arXiv:1611.09848 Dragomir, D., Benneke, B., Pearson, K. A., Crossfield, I. J. M., Eastman, J., Barman, T., Biddle, L. I., 2015, ApJ, 814, 102 Dragomir, D., et al., 2013, ApJ, 772, L2 Dressing, C. D., Charbonneau, D., 2015, ApJ, 807, 45 Eastman, J., Siverd, R., Gaudi, B. S., 2010, PASP, 122, 935 Fukugita, M., Ichikawa, T., Gunn, J. E., Doi, M., Shimasaku, K., Schneider, D. P., 1996, AJ, 111, 1748 Gaidos, E., Mann, A. W., Kraus, A. L., Ireland, M., 2016, MNRAS, 457, 2877 Gillon, M., et al., 2007, A&A, 472, L13 Gillon, M., et al., 2016, Nature, 533, 221 Greiner, J., et al., 2008, PASP, 120, 405 Hartman, J. D., et al., 2015, AJ, 149, 166 Hilditch, R. W., 2001, An Introduction to Close Binary Stars, Cambridge University Press, Cambridge, UK Jackson, R. J., Jeffries, R. D., 2012, MNRAS, 423, 2966 Jackson, R. J., Jeffries, R. D., 2013, MNRAS, 431, 1883 Oshagh, M., Santos, N. C., Boisse, I., Bou´e, G., Montalto, M., Dumusque, X., Haghighipour, N., 2013, A&A, 556, A19 Oshagh, M., Santos, N. C., Ehrenreich, D., Haghighipour, N., Figueira, P., Santerne, A., Montalto, M., 2014, A&A, in press, arXiv:1407.2066 Parmentier, V., Fortney, J. J., Showman, A. P., Morley, C., Marley, M. S., 2016, ApJ, 828, 22 Rogers, L. A., 2015, ApJ, 801, 41 Schaefer, L., Wordsworth, R., Berta-Thompson, Z., Sasselov, D., 2016, ApJ, 829, 63 Seager, S., Mall´en-Ornelas, G., 2003, ApJ, 585, 1038 Shields, A. L., Ballard, S., Johnson, J. A., 2016, arXiv:1610.05765 Southworth, J., 2008, MNRAS, 386, 1644 Southworth, J., 2010, MNRAS, 408, 1689 Southworth, J., 2011, MNRAS, 417, 2166 Southworth, J., 2013, A&A, 557, A119 Southworth, J., 2017, PASP, 129, 024401 Southworth, J., Evans, D. F., 2016, MNRAS, 463, 37 Southworth, J., Maxted, P. F. L., Smalley, B., 2004, MNRAS, 351, 1277 Southworth, J., Mancini, L., Maxted, P. F. L., Bruni, I., Tregloan-Reed, J., Barbieri, M., Ruocco, N., Wheatley, P. J., 2012, MNRAS, 422, 3099 16 Southworth, J., et al., 2009, MNRAS, 396, 1023 Southworth, J., et al., 2014, MNRAS, 444, 776 Sozzetti, A., Torres, G., Charbonneau, D., Latham, D. W., Holman, M. J., Winn, J. N., Laird, J. B., O'Donovan, F. T., 2007, ApJ, 664, 1190 Stetson, P. B., 1987, PASP, 99, 191 Thomas, S. W., Madhusudhan, N., 2016, MNRAS, 458, 1330 Tregloan-Reed, J., Southworth, J., Tappert, C., 2013, MNRAS, 428, 3671 Tsiaras, A., et al., 2016, ApJ, 820, 99 Van Grootel, V., et al., 2014, ApJ, 786, 2 Weiss, L. M., Marcy, G. W., 2014, ApJ, 783, L6 Winn, J. N., et al., 2011, ApJ, 737, L18
1111.3628
1
1111
2011-11-15T20:35:23
The Cratering History of Asteroid (21) Lutetia
[ "astro-ph.EP" ]
The European Space Agency's Rosetta spacecraft passed by the main belt asteroid (21) Lutetia the 10th July 2010. With its ~100km size, Lutetia is one of the largest asteroids ever imaged by a spacecraft. During the flyby, the on-board OSIRIS imaging system acquired spectacular images of Lutetia's northern hemisphere revealing a complex surface scarred by numerous impact craters, reaching the maximum dimension of about 55km. In this paper, we assess the cratering history of the asteroid. For this purpose, we apply current models describing the formation and evolution of main belt asteroids, that provide the rate and velocity distributions of impactors. These models, coupled with appropriate crater scaling laws, allow us to interpret the observed crater size-frequency distribution (SFD) and constrain the cratering history. Thanks to this approach, we derive the crater retention age of several regions on Lutetia, namely the time lapsed since their formation or global surface reset. We also investigate the influence of various factors -like Lutetia's bulk structure and crater obliteration- on the observed crater SFDs and the estimated surface ages. From our analysis, it emerges that Lutetia underwent a complex collisional evolution, involving major local resurfacing events till recent times. The difference in crater density between the youngest and oldest recognized units implies a difference in age of more than a factor of 10. The youngest unit (Beatica) has an estimated age of tens to hundreds of Myr, while the oldest one (Achaia) formed during a period when the bombardment of asteroids was more intense than the current one, presumably around 3.6Gyr ago or older.
astro-ph.EP
astro-ph
The Cratering History of Asteroid (21) Lutetia S. Marchia,∗, M. Massironib, J.-B. Vincentc, A. Morbidellia, S. Mottolad, F. Marzarie, M. Kuppersf, S. Besseg, N. Thomash, C. Barbierii, G. Nalettoj, H. Sierksc aDepartement Cassiop´ee, Universite de Nice - Sophia Antipolis, Observatoire de la Cote d'Azur, CNRS, Nice, France bDepartment of Geosciences, Padova University, Italy cMax Planck Institute for Solar System Research, Lindau, Germany dInstitut fur Planetenforschung, DLR-Berlin, Germany eDepartment of Physics, Padova University, Italy fESA-ESAC, Villanueva de la Canada (Madrid), Spain gLaboratoire d'Astrophysique de Marseille, France hPhysikalisches Institut, University of Bern, Switzerland iDepartment of Astronomy, Padova University, Italy jDepartment of Information Engineering, Padova University, Italy Abstract The European Space Agency's Rosetta spacecraft passed by the main belt asteroid (21) Lute- tia the 10th July 2010. With its ∼ 100 km size, Lutetia is one of the largest asteroids ever imaged by a spacecraft. During the flyby, the on-board OSIRIS imaging system acquired spectacular im- ages of Lutetia's northern hemisphere revealing a complex surface scarred by numerous impact craters, reaching the maximum dimension of about 55 km. In this paper, we assess the cratering history of the asteroid. For this purpose, we apply cur- rent models describing the formation and evolution of main belt asteroids, that provide the rate and velocity distributions of impactors. These models, coupled with appropriate crater scaling laws, allow us to interpret the observed crater size-frequency distribution (SFD) and constrain the cratering history. Thanks to this approach, we derive the crater retention age of several re- gions on Lutetia, namely the time lapsed since their formation or global surface reset. We also investigate the influence of various factors -like Lutetia's bulk structure and crater obliteration- on the observed crater SFDs and the estimated surface ages. From our analysis, it emerges that Lutetia underwent a complex collisional evolution, involv- ing major local resurfacing events till recent times. The difference in crater density between the youngest and oldest recognized units implies a difference in age of more than a factor of 10. The youngest unit (Beatica) has an estimated age of tens to hundreds of Myr, while the oldest one (Achaia) formed during a period when the bombardment of asteroids was more intense than the current one, presumably around 3.6 Gyr ago or older. Keywords: Asteroid (21) Lutetia, Asteroid cratering, Asteroid evolution, Main Belt Asteroids ∗Corresponding author Email address: [email protected] (S. Marchi) Preprint submitted to Planetary and Space Science May 23, 2018 1. Introduction The European Space Agency's (ESA) Rosetta spacecraft passed by the main belt asteroid (21) Lutetia with a relative velocity of ∼ 15 km/s on 10 July 2010 at 15:44:56 UTC. The Rosetta- Lutetia distance at closest approach (CA) was 3170 km. During the flyby the solar phase angle (sun-object-observer) decreased from the initial 11◦ to a minimum of 0.15◦ 18 minutes before CA, then increased again to 80◦ at CA and finally reached a maximum of 139◦ when the ob- servations were stopped. A total of 400 images were obtained by the Optical, Spectroscopic, and Infrared Remote Imaging System (OSIRIS), which consists of two imagers: the Wide Angle Camera (WAC) and the Narrow Angle Camera (NAC) [10]. The best resolution at CA corre- sponded to a scale of 60 m/px at the asteroid surface. Lutetia has an orbital semi-major axis of about 2.43 AU, an eccentricity of 0.16 and an inclina- tion of 3.06◦. Its shape can be fitted by an ellipsoid having axes of 121 × 101 × 75 km [22]. Previous space missions have visited and acquired detailed data for a total of 6 asteroids, namely four main belt asteroids [951 Gaspra, 243 Ida, 253 Mathilde, 2867 Steins; 23, 2, 1, 9] and two near-Earth objects [433 Eros, 25143 Itokawa; 24, 21]. Itokawa is the smallest of them, with dimensions of 0.45 × 0.29 × 0.21 km. The other asteroids have average sizes ranging from ∼ 5 km to ∼ 53 km. In this respect, Lutetia with its average size of 98 km is the second largest asteroid ever visited by a spacecraft so far (at the moment of the writing -October 2011-, Dawn mission is orbiting around the 500-km sized asteroid (4) Vesta). This paper analyzes some of the highest resolution OSIRIS images with the aim to study the crater size-frequency distributions (SFDs) on the different units that have been identified on the basis of geological investigations [22, 16, 26]. This analysis provides constraints on Lutetia's bulk structure and surface evolution. The observed crater SFDs are also used to compare the cratering process among the different units, to derive absolute ages and provide a chronology of the major events that affected Lutetia evolution. 2. Lutetia crater population The NAC high resolution images acquired during the flyby where used to identify major re- gions on Lutetia (see Fig. 1). These regions have been defined by taking into account several factors, including local topography, geological features, surface texture, crater spatial density and stratigraphic relationships [26, 16]. In this respect, each region is characterized by distinct properties of one or more of the above listed factors. The regions indentified have been further divided into several units. Thanks to this selection criterion, the defined units reflect major dif- ferences in their evolution [16, 26]. Note that the actual unit boundaries are in some cases not well established due to the lack of resolution and/or unfavorable illumination conditions [16, 26]. Among the major regions, only 4 were imaged with enough quality for accurate crater count- ing to be performed. These are Achaia, Narbonensis, Noricum and Baetica. Their geological properties show remarkable differences, therefore they will described individually in the follow- ing sections. 2 Achaia. This region is defined by a remarkably flat and uniform area. It is bounded by Baet- ica, Narbonensis and Etruria. Its boundaries with Baetica and Narbonensis are defined mainly by texture and topography, respectively. The boundary with Etruria is defined by the same means but, due to low contrast of the images in these regions, it is less precisely established1. The illumination conditions within Achaia are very good and uniform, therefore craters are clearly visible and their size estimate is performed with precision [25]. The Achaia region (Ac1+Ac2) is heavily cratered, showing a large range of crater sizes, from 21.6 km (Nicea crater) down to the resolution limit (we used a minimum of 4 pixels to identify craters, thus about 0.2 km). The overall spatial distribution of the 157 craters > 0.6 km is uni- form and there appears to be no evident contamination from adjacent units (see Fig. 2, panel b). At smaller sizes, several crater-like features may not be of impact origin. Many circular depres- sions are close to, or overlap linear features, therefore may not represent bona fide craters. The presence of secondary craters (formed by boulders ejected during the formation of other craters) can also be possible at these small crater sizes, although it is unclear how likely can secondary craters form on Lutetia, given its low escape velocity. For the purpose of age assessment, we are interested in primary craters (i.e. formed by im- pacts with asteroids), therefore our analysis focuses on craters > 0.6 km. The resulting crater SFD is shown in Fig. 3 (panel a). An interesting result is that Achaia's crater SFD exhibits a marked flexure point at about 4 − 7 km. Note that the observed flexure point is unlikely due to observational biases, like un- certainties in the identification of craters or resolution issues. This is because Achaia region is a remarkably flat area and it has been imaged with uniform conditions of illumination, while the flexure point is well above the image resolution. Moreover, thanks to the boundary selection, we also exclude that the observed flexure is due to obliteration of small craters due to crater ejecta coming from nearby units (e.g., Beatica). For the same reason, it seems also unlikely that the formation of the large crater Massalia (see next sections) played a role in the formation of the flexure point in Achaia crater SFD. Noricum. This unit has a very complex topography. It contains a number of closely packed and prominent circular features, likely impact craters, showing several stages of degradation [25]. Moreover, this unit looks "compressed" among the impact craters of Baetica, Massalia crater, and possibly another large crater on the dark side of Lutetia (namely, Pannonia region; see Fig. 1), the presence of which may be inferred thanks to the circular terminator of part of Noricum. These factors are likely at the origin of Noricum complex topography. Crater counts have been performed in unit Nr1+Nr2 (for semplicity we will refer to Noricum region in the rest of the work). The overall viewing geometry is not optimal (i.e. nearly edge- on), therefore the size estimate of some of the 76 identified craters (> 0.6 km) is problematic (see Fig. 2, panel a). The resulting crater SFD shows a clear transition at about 2 km (see Fig. 3, panel b): the slope of the crater SFD for D > 2 km is considerably shallower than that for D < 2 km. The feature resembles somewhat the flexure seen on Achaia crater SFD, although in this case it 1Note that several choices of the Etruria-Achaia boundary have been performed in our analysis. The influence on the actual choice on the resulting Achaia crater SFD is negligible. 3 may be due to imprecise size estimate for several large craters due to their nearly edge-on view. The crater spatial density for D < 2 km is very similar to that of Achaia. Narbonensis. This region corresponds to the interior of the 55-km-sized crater Massalia, the largest impact structure detected on Lutetia. Crater count has been performed in unit Nb1 (for simplicity we will refer to Narbonensis region in the rest of the work). A total of 47 craters > 0.6 km have been identified (see Fig. 2, panel c). Notably, several craters appear deformed by sliding of their rims due to the relatively high topographic slope present in large part of the unit [25, see also Fig. 4, upper panels]. In these cases, the determination of the actual crater size is not very accurate. Overall, the crater spatial density of Narbonensis is lower than that of Achaia (see Fig. 3, panel c). The shapes of the crater SFDs of the two units also differ. In particular, the Narbonensis crater SFD has a shallower slope at small sizes than Achaia. It is not clear whether this difference is due to poor count statistics or it is a real feature. In the latter case, it might be due to variation in the local properties of the terrains or due to some later modification (as we will discuss later). Baetica. This region, unlike the previous ones, shows marked evidence of several major modification processes (landslides, ejecta blanketing etc) that have been used to establish sub- units that likely formed at different epochs [26, 16]. Moreover, this region is also characterized by large topographical slope variations (from 0 to 45 deg, see Fig. 4, lower panels), and by the presence of many large boulders [11]. Overall, the Baetica region presents much fewer craters than adjacent regions. Some Baet- ica's units appear extremely young, showing no detectable impact craters. For these reasons, we restrict our analysis to a unit, named Bt1a (see Fig. 1), which apparently has not been affected by recent geological processes (e.g., landslides), it is relatively flat and uniform, and does contain a fair number of small impact craters. In this case, we boost crater detection by using Laplacian- filtered images2. We identify 62 craters in the range 0.2 − 1 km (see Fig. 2, panel d). The Bt1a crater SFD shows an overall shape consistent with those of other units, and it is characterized by a much lower crater spatial density (see Fig. 3, panel d). Interestingly, Bt1a contains a fresh and large (∼ 7 km) crater, plus a second highly degraded crater having similar dimensions that has not been counted since it probably formed before Bt1a [26, 25]. 3. The Model Production Function chronology The crater SFDs of the units presented in the previous section can be used to derive their crater retention ages. The age of units is crucial information, since it provides constraints on the formation and evolution of Lutetia. In this respect, Lutetia stands out with respect to all previ- ously visited asteroids (except Vesta), for its complex geological evolution. Therefore, the crater retention ages of its units are important to set a timeline for this evolution. Moreover, the study 2 The Laplacian filter technique uses secondary derivatives in two directions to enhance the contrast of the input image and it is known to be very effective in revealing small, high frequency features [3]. 4 of the cratering process along with geological assessment can be used to constrain the physical properties of the target. In this work, crater retention ages are derived in the framework of the Model Production Function (MPF) chronology [12]. With this approach Lutetia's crater production function (i.e., the expected number of craters per year per unit surface) is computed by modeling its impactor flux and by using a crater scaling law in order to compute the resulting crater population. The resulting crater MPF gives the cumulative density of craters (per year) as a function of the crater size. In analogy with previous work, the impactor flux is characterized by its size-frequency dis- tribution and impact velocity distribution. The impactor SFD is taken from the model population of main belt asteroids of [4]. In this work, we will also consider a second MBA population derived by the Sub-Kilometer Asteroid Diameter Survey (SKADS) [5, see Fig. 5]. Using the [7] algorithm, we computed that the intrinsic collision probability between MBAs and Lutetia is Pi = 4.21 · 10−18 km−2yr−1. Note that this Pi value is significantly higher that the average value for the main belt, namely 2.86 · 10−18 km−2yr−1. Using the same algorithm, we also computed the Lutetia's impact velocity distribution (see Figure 6). Concerning the crater scaling law, we adopted a Pi-group scaling law [8]. These scaling laws allow us to estimate the size of a crater given the dimension (d) and velocity (v) and den- sity (δ) of the impactor along with the density (ρ) and strength (Y) of the target. In addition to these quantities, two parameters (ν, µ) account for the nature of the terrains (hard rock, cohesive soil, porous material). In this paper, we investigate both hard rock and cohesive soils scaling laws, whose parameters are ν = 0.4, µ = 0.55 and ν = 0.4, µ = 0.41, respectively. We assume Y = 2 · 108 dyne/cm2 for typical hard rock and an impactor density of δ = 2.6 g/cm3 [13]. The bulk density of Lutetia is ρ = 3.4 g/cm3 [22]. Values of density and strength for cohesive soils will be given in Section 4. Further details about the crater scaling law can be found in [14]. Note that no correction for the transient-to-final crater size has been applied, because the crater modification stage is not likely to occur on Lutetia given its low gravity. Absolute ages can be computed by knowing the time dependence of the impactor flux in the past. Unfortunately, such time-dependence is not known for main belt asteroids. Two approaches can be used to overcome such a limitation. First, one can assume that the present impact rate for main belt asteroids remained constant over the age of the solar system. This scenario requires a constant main belt population, where no big modification (e.g., in its orbital architecture and total mass) occured. However, it is known that the main belt was more massive in the past and that during the early phases of the solar system it was shaped by major events [e.g., 17]. However, these processes have not yet been modeled with enough certainty and accuracy to enable the determination of the time evolution of the impact rate. An alternative approach is to refer to the lunar impactor flux, which has been calibraterd on the basis of radiometric ages of lunar samples [18, 12]. This scenario assumes that the impactor flux variation experienced by the Moon also applies to main belt asteroids. In reality, since the Moon is not embedded in the main belt, it is likely that the Moon and MBAs had very different impact histories. For instance, consider the case that the lunar impact cataclysm between 4.1 and 3.8 Gy ago was due to a temporary destabilization of the main belt that removed a part of its asteroids. Then the Moon would have suffered an impact spike, while the impact rate in the asteroid belt would have decreased (i.e., 5 without impact spike) from an initial higher but rougly constant value in the 4.5 − 4.1 Ga time- interval, to the current value. Thus, the time evolutions of the impact rate on the Moon and in the asteroid belt would have been totally different. On the other hand, in the case of a large cometary contribution to the lunar cataclysm (i.e., from a source region outside the main belt) these bodies would have produced an impact spike on both the Moon and MBAs. In the assumption that the evolution of the impact rate in the asteroid belt and on the Moon was the same, and assuming that all craters that are formed are retained on the surface, the crater MPF function for an asteroid at a time t is given by: MPF(D, t) = MPF(D, 1yr) · N1(t) N1(1yr) (1) where D is the crater size and N1(t) expresses the lunar crater cumulative number at 1 km as a function of time according to the following equation: N1(t) = a(ebt − 1) + ct (2) where t is in Gyr (t = 0 is the present time), a = 1.23 × 10−15, b = 7.85, c = 1.30 × 10−3 [12]3. Note that setting a = 0 would correspond to the constant flux scenario. The MPF(D, t) is used to derive the model cratering age by a best fit of the observed crater SFD that minimizes the reduced chi squared value, χ2 r . Data points are weighted according to their measurement errors. The formal errors on the best age correspond to a 50% increase of the χ2 r around the minimum value. Other sources of uncertainties are neglected [see 14, for more details]. Equation 1 basically implies that MPF(D, t) is obtained by simply y-axis shifting MPF(D,1yr) by a proper amount. It has been shown by previous studies on asteroid cratering, however, that several crater obliteration processes may be at work [e.g., 19]. In the case that crater oblitera- tion occurs, the shape of the MPF changes over time and may reach a steady-state in the case that crater saturation occurs (namely, the newly formed craters erase previous ones leaving the overall crater spatial density unchanged). In this paper, we take into account crater obliteration processes as described in [13]. 4. Crater retention age estimates One important aspect of MPF methodology is that it depends on the assumed properties of the target body [15, 14]. Therefore, the analysis of crater SFDs on different terrains on the same body (or different asteroids) should be done with caution, since changes in the material properties may invalidate direct comparison [14]. Generally speaking, material properties are not known in detail, however, in some cases, they can be constrained on the basis of geomorphological and geological analysis. Therefore, whenever possible, MPF chronology allows to derive cratering ages taking into account explicitly the effect of the inferred material properties. In this section 3The parameters a, b, c for the lunar chronology curve are determined by best fit of lunar calibration data and their actual values may vary according to different authors [e.g., 18]. However, the variation of the actual values for a, b, c has a negligible impact on the age determination. 6 we present the results of our MPF-based age estimate for each unit investigated. Achaia. As described in the previous section, Achaia crater SFD is characterized by a flexure point located at 4 < D < 7 km. Figure 7 reports the results of MPF best fitting of the observed crater SFD. The left panel shows the best fits obtained by using [4] population (P1 hereinafter) and the crater scaling law for hard rock both with and without crater obliteration. Concerning the crater obliteration process, we took into account local regolith jolting and crater superposition and adopted the same parameters used by [19]. Global seismic effects and cumulative seismic shaking have not been considered because of the large size of Lutetia4. The present fits are achieved anchoring the MPF to the large crater end of the crater SFD. The quality of the fit is ba- sically the same in the two scenarios, except for a slightly older age in case of crater obliteration. These results clearly show that P1 is not able to accurately reproduce the observed cratering. A similar conclusion is reached also using the [5] population (P2 hereinafter). In particular, the observed flexure in the crater SFD has no correspondence is either MBA populations. Indeed, the impactor population is not known at the impactor size relevant for the flexure (∼ 0.5 −0.8 km) and therefore it is possible that the real main belt SFD may account for it. Nevertheless, the fact that such a feature has not been observed on other large asteroids, like Ida and Mathilde [22], makes this unlikely. We have excluded that the flexure is due to the impactor flux, global and local obliteration processes, and observational biases (see also discussion in Section 2). A further possibility is that the flexure is related to terrain properties. As shown for Mercury [14], the presence of a strat- ified target having fractured material at the surface overlying a more competent interior would produce a crater SFD showing a characteristic flexure. Such a flexure is the combined result of i) adopting different material parameters for the fractured layer and the competent interior and ii) using cohesive soil and hard rock scaling laws for the two layers [14]. The position of the flexure is mainly determined by the thickness of the fractured material, which can be chosen in order to produce a best fit of the observed crater SFD. We investigated this possibility, by modeling a transition in the Achaia properties, as done in [14]. The results are shown in Fig- ure 7 (right panel). The P1 best fit is now improved, being in overall good agreement with the crater SFD. The resulting age is 3.6 ± 0.1 Ga, obtained for a fractured layer depth of 3 km. It must be clear that the above age derives from the lunar chronology (Equ. 1), whose applicatibily to main belt asteroids is unclear. It is also noteworthy that extrapolating the present main belt impact rate in the past would lead to an age older than that of the solar system. This suggests that the main belt experienced a heavy bombardment in the past, although not necessarily with the time-dependence described by equ. 1. The use of the lunar chronology probably provides a lower bound to the real age, whereas the age computed assuming a constant flux provides an upper bound (in this case a trivial one). The best fit presented in figure 7 shows a residual mismatch for craters 0.6 − 2 km (much above resolution limit), the origin of which is unclear. Here we show that, using a shallower MBA population -such as P2- would produce a better match of the observed crater SFD. The resulting age is 3.7 ± 0.1 Ga. Note, however, even in the presence of a shallower population a 4 To see this, we rescaled the Ida's global erasing curve from the Figure 4 of [19] to Lutetia. It results that a crater of about 100 km would be needed to globally erase craters ≥ 1 km. This result also suggests that it is unlikely that the formation of Massalia crater triggered global surface reset. 7 stratified target is needed in order to explain the flexure (Figure 7, left panel). It must be clear that the SKAD survey is valid down to an absolute magnitude of ∼ 18 (corresponding to a size of 0.8 km for a geometric albedo of 0.15). Such impactors would produce crater sizes of the order of several km, therefore in our fit we extrapolated P2 slope outside its range of validity. We also find that, independently of the MBA population used, the Achaia crater SFD is not saturated. Indeed, at least with the crater obliteration parameters adopted here, the saturation occurs at an higher crater density than observed on Achaia. However, we caution that this conl- cusion depends on the not-well-known process of crater obliteration. A more thorough analysis of this issue is deferred to future work. Baetica. This unit is characterized by the presence of a widespread regolith layer. The thick- ness of this layer is unknown, although both crater and landslide morphologies have been used to constrain its depth to be at least 100s of meters [25]. Therefore, it seems likely that all the craters (except maybe for the few largest ones) detected in Bt1a formed in highly granular, cohesive soils. As for the strength, reference values are from the lunar regolith (Y ∼ 3 · 104 at a depth of 3 meters) and terrestrial alluvium (Y ∼ 7 · 105). Here, we investigate strength values ranging from 105 to 107 dyne/cm2. We also take a density of 2 g/cm3, typical of lunar regolith. The resulting MPF best fit, using P1 population, is shown if Fig. 8. The main conclusion is that Bt1a is very young, ranging from ∼ 4 to ∼ 50 Ma, according to the value of the strength used. The same figure also shows the best fit achieved with P2. The quality of the fit is now much improved, given the SKADS' shallower SFD slope. In this case the derived ages range from ∼ 50 to ∼ 220 Ma. Note that the last age is in better agreement with the boulder lifetime estimated for the central (and youngest) unit of Baetica [11]. The overall wavy shape of the observed crater SFD is not accurately reproduced by the MPFs. This may have several explanations, including low crater statistics and a poor knowledge of the MBA SFD at these small sizes. Note that it is also possible, given the large topographical slopes present in this area, that small craters are not well preserved (see Fig. 4, lower panels). Noricum and Narbonensis. These two units present several difficulties in their age assess- ment. Both crater SFDs are not well fit by the MPFs, possibly because of errors in the crater size measurements (Noricum) and poor statistics (Narbonensis). Some constraints on the expected evolution of these units come for geological analysis. First of all, it is clear from stratigraphical arguments that Massalia crater formed later in time with respect to both Achaia and Noricum [16]. Therefore, the Narbonensis unit is younger than Achaia and Noricum. Moreover, in the light of the arguments discussed in previous sections, it appears difficult that the formation of Massalia globally reset Lutetia's surface. This conclusion is also in agreement with hydrocode simulations of Massalia formation. Note that these simulations [6] predict that the Massialia event triggered the formation of a fractured layer generated all over the surface of Lutetia. The actual damage of the fractured layer depends on the resolution of the simulations, nevertheless it is believed to be not sufficient to cause global resurfacing (K. Wunnemann pers. comm. on 27 June 2011)5. 5 We also acknowledge the fact that these conclusions are based on scaling laws and simulations which depend on 8 If the above scenario is correct, then we expect that Noricum has similar properties as Achaia. Figure 9 (left panel) shows the MPF best fit using a stratified target model and crater oblitera- tion. The best fit is achieved with a fractured layer depth of 1.3 km. The resulting Noricum age is ∼ 3.4 and ∼ 3.7 Ga for P1 and P2 respectively, which is consistent with being coeval with Achaia (again, we point out that these ages are derived using the lunar chronology). The rela- tively shallow layer of fractured material may also be consistent with the complex topography of the units (possibly reflecting a more competent near-surface interior). Figure 9 (right panel) shows the MPF best fit of Narbonensis. In this case, the observed crater SFD does not show evidences of a flexure, possibly due to the poor crater statistics, that would suggest the presence of stratified terrains. Nevertheless, according to [6], the interior of the Massalia crater is expected to be fractured up to the depth of several km. Therefore, by using the same crater scaling law and a fractured layer depth of 3.5 km (although larger depths are also possible), we obtain a best-fit age of ∼ 0.95 and ∼ 1.3 Ga for population P1 and P2, respectively. This latter result is puzzling, since such a large crater is not expected to be so young. Note that the inferred age is quite insensitive to the adopted scaling law or impactor population. We estimated that the impactor that formed Massalia was in the size range 7-9 km [22]. The current frequency of such impacts is about one every 9 Ga. The computed a priori probability that such an event happened in the last 1 Ga, is ∼ 11%. On the other hand, knowing that the Massalia event did happen within the last 4.5 Ga, the probability that such event occurred in the last 1 Ga, is ∼ 25%. These numbers apply for the present main belt impact rate, and thus certainly represent an upper limit because it is believed that the impact rate in the primordial main belt was at least a factor ∼ 2 − 4 more intense than today [17]. This would imply that Massalia event more likely happened early on rather than recently. Thus, in conclusion, it is likely that other processes may be responsible for a lack of craters within Narbonensis [16, 26]. As discussed in the previous section, this unit has relatively high topographical slopes and episodes of slopes slumping may have induced crater erasing (see Fig. 4, upper panels). The presence of significant rims slumping is supported by the V-shaped topographical profile of the crater [20]. Comparing the observed profile with a typical profile of a fresh crater [6], we derive that several 100s m of rim material may have been displaced toward the center of the crater, which may be enough to explain the relatively young age of this unit. 5. Discussions and conclusions The main result of our crater retention age analysis is that it confirms a prolonged and com- plex collisional evolution of Lutetia. As shown for previous asteroids visited by spacecraft, collisions play a major role in the evolution of any asteroid, being largely responsible of their shapes, internal structure and geomorpohological features. The latter play also an important role for the understanding of surface specrophotometric properties. All these collisional-related processes are well documented on Lutetia, and can be used to constrain its evolution. The derived ages of the main units of Lutetia show its active collisional poorly constrained parameters. Thus, it is possible -although unlikely- that the formation of Massalia crater triggered major crater reset on nearby regions. 9 history, lasting for about 4 Ga. The extremely young Bt1a unit, with an age of < 220 Ma, indi- cates that major (collisional) events occurred until very recent times. We also find evidence on the oldest Achaia region -and possibly Noricum region-, of a non-uniform radial strength profile, possibly due to the effects of previous collisions that produced a highly fractured surface on top a competent interior. In this respect, Lutetia resembles what has been found on other much larger bodies like Mercury and the Moon [14]. It is also noteworthy that the observed cratering seems to be produced by a population having a shallower cumulative slope than predicted by the [4] model. The overall slope seems to be consistent with recent observations [5], although our size range of interest extends beyong their obsevational limits. This result, if confirmed by further studies, will require a revision of the present collisional models. On the other hand, according to the present theories of main belt evolution, Lutetia should be a primordial object [4]. This is also confirmed by Lutetia's high density that makes it unlikely to be a fragment of a larger body [27]. This consideration is, however, in contradiction with the derived crater retention ages. Either the chronology scheme is not accurate, or some major event occurred in Lutetia history to reset its surface. Concerning the adopted lunar chronology, it likely underestimates the real ages of main belt asteroids. Indeed, the exponential increase of the lunar impactor flux for ages older than 3.5 Ga did not likely take place in the main belt unless the main belt suffered an intense cometary bombardment. In fact, in the current scenario of main belt evolution during the Late Heavy Bombardment (LHB) at ∼ 3.9 Ga, the main belt depleted by a factor of 2-4 at most [17]. On the other hand, the lunar chronology (Equ. 2) predicts an increase in the impactor flux of about a factor of ∼ 5 and ∼ 40 in the time spans 3.5 − 3.9 Ga and 3.5 − 4.2 Ga, respectively. This steep increase in the impactor flux results in too young crater retention ages of asteroid surfaces. However, dynamical models of the early evolution of the main belt are not yet robust enough to successfully be used for precise age determination. Concerning possible resetting event(s), the most energetic event that we can infer is the for- mation of Massalia crater, which, according to the previous discussion, was not able to reset the whole surface. Unless this conclusion is affected by poorly constrained parameters or other more energetic event(s) took place in the Lutetia southern hemisphere (not imaged by OSIRIS), this option appears untenable. Lutetia's crater age conundrum still remains unsolved. Nevertheless, we expect major im- provements in our theoretical and observational understandings of the main belt in the near fu- ture. In particular, the Dawn mission arrived at Vesta, the second largest asteroid, in July 2011. High resolution imaging of Vesta will help to constrain the early impact history of the main belt and the evolution of its primordial asteroids, Lutetia included. Acknowledgments We thank the referee D. O'Brien and an anonymous referee for helpful comments which im- proved the manuscript. OSIRIS was built by a consortium of the Max-Planck-Institut fur Sonnensystemforschung, Katlenburg- Lindau, Germany, CISAS - University of Padova, Italy, the Laboratoire d'Astrophysique de Mar- seille, France, the Instituto de Astrofsica de Andalucia, CSIC, Granada, Spain, the Research and Scientific Support Department of the European Space Agency, Noordwijk, The Netherlands, 10 the Instituto Nacional de Tecnica Aeroespacial, Madrid, Spain, the Universidad Politechnica de Madrid, Spain, the Department of Physics and Astronomy of Uppsala University, Sweden, and the Institut fur Datentechnik und Kommunikationsnetze der Technischen Universitat Braun- schweig, Germany. The support of the national funding agencies of Germany (DLR), France (CNES), Italy (ASI), Spain (MEC), Sweden (SNSB), and the ESA Technical Directorate is grate- fully acknowledged. We thank the Rosetta Science Operations Centre and the Rosetta Mission Operations Centre for the successful flyby of (21) Lutetia. 11 References [1] Belton, M. J. S., Chapman, C. R., Veverka, J., Klaasen, K. P., Harch, A., Greeley, R., Greenberg, R., Head, III, J. W., McEwen, A., Morrison, D., Thomas, P. C., Davies, M. E., Carr, M. H., Neukum, G., Fanale, F. P., Davis, D. R., Anger, C., Gierasch, P. J., Ingersoll, A. P., Pilcher, C. B., Sep. 1994. First Images of Asteroid 243 Ida. Science 265, 1543 -- 1547. [2] Belton, M. J. S., Veverka, J., Thomas, P., Helfenstein, P., Simonelli, D., Chapman, C., Davies, M. E., Greeley, R., Greenberg, R., Head, J., Sep. 1992. Galileo encounter with 951 Gaspra - First pictures of an asteroid. Science 257, 1647 -- 1652. [3] Besse, S. et al. 2010. Icarus, submitted. [4] Bottke, W. F., Durda, D. D., Nesvorn´y, D., Jedicke, R., Morbidelli, A., Vokrouhlick´y, D., Levison, H. F., Dec. 2005. Linking the collisional history of the main asteroid belt to its dynamical excitation and depletion. Icarus 179, 63 -- 94. [5] Gladman, B. J., et al. 2009, Icarus 202, 104 [6] Cremonese, G. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [7] Farinella, P., & Davis, D. R., 1992. Collision rates and impact velocities in the Main Asteroid Belt. Icarus, 97, 111. [8] Holsapple, K. A., Housen, K. R., Mar. 2007. A crater and its ejecta: An interpretation of Deep Impact. Icarus 187, 345 -- 356. [9] Keller, H. U. et al. 2010. E-Type Asteroid (2867) Steins as Imaged by OSIRIS on Board Rosetta. Science 327, 190. [10] Keller, H. U., Barbieri, C., Lamy, P., Rickman, H., Rodrigo, R., Wenzel, K., Sierks, H., A'Hearn, M. F., Angrilli, F., Angulo, M., Bailey, M. E., Barthol, P., Barucci, M. A., Bertaux, J., Bianchini, G., Boit, J., Brown, V., Burns, J. A., Buttner, I., Castro, J. M., Cremonese, G., Curdt, W., da Deppo, V., Debei, S., de Cecco, M., Dohlen, K., Fornasier, S., Fulle, M., Germerott, D., Gliem, F., Guizzo, G. P., Hviid, S. F., Ip, W., Jorda, L., Koschny, D., Kramm, J. R., Kuhrt, E., Kuppers, M., Lara, L. M., Llebaria, A., L´opez, A., L´opez-Jimenez, A., L´opez-Moreno, J., Meller, R., Michalik, H., Michelena, M. D., Muller, R., Naletto, G., Orign´e, A., Parzianello, G., Pertile, M., Quintana, C., Ragazzoni, R., Ramous, P., Reiche, K., Reina, M., Rodr´ıguez, J., Rousset, G., Sabau, L., Sanz, A., Sivan, J., Stockner, K., Tabero, J., Telljohann, U., Thomas, N., Timon, V., Tomasch, G., Wittrock, T., Zaccariotto, M., Feb. 2007. OSIRIS The Scientific Camera System Onboard Rosetta. Space Science Reviews 128, 433 -- 506. [11] Kueppers, M. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [12] Marchi, S., Mottola, S., Cremonese, G., Massironi, M., Martellato, E., Jun. 2009. A New Chronology for the Moon and Mercury. Astronomical Journal 137, 4936 -- 4948. [13] Marchi, S., et al. 2010, PSS, 58, 1116 [14] Marchi, S., Massironi, M., Cremonese, G., Martellato, E., Giacomini, L., & Prockter, L. 2011, arXiv:1105.5272 [15] Massironi, M., Cremonese, G., Marchi, S., et al. 2009, GRL, 362, 21204 [16] Massironi, M. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [17] Morbidelli, A., Brasser, R., Gomes, R., Levison, H. F., & Tsiganis, K. 2010, Astronomical Journal 140, 1391 -- 1401. [18] Neukum, G., & Ivanov, B. A. 1994, Hazards Due to Comets and Asteroids, 359. [19] O'Brien, D. P., Greenberg, R., Richardson, J. E., Jul. 2006. Craters on asteroids: Reconciling diverse impact records with a common impacting population. Icarus 183, 79 -- 92. [20] Preusker, M. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [21] Saito, J., Miyamoto, H., Nakamura, R., Ishiguro, M., Michikami, T., Nakamura, A. M., Demura, H., Sasaki, S., Hirata, N., Honda, C., Yamamoto, A., Yokota, Y., Fuse, T., Yoshida, F., Tholen, D. J., Gaskell, R. W., Hashimoto, T., Kubota, T., Higuchi, Y., Nakamura, T., Smith, P., Hiraoka, K., Honda, T., Kobayashi, S., Furuya, M., Matsumoto, N., Nemoto, E., Yukishita, A., Kitazato, K., Dermawan, B., Sogame, A., Terazono, J., Shinohara, C., Akiyama, H., Jun. 2006. Detailed Images of Asteroid 25143 Itokawa from Hayabusa. Science 312, 1341 -- 1344. [22] Sierks, H. et al, 2011. xxxxxxx. Science. [23] Veverka, J., Thomas, P., Harch, A., Clark, B., Bell, J. F., Carcich, B., Joseph, J., Murchie, S., Izenberg, N., Chapman, C., Merline, W., Malin, M., McFadden, L., Robinson, M., Jul. 1999. NEAR Encounter with Asteroid 253 Mathilde: Overview. Icarus 140, 3 -- 16. [24] Veverka, J., Thomas, P. C., Bell, III, J. F., Bell, M., Carcich, B., Clark, B., Harch, A., Joseph, J., Martin, P., Robinson, M., Murchie, S., Izenberg, N., Hawkins, E., Warren, J., Farquhar, R., Cheng, A., Dunham, D., Chapman, C., Merline, W. J., McFadden, L., Wellnitz, D., Malin, M., Owen, Jr., W. M., Miller, J. K., Williams, B. G., Yeomans, D. K., Jul. 1999. Imaging of asteroid 433 Eros during NEAR's flyby reconnaissance. Science 285, 562 -- 564. [25] Vincent, J.B. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [26] Thomas, N. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. [27] Weiss, B. et al, 2011. xxxxxxx. PSS this issue 00, 0 -- 0. 12 Figure 1: The 6 major regions identified on Lutetia. Ac: Achaia, Nr: Noricum, Nb: Narbonensis, Bt: Baetica, Et: Etruria, Pa: Pannonia. Note that some bondaries may sligtlhy vary according to different authors. For a more detailed definition of the regions see [16] and [26]. Some of the major units (i.e. subdivisions of the regions) are also reported. Colored units are those used for crater counts. The corresponding areas (km2) are: 760 (Bt1a), 2875 (Ac1+Ac2), 2042 (Nr1+Nr2), 2647 (Nb1). The blue "+" at the center of the image indicates the north pole. 13 Figure 2: The panels indicate the craters counted on the 4 units investigated. Crater counts have been performed on image NAC.15.42.41.240 for Achaia (b), Noricum (a) and Narbonensis (c) regions, and on image NAC.15.44.41.262 for Bt1a region (d). 14 Achaia Region a) Achaia Region Noricun Region b) 10-1 10-2 10-3 10-1 10-2 10-3 ) 2 - m k ( s r e t a r c f o r e b m u n e v i t a l u m u C 10-4 0.1 1 10 10-4 0.1 100 1 10 100 Crater diameter (km) Achaia Region Bt1a Unit d) Achaia Region Narbonensis Region c) 10-1 10-2 10-3 10-1 10-2 10-3 ) 2 - m k ( s r e t a r c f o r e b m u n e v i t a l u m u C 10-4 0.1 1 10 10-4 0.1 100 1 10 100 Crater diameter (km) Figure 3: Crater size-frequency distributions of the 4 units reported in Figure 2 (Achaia region (a), Noricum region (b), Narbonensis region (c), Bt1a unit (d)). Achaia crater SFD is reported in all panels for a better comparison. 15 Figure 4: Close view of Narbonensis and Baetica regions (top-left and bottom-left panels, respectively). Topographical slope (i.e., the angle between local shape and local gravity) for Narbonensis (top-right panel) and Baetica (bottom-right panel) regions. 16 Bottke et al. (2005) Gladman et al. (2009) 1011 1010 109 108 107 106 105 104 103 r e b m u n e v i t a l u m u C 102 10-2 10-1 100 Diameter (km) 101 102 Figure 5: Main belt size-frequency distributions used in this work. The model [4] distribution is based on the observed debiased main belt population down to about 1 km. The [5] distribution is obtained in the following manner: for impactor sizes ≥ 3 km (corresponding to the completeness limit of H-magnitute ∼ 15, for an assumed albedo of 0.2), it overlaps with the [4] SFD; while for sizes < 3 km it is has a cumulative slope of -1.5. Note that the SKADS survey is valid down to H ∼ 18 (correponding to about 0.8 km), nevertheless we extrapolated the -1.5 slope to smaller sizes. 17 0.5 ) v ( f , y t i l i b a b o r p t c a p m I 0.25 0 0 5 Average Main Belt Steins Lutetia Moon 10 Impact velocity (km/s) 15 20 Figure 6: Lutetia's impact velocity distribution. For a comparison, the distribution for Steins, the average main belt (shaded area) and the Moon (largely out of scale) are also reported. The average impact velocity for Lutetia is 4.3 km/s. 18 100 10-1 10-2 10-3 ) 2 - m k ( s r e t a r c f o r e b m u n e v i t a l u m u C Achaia Region 3.2+0.2 3.7+0.1 3.8+0.1 -0.5 Ga (P1, hard-rock, w/o CO) -0.1 Ga (P1, hard-rock, CO) -0.1 Ga (P2, hard-rock, CO) Achaia Region 3.6+0.1 3.7+0.1 -0.1 Ga (P1, CO) -0.1 Ga (P2, CO) 100 10-1 10-2 10-3 10-4 0.1 1 10 10-4 0.1 100 1 10 100 Crater diameter (km) Figure 7: Achaia MPF best fit. Left panel: Best fits obtained using hard rock scaling law, with and without crater obliteration (CO), and using the MBA population from [4] (P1) and [5] (P2). Right panel: Best fits obtained modeling a transition in the physical properties of Achaia region, namely adopting a fractured layer onto a more competent interior (see text for more details). 19 Bt1a Unit -2 - 51+11 7+2 51+7 -7 - 220+25 -11 Ma (P1) -25 Ma (P2) 100 10-1 10-2 10-3 10-4 ) 2 - m k ( s r e t a r c f o r e b m u n e v i t a l u m u C 10-5 0.1 1 10 100 Crater diameter (km) Figure 8: Baetica MPF best fit for two impactor populations. The crater scaling law for cohesive soil has been used. The derived age ranges correspond to two limiting values of strength, namely 105 and 107 dyne/cm2, respectively. No crater obliteration has been applied here, given the very young ages involved (see text for further details). 20 Noricum Region 3.4+0.1 3.7+0.1 -0.5 Ga (P1, CO) -0.1 Ga (P2, CO) 10-1 10-2 10-3 ) 2 - m k ( s r e t a r c f o r e b m u n e v i t a l u m u C Narbonensis Region 0.95+0.2 -0.2 Ga (P1, CO) 1.26+0.4 -0.2 Ga (P2, CO) 10-1 10-2 10-3 10-4 0.1 1 10 10-4 0.1 100 1 10 100 Crater diameter (km) Figure 9: Noricum and Narbonensis MPF best fits (see text for further details). 21
1008.2484
1
1008
2010-08-14T23:52:56
Orbital evolution under action of fast interstellar gas flow
[ "astro-ph.EP" ]
Orbital evolution of an interplanetary dust particle under action of an interstellar gas flow is investigated. Secular time derivatives of the particle orbital elements, for arbitrary orbit orientation, are presented. An important result concerns secular evolution of semi-major axis. Secular semi-major axis of the particle on a bound orbit decreases under the action of fast interstellar gas flow. Possible types of evolution of other Keplerian orbital elements are discussed. The paper compares influences of the Poynting-Robertson effect, the radial solar wind and the interstellar gas flow on dynamics of the dust particle in outer planetary region of the Solar System and beyond it, up to 100 AU. Evolution of putative dust ring in the zone of the Edgeworth-Kuiper belt is studied. Also non-radial solar wind and gravitational effect of major planets may play an important role. Low inclination orbits of micron-sized dust particles in the belt are not stable due to fast increase of eccentricity caused by the interstellar gas flow and subsequent planetary perturbations - the increase of eccentricity leads to planet crossing orbits of the particles. Gravitational and non-gravitational effects are treated in a way which fully respects physics. As a consequence, some of the published results turned out to be incorrect. Moreover, the paper treats the problem in a more general way than it has been presented up to now. The influence of the fast interstellar neutral gas flow might not be ignored in modeling of evolution of dust particles beyond planets.
astro-ph.EP
astro-ph
Orbital evolution under action of fast interstellar gas flow P. P´astor1,2, J. Klacka1, and L. K´omar1 1 Department of Astronomy, Physics of the Earth, and Meteorology, Faculty of Mathematics, Physics and Informatics, Comenius University, Mlynsk´a dolina, 842 48 Bratislava, Slovak Republic e-mail: [email protected], [email protected], [email protected] 2 Tekov Observatory, Sokolovsk´a 21, 934 01, Levice, Slovak Republic ABSTRACT Orbital evolution of an interplanetary dust particle under action of an interstellar gas flow is investigated. Secular time derivatives of the particle orbital elements, for arbitrary orbit orientation, are presented. An important result concerns secular evolution of semi-major axis. Secular semi-major axis of the particle on a bound orbit decreases under the action of fast interstellar gas flow. Possible types of evolution of other Keplerian orbital elements are discussed. The paper compares influences of the Poynting-Robertson effect, the radial solar wind and the interstellar gas flow on dynamics of the dust particle in outer planetary region of the Solar System and beyond it, up to 100 AU. Evolution of putative dust ring in the zone of the Edgeworth-Kuiper belt is studied. Also non-radial solar wind and gravitational effect of major planets may play an important role. Low inclination orbits of micron-sized dust particles in the belt are not stable due to fast increase of eccentricity caused by the interstellar gas flow and subsequent planetary perturbations -- the increase of eccentricity leads to planet crossing orbits of the particles. Gravitational and non-gravitational effects are treated in a way which fully respects physics. As a consequence, some of the published results turned out to be incorrect. Moreover, the paper treats the problem in a more general way than it has been presented up to now. The influence of the fast interstellar neutral gas flow might not be ignored in modeling of evolution of dust particles beyond planets. Key words. ISM: general, Celestial mechanics, Interplanetary medium 1. Introduction Motion of stars relative to their local interstellar medium is frequent/usual process in galaxies. Neutral atoms pene- trate into the Solar System due to the relative motion of the Sun with respect to the interstellar medium. This flow of neutral atoms through a heliosphere has been investigated in many papers, e.g. Fahr (1996), Lee et al. (2009), Mobius et al. (2009). Motion of dust in interplanetary space can be affected by the neutral gas penetrating into the heliosphere. Influence of this effect on dynamics of dust particles is usu- ally ignored in literature. The Poynting-Robertson effect, the radial solar wind and the gravitational perturbation of planet(s) are usually taken into account (Sidlichovsk´y & Nesvorn´y 1994; Liou & Zook 1997; Liou & Zook 1999; Kuchner & Holman 2003). Scherer (2000) has calculated secular time derivatives of angular momentum and Laplace-Runge-Lenz vector of a dust particle under the action of interstellar gas flow. But Scherer's calculations contain several incorrectnesses. He has come to the conclusion that semi-major axis of the dust particle increases exponentially (Scherer 2000, p. 334). This paper presents that semi-major axis of the dust particle decreases under the action of interstellar gas flow, in the framework of the perturbation theory. Motion of dust particles in the zone of the Edgeworth- Kuiper belt under the action of the interstellar flow of gas has been investigated by Klacka et al. (2009a). The authors have calculated secular time derivatives of orbital elements only for the case when interstellar gas velocity vector lies in the orbital plane of the dust particle and direction of the velocity vector is parallel with y-axis. This paper overcomes these restrictions. Moreover, it presents some main proper- ties of dust dynamics under the action of the interstellar gas. 2. Secular evolution Acceleration of a spherical dust particle caused by the flow of neutral gas can be given in the form (Scherer 2000) dv dt = − cD γH v − vH (v − vH ) , (1) where vH is velocity of the neutral hydrogen atom, v is velocity of the dust grain, cD is the drag coefficient, γH is the collision parameter. For the collision parameter we can write γH = nH mH m A , (2) where mH is mass of the neutral hydrogen atom, nH is the concentration of interstellar neutral hydrogen atoms, A = πR2 is the geometrical cross section of the spherical dust grain of radius R and mass m. The concentration of interstellar hydrogen nH is not constant in the entire helio- sphere. For heliocentric distances r less than 4 AU nH de- creases precipitously from its value in the outer heliosphere toward the Sun, due to ionization (Lee et al. 2009). But in the outer heliosphere, r ∈ (30 AU, 80 AU), we can assume that the concentration of the neutral hydrogen atoms is constant nH = 0.05 cm−3 (Fahr 1996). The same assump- tion can be used also behind the solar wind termination 2 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow shock. The shock was crossed by Voyager 1 at a heliocen- tric distance 94 AU and by Voyager 2 at 84 AU (Richardson et al. 2008). We will assume that the speed of interstellar gas is much greater than the speed of the dust grain in the stationary frame associated with the Sun (v ≪ vH ). This approxi- mation leads to approximately constant value of cD ≈ 2.6 (Baines et al. 1965; Banaszkiewicz et al. 1994; Klacka et al. 2009a). We want to find influence of the flow of interstellar gas on secular evolution of particle's orbit. We will assume that the dust particle is under the action of solar gravitation and the flow of neutral gas. Hence we have equation of motion (3) µ r3 = − r − cD γH v − vH (v − vH ) , dv dt where µ = GM⊙, G is the gravitational constant, M⊙ is mass of the Sun, r is position vector of the dust particle with respect to the Sun and r = r. The acceleration caused by the flow of interstellar gas will be considered as a per- turbation acceleration to the central acceleration caused by the solar gravity. In order to compute secular time deriva- tives of Keplerian orbital elements (a - semi-major axis, e - eccentricity, ω - argument of perihelion, Ω - longitude of ascending node, i - inclination) we want to use Gauss perturbation equations of celestial mechanics. To do this, we need to know radial, transversal and normal compo- nents of acceleration given by Eq. (1). Orthogonal radial, transversal and normal unit vectors of the dust particle on the Keplerian orbit are (see Fig. 1 and e.g. P´astor 2009) eR = (cos Ω cos(f + ω) − sin Ω sin(f + ω) cos i , sin Ω cos(f + ω) + cos Ω sin(f + ω) cos i , sin(f + ω) sin i) , (4) Fig. 1. A particle on an elliptical orbit depicted together with the radial, transversal and normal unit vectors. Angles characterizing position of the particle on the orbit are also shown. where , vH σ = pµ/p and vH = vH. Hence v − vH2 = µ p (1 + 2e cos f + e2) (12) (5) (6) [B + e(A sin f + B cos f )] + v2 where the identity √A2 + B 2 + C 2 = vH was used. p − 2r µ H , (13) If we denote components of the velocity vector of hydro- gen gas in the stationary Cartesian frame associated with the Sun as vH = (vHX , vHY , vHZ ), then we obtain A sin f + B cos f = (− cos Ω sin ω − sin Ω cos ω cos i) vHX + (− sin Ω sin ω + cos Ω cos ω cos i) vHY + cos ω sin i vHZ = I . (14) eT = (− cos Ω sin(f + ω) − sin Ω cos(f + ω) cos i , − sin Ω sin(f + ω) + cos Ω cos(f + ω) cos i , cos(f + ω) sin i) , eN = (sin Ω sin i, − cos Ω sin i, cos i) , where f is true anomaly. Thus, we need to calculate the values of aR = dv/dt · eR, aT = dv/dt · eT and aN = dv/dt· eN . Velocity of the particle in an elliptical orbit can be calculated from v = dr dt = d dt (reR) = r e sin f 1 + e cos f df dt where r = p 1 + e cos f eR + r eT df dt , (7) (8) and p = a(1 − e2). In this calculation also the second Kepler's law df /dt = √µp/r2 must be used. Now, one can easily verify that (v − vH ) · eR = vH σ e sin f − vH · eR = vH σ e sin f − A , (v − vH ) · eT = vH σ (1 + e cos f ) − vH · eT = vH σ (1 + e cos f ) − B , (v − vH ) · eN = − vH · eN = − C , Now we consider only such orbits for which σ ≪ 1 , or, more precisely, the value σ2 is negligible in comparison with σ. This is reasonable for orbits with not very large eccentricities, since v ≪ vH . Using this approximation, Eqs. (13)-(14) yield (15) σ vH (B + eI)] . v − vH = vH [1 − For radial, transversal and normal components of accelera- tion we obtain from Eq. (1), Eqs. (9)-(11) and Eq. (16) (16) aR = − cD γH v2 H" A vH (cid:18) σeI vH − 1(cid:19) (9) (10) (11) P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 3 µ = 2 a 1 − e2 r p µ (cid:20)aR sin f + aT (cid:18)cos f + = r p e r p = − √µ p r µ (cid:18)aR cos f − aT sin i sin(f + ω) sin(f + ω) cos i , − aN 1 r = √µ p aN , sin i e + cos f 1 + e cos f(cid:19)(cid:21) , sin f(cid:19) 2 + e cos f 1 + e cos f da dt de dt dω dt dΩ dt di dt AB v2 H (cid:19)# , + σ(cid:18)e sin f + H" B vH (cid:18) σeI vH − 1(cid:19) aT = − cD γH v2 H(cid:19)# , + σ(cid:18)1 + e cos f + aN = − cD γH vH C σeI vH − 1 + σ B 2 v2 (17) (18) (19) B vH! . Now we can use Gauss perturbation equations of celestial mechanics to compute time derivatives of orbital elements. The perturbation equations have the form [aR e sin f + aT (1 + e cos f )] , = r √µ p aN cos(f + ω) . (20) 0 0 1 = df g dt = hgi = dt(cid:19)−1 g(cid:18) df Time average of any quantity g during one orbital period T can be computed using √µ 2πa3/2 Z 2π T Z T √µ g(cid:18)√µp r2 (cid:19)−1 2πa3/2 Z 2π 2πa2√1 − e2 Z 2π (21) where the second (√µp = r2df /dt) and the third (4π2a3 = µT 2) Kepler's laws were used. From Eqs. (17)-(21) we fi- nally obtain for the secular time derivatives of the Keplerian orbital elements g r2 df , df = 1 0 0 (cid:28) da dt(cid:29) = − 2 a cD γH v2 H r p × "I 2 − (I 2 − S 2) µ " 3I (cid:28) de dt(cid:29) = cD γH vH r p 2 µ σ (1 + 1 − √1 − e2 e2 1 v2 H #) , + σ(I 2 − S 2)(1 − e2) vH e3 × (cid:18)1 − e2 2 −p1 − e2(cid:19)# , µ (− r p 3S e dt(cid:29) = (cid:28) dω cD γH vH 2 + σSI vH e4(cid:20)e4 − 6e2 + 4 − 4(1 − e2)3/2(cid:21) (23) + C σ vH − cos i sin i (cid:20) 3e sin ω 1 − e2 (S cos ω − I sin ω)(cid:21)) , 3e sin ω 2 sin i r p 1 − e2 (S cos ω − I sin ω)# , µ "− σ vH cD γH vH C cD γH vH C 2 µ " 3e cos ω r p 1 − e2 (S sin ω + I cos ω)# , + σ vH dt (cid:29) = (cid:28) dΩ + dt(cid:29) = − (cid:28) di (24) (25) (26) where the quantities S = (cos Ω cos ω − sin Ω sin ω cos i) vHX + (sin Ω cos ω + cos Ω sin ω cos i) vHY + sin ω sin i vHZ , I = (− cos Ω sin ω − sin Ω cos ω cos i) vHX + (− sin Ω sin ω + cos Ω cos ω cos i) vHY + cos ω sin i vHZ , C = sin Ω sin i vHX − cos Ω sin i vHY + cos i vHZ , are values of A = vH · eR, B = vH · eT and C = vH · eN at perihelion of particle's orbit (f = 0), respectively. The value of C is a constant on a given oscular orbit. The values of S, I and C are depicted in Fig. 2. One can use also the (27) Fig. 2. A schematic representation of the values S, I and C for a given orbit. relations (22) S cos ω − I sin ω = cos Ω vHX + sin Ω vHY = eP A · vH , S sin ω + I cos ω = − sin Ω cos i vHX + cos Ω cos i vHY + sin i vHZ = eP E · vH , eP A · eP E = 0 , eN × eP A = eP E . (28) Vector eP A is directed from the Sun to the ascending node. eP A · vH = S cos ω − I sin ω is magnitude of vH compo- nent parallel with the line of nodes. The orbital plane is defined by its normal unit vector eN . eN × eP A = eP E. 4 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow ~vH ~eT P ~vH ~eRP c ~vH d ~vH ~eRP S > 0 I > 0 S > 0 I < 0 h dS dt i < 0 ~eT P h dS dt i > 0 ~eRP S < 0 I < 0 h dS dt i < 0 ~eT P S < 0 I > 0 h dS dt i > 0 a b ~eT P ~eRP Fig. 3. Secular time derivatives of S (see Eqs. 27) for a dust particle on prograde orbit in planar case. Origins of these Cartesian coordinate systems are in the Sun and vertical axes are aligned with the direction of the hydrogen gas velocity vector. eRP and eT P are unit radial and transversal vectors in perihelion of the particle orbit (see text). eP E · vH = S sin ω + I cos ω is magnitude of vH compo- nent perpendicular to the line of nodes and lying in the orbital plane. 3. Theoretical discussion Eqs. (22)-(26) enable to deduce some properties of secular evolution of the dust particle under the action of the flow of interstellar gas. C = 0 for a special case when the velocity of hydrogen gas vH lies in the orbital plane of the particle. In this case we get that the inclination and the longitude of ascending node are constant. Secular time derivatives for a special case of this kind (i = Ω = 0, vHX = vHZ = 0, vHY = vH in Eq. 27) have been derived in Klacka et al. (2009a). Secular time derivatives of a, e and ω are given in Klacka et al. (2009a) without generalization represented by Eqs. (27) and Fig. 2. Putting σ = 0 in Eqs. (22)-(26) one obtains a solution equivalent to the solution of Eq. (1) with the RHS inde- pendent of the particle's velocity (constant force). Eq. (22) yields constant semi-major axis for the unrealistic case σ = 0. We will show, using Eq. (22), that secular semi-major axis decreases under the action of interstellar gas flow for σ > 0. If secular increase of the semi-major axis would occur, then the value of curl braces in Eq. (22) would be negative. The value of 1 − √1 − e2 is always positive or zero. Thus, the curly braces could be negative only for I 2 − S 2 > 0. Since I and S are radial and transversal components of the constant vector vH in the perihelion (Fig. 2) of particle's orbit, we obtain maximal value of I 2 − S 2 for orbit orienta- tion characterized by S = 0. Using these assumptions, the minimal value (M V ) in the curl braces is M V = 1 + = 1 + = 1 + e2 1 v2 H I 2 − I 2 1 − √1 − e2 e2 − 1 + √1 − e2 Hp1 − e2 1 − √1 − e2 I 2 v2 H I 2 v2 e2 e2 ! > 0 . (29) The positiveness of M V means that the secular semi-major axis is a decreasing function of time. Eq. (22) was de- rived under the assumption that Solar gravity represents dominant acceleration in comparison to the interstellar gas disturbing acceleration. Since gravitational acceleration de- creases with square of a heliocentric distance, the assump- tion about the disturbing interstellar gas acceleration may not be valid far from the Sun. Behind the heliocentric dis- tance at which solar gravitational acceleration equals the in- terstellar gas acceleration, the particle's secular semi-major axis can also be an increasing function of time. The two ac- celerations are equal at heliocentric distance ≈ 1 × 104 AU for dust particle with R = 1 µm and = 2 g/cm3, and, at heliocentric distance ≈ 7 × 103 AU for dust particle with R = 1 µm and = 1 g/cm3. If Eq. (22) can be used, then secular time derivative of the semi-major axis is propor- tional to the value of semi-major axis (value of pp/µ σ is independent of semi-major axis). Our result differs from the Scherer's statement (semi-major of the particle can increase exponentially) at least due to an error in his expression for secular time derivative of magnitude of angular momentum. Defining the function w(e) ≡ 1 − e2 / 2 − √1 − e2, e ∈ [0,1), present in Eq. (23), we obtain dw de ≡ e2 d de(cid:18)1 − = − e + 2 −p1 − e2(cid:19) √1 − e2 ≥ 0 . e (30) Thus, w(e) is an increasing function of eccentricity. We ob- tain that w(e) ≥ 0 for all e ∈ [0,1), since w(e) is an in- creasing function of eccentricity and w(0) = 0. Hence, the sign of the second term in the square braces in Eq. (23) will depend on the sign of I 2 − S 2. For small values of the eccentricity we get that the secu- lar time derivative of eccentricity is, approximately, propor- tional only to the first term multiplied by I in the square braces in Eq. (23). Eq. (24) yields that the argument of perihelion is con- stant for the planar case C ≡ 0 and for the orbit orientation characterized by S = 0. The longitude of the ascending node and the inclination are constant in this case. Hence differ- entiation of the second equation of Eqs. (27) with respect to time gives dI dt = − S dω dt . (31) Thus, I is also constant in this case. Since σ is a small number, the first term in square braces in Eq. (23) is the dominant term on evolution eccentricity. P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 5 into averaged time derivatives of the quantities S, I and C defined by Eqs. (27), one obtains Now, let us consider the planar case (C ≡ 0) with S 6= 0. The differentiation of the first of Eqs. (27) with respect to time gives dS dt = I dω dt . This quantity can be averaged using Eq. (21). We get (cid:28) dS dt(cid:29) = I(cid:28) dω dt(cid:29) . (cid:28) dS dt(cid:29) = (32) (33) (cid:28) dI dt(cid:29) = If σ is a small number and I is not close to zero, then we can use the following approximations of Eqs. (23)-(24) + − cD γH vH S 2 σ I 2 3I e − µ (− r p vH(cid:20)C 2 e4(cid:18)e4 − 6e2 + 4 − 4(1 − e2)3/2(cid:19)(cid:21)) , 3S 2 e − e4(cid:18)e4 − 6e2 + 4 − 4(1 − e2)3/2(cid:19)(cid:21)) , µ (− r p 1 − e2 + 3eC 2 S 2 2 cD γH vH σI vH(cid:20)C 2 3 cD γH vH 2 dt(cid:29) ≈ (cid:28) de dt(cid:29) ≈ − (cid:28) dω I , µ r p r p µ 3 cD γH vH 2 S e . Inserting Eqs. (34)-(35) into Eq. (33) one can obtain (cid:28) de dt(cid:29) ≈ − e S (cid:28) dS dt(cid:29) . (34) (35) (36) (cid:28) dC dt (cid:29) = cD γH vH C 2 µ " 3eI r p 1 − e2 (S 2 + I 2)# . + σ vH Eq. (39) yields for the planar case (C ≡ 0) dt(cid:29) = (cid:28) dS µ "− r p cD γH vH S 3I e 2 This equation leads to the differential equation + (39) (40) (41) (42) de e ≈ − dS S , with the solution e ≈ . D S D is a constant which can be determined from initial con- ditions. Thus, eccentricity is close to its minimal value if the major axis of the orbit is aligned with the direction of the hydrogen gas velocity vector. If we consider the system Eqs. (23)-(26) with σ = 0, then we obtain hdS/dti = − 3 cD γH vH pp/µ I S / (2e) from the first of Eqs. (27). If we use Eq. (23) with σ = 0, then we get equation hde/dti = − e hdS/dti / S. This equation leads to the solution e = D / S also for this non- planar case. But approximation σ ≈ 0 is not allowed in Eqs. (24)-(26) since the first terms in square braces multiplied by C are multiplied by e, sin ω and cos ω which can be close to zero. The case σ = 0 (the RHS of Eq. 1 is independent of the particle's velocity -- constant force) is also significant for a motion of a rocket in a central gravitational field with a constant-reaction acceleration vector. We obtain the fol- lowing important result: if the constant acceleration can be considered as a perturbation acceleration to the central gravitation and S 6= 0, then equation e = D / S holds during orbital motion of the rocket. This result fills up the results of Betelsky (1964) and Kunitsyn (1966). Parameters S, I and C determine position of the orbit with respect to the hydrogen gas velocity vector. Therefore their time derivatives are useful for description of evolu- tion of the orbit position in space. Putting Eqs. (24)-(26) (37) = (38) σI 2 vH e4(cid:18)e4 − 6e2 + 4 − 4(1 − e2)3/2(cid:19)# k(e)(cid:19) . µ (cid:18) − r p σI 2 vH 3I e + 2 cD γH vH S Function k(e) is a decreasing function of eccentricity for e ∈ (0,1). This can be shown in a similar procedure as we used in Eq. (30) for the function w(e). The procedure must be used more than once for the function k(e). The function k(e) obtains values from lime→0 k(e) = − 0.5 to lime→1 k(e) = − 1, for e ∈ (0,1). For the eccentricity e = 1 we obtain lime→1 σ = ∞. Hence we use maximal value of eccentricity em for which Eq. (15) holds. The first term in parenthesis in Eq. (42) is minimal for e = em for a given value of I. Value of the parenthesis in Eq. (42) will be maximally affected by the second term for e = em. The second term in parenthesis yields for e = em , em σI 2 (43) k(em) ≥ − vH ≥ − σI ≥ − I ≥ − 3 I σI 2 vH since σ ≪ 1 and I ≤ vH . Thus, the sign of hdS/dti de- pends only on the sign of the first term in the parenthesis in Eq. (42). Fig. 3 depicts unit vectors eRP = eR(f = 0) and eT P = eT (f = 0) in the orbital plane of a particle with prograde orbit. The hydrogen gas velocity vector vH for the planar case lies in the orbital plane. Direction of vH is also shown. The unit vector eRP has direction and orientation from the Sun to perihelion of the particle orbit. In Fig. 3a eRP lies in the first quadrant of the Cartesian coordinate system with origin in the Sun and vertical axis aligned with the direction of the hydrogen gas velocity vector. Both S and I are greater than 0, for these positions of the unit vec- tors eRP and eT P . Hence hdS/dti is negative. If eRP lies in the second quadrant (Fig. 3b), then hdS/dti is positive. If eRP lies in the third quadrant (Fig. 3c), then hdS/dti is negative. Finally, if eRP lies in the fourth quadrant (Fig. 6 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow ] U A [ a 60 50 40 30 300 240 ] ° [ 180 120 60 0,0 0,5 1,0 1,5 2,0 2,5 3,0 t [ 106 years ] 1,0 0,8 0,6 0,4 0,2 0,0 90 80 70 60 ] - [ e ] ° [ 0,0 0,5 1,0 1,5 2,0 2,5 3,0 t [ 106 years ] 0,0 0,5 1,0 1,5 2,0 2,5 3,0 0,0 0,5 1,0 1,5 2,0 2,5 3,0 t [ 106 years ] t [ 106 years ] 20,0 19,5 ] ° [ i 19,0 18,5 18,0 0,0 0,5 1,0 1,5 2,0 2,5 3,0 t [ 106 years ] Fig. 4. Two evolutions of orbital elements of a dust particle with R = 2 µm and mass density = 1 g/cm3 under the action of interstellar gas flow. Evolution depicted by a solid line is calculated from the equation of motion. Evolution depicted by a dashed line corresponds to Eqs. (22)-(26). 3d), then hdS/dti is positive, as it is in the second quadrant. Therefore, the vector eRP rotates counterclockwise in the first and the second quadrants, and, clockwise in the third and the fourth quadrants. If the vector eRP is parallel with vertical axis in Fig. 3 (I = 0 and C = 0), then Eq. (40) yields hdI/dti > 0. Thus, positions of the vector eRP par- allel with the vertical axis in Fig. 3 are not stable. However, if eRP is parallel with the horizontal axis (S = 0 and C = 0), then hdS/dti = 0 and hdI/dti = 0. Thus, eRP parallel with the horizontal axis yields stable positions of the vector eRP . The stable position of the vector eRP parallel with the horizontal axis and directed to the left in Fig. 3 is of the- oretical importance, only. In reality, no particles should be observed with perihelia in this direction. However, all unit vectors eRP of particles in prograde orbits will approach the right direction in Fig. 3. For retrograde orbits we have to use transformation S → S, I → − I. We obtain hdS/dti → − hdS/dti. Hence, all unit vectors eRP of particles in retrograde orbits will approach the left direction in Fig. 3. This result was obtained, in a different way, also by Scherer (2000). Scherer states that the approaches of unit vectors eRP to one direction holds also for the non-planar case. However, the statement of Scherer is incorrect, in general. If I = 0 and C 6= 0, then Eq. (39) implies that hdS/dti is P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 7 proportional to SC 2 and not to SI. This leads to a behav- ior which differs from the behavior discussed above using Fig. 3. Scherer (2000, p. 332) furthermore states that orbital plane under the effect of the interstellar gas flow will be rotated into a plane coplanar to the flow vector vH , inde- pendent of initial position of the orbital plane. We showed, using numerical integrations, that the statement of Scherer is not correct. Retrograde orbits were not discussed by Scherer. We found an interesting orbit behavior. It depends on the orbit orientation with respect to the hydrogen gas velocity vector vH. If vH lies in a plane i = 0 and vH is perpendicular to the vector eRP (in this case S = 0 and hdS/dti = 0), then the interstellar gas flow can change pro- grade orbit into a retrograde one (even more times for one particle) and inclination does not approach the value i = 0. 4. Numerical results 4.1. Comparison of the numerical solution of equation of motion and the solution of Eqs. (22)-(26) We numerically solved Eq. (3) and the system of differen- tial equations represented by Eqs. (22)-(26). Solutions are compared in Fig. 4. We assumed that the direction of the interstellar gas velocity vector is identical to the direction of the velocity of the interstellar dust particles entering the Solar System. The interstellar dust particles enter the Solar System with a speed of about v∞ = 26 km/s (Landgraf et al. 1999) and they are arriving from direction of λecl = 259◦ (heliocentric ecliptic longitude) and βecl = 8◦ (heliocentric ecliptic latitude) (Landgraf 2000). Thus, components of the velocity in the ecliptic coordinates with x-axis aligned to- ward the equinox are vH = − 26 km/s (cos(259◦) cos(8◦), sin(259◦) cos(8◦), sin(8◦)). As an initial conditions for a dust particle with R = 2 µm and mass density = 1 g/cm3 we used ain = 60 AU, ein = 0.5, ωin = 90◦, Ωin = 90◦, iin = 20◦ for Eqs. (22)-(26). The initial true anomaly of the dust particle was fin = 180◦ for Eq. (3). Fig 4. shows that the obtained evolutions are in a good agreement. Evolutions be- gin separate as the eccentricity approaches 1. This is caused by the fact that approximation σ ≪ 1, see Eq. (15), does not hold for large eccentricities. Detailed numerical solution of the equation of motion (Eq. 3) yields that the secular semi-major axis is a decreasing function of time also when eccentricity approaches 1. We can summarize, on the basis of the previous para- graph. If there is no other force, besides solar gravity and the flux of interstellar gas, then the semi-major axis of an interplanetary dust particle decreases and the particle can hit the Sun. However, the particle can hit the Sun also by another possibility: particle's eccentricity increases to 1. These mathematical possibilities probably do not occur in reality, since other forces can act on the dust particle and the interstellar gas is ionized below the heliocentric distance of about 4 AU. Let us return, once again, to the planar case (C ≡ 0) in which S = 0 and the dominant term in the square braces in Eq. (23), the term (3/2) I, is negative. Numerical inte- gration of Eq. (3) shows that if the eccentricity decreases to 0, then the argument of perihelion ω "shifts" its value to the value ω + π. This means that the negative value of I ] U A [ a 60 50 40 30 20 0.7 0.6 ] - [ e 0.5 0.4 90 60 ] ° [ 30 0 2 4 6 8 10 t [ 105 years ] 0 2 4 6 8 10 t [ 105 years ] PR + SW + IG PR + SW PR + IG PR PR + SW + IG PR + SW PR + IG PR PR + SW + IG PR + SW PR + IG PR 0 2 4 6 8 10 t [ 105 years ] Fig. 5. Orbital evolution of dust particle with R = 2 µm, mass density = 1 g/cm3 and ¯Q′ pr = 0.75 under the action of the Poynting-Robertson effect (PR), the radial solar wind (SW) and the flow of interstellar gas (IG). changes to positive and the eccentricity begins to increase with the same slope. The approximative solution represented by Eq. (38) is in a good agreement with the detailed numerical solution of Eqs. (22)-(26) for the planar case with S 6= 0. This holds for the whole time interval, also for I close to zero. Eq. (38) holds, approximately, also for the non-planar evolution depicted in Fig. 4. In this case i is close to zero, vH ≈ vHY and Ω ≈ 90◦ at the eccentricity minimum. Eq. (38) gives e ≈ D / vH cos ω. The evolutionary minimum of eccentricity occurs for the case when ω is close to 180◦. This is in accordance with Eq. (38). 8 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 4.2. Comparing of influences of interstellar gas flow, Poynting-Robertson effect and radial solar wind on dynamics of dust particles We have considered only the effect of interstellar gas flow, up to now. In reality, some other non-gravitational effects play non-negligible role. Thus, we want to compare the ef- fect of the interstellar gas flow with the other effects influ- encing dynamics of dust grains in the Solar System. For this purpose we included the Poynting-Robertson effect (P-R ef- fect) and the radial solar wind into the equation of motion. The P-R effect is electromagnetic radiation pressure force acting on a spherical particle (Klacka 2004; arXiv:astro- ph/0807.2915; arXiv:astro-ph/0904.0368). Equation of mo- tion for the P-R effect, the effect of the radial solar wind and the effect of gas flow has the form (e.g. Klacka et al. 2009b) for the P-R effect and the radial solar wind (γH = 0 in Eq. 44). Finally, the evolution depicted by the solid line holds for the case when all three effects act together. The evolution of semi-major axis depicted in Fig. 5 shows that the flow of interstellar gas is more important than the radial solar wind, as for the effects on the dynam- ics of the dust particle. The secular eccentricity is always a decreasing function of time for the P-R effect and the radial solar wind (e.g., Wyatt & Whipple 1950, Klacka et al. 2009b). The growth in eccentricity depicted in Fig. 5 is due to the interstellar gas. Fast decrease of the semi-major axis in Fig. 5 may also be, at least partially, caused by the fact that higher eccentricities decrease the value of hda/dtiP R and the P-R effect becomes stronger. The secular evolution of eccentric- ity can be also an increasing function of time if the flow of interstellar gas is taken into account. We have dv dt = − eR µ (1 − β) r2 1 + r2 µ η ¯Q′ pr!(cid:16) v · eR − β − cD γH v − vH (v − vH ) , c eR + v c(cid:17) (44) where decrease of particle's mass (corpuscular sputtering) and higher orders in v/c are neglected. c is the speed of light in vacuum. Parameter β is defined as the ratio of the radial component of the electromagnectic radiation pressure force and the gravitational force between the Sun and the particle in rest with respect to the Sun: β = 3L⊙ ¯Q′ pr 16πcGM⊙R . = 5.763 × 10−4 ¯Q′ pr R [m] [kg/m3] . (45) L⊙ is the solar luminosity, L⊙ = 3.842 × 1026 W (Bahcall 2002), ¯Q′ pr is the dimensionless efficiency factor for radia- tion pressure integrated over the solar spectrum and cal- culated for the radial direction ( ¯Q′ pr = 1 for a perfectly absorbing sphere) and is mass density of the particle. η is the ratio of solar wind energy to electromagnetic solar energy, both radiated per unit of time η = 4πr2u L⊙ N Xi=1 nimic2 , (46) where u is the speed of the solar wind, u = 450 km/s, mi and ni, i = 1 to N , are masses and concentrations of the solar wind particles at a distance r from the Sun, η = 0.38 for the Sun (Klacka et al. 2009b). Four numerical integrations of Eq. (44) are shown in Fig. 5. We used dust particle with R = 2 µm, mass density = 1 g/cm3 and ¯Q′ pr = 0.75. We took into account only planar case when the interstellar gas velocity lies in the orbital plane of the dust particle (C = 0), for the sake of simplicity. Initial position is rin = (0, −90 AU, 0) and initial velocity vector is vin = (2 km/s, 0, 0). Orbital evolution is given by evolution of orbital elements calculated for the case when a central acceleration is defined by the first Keplerian term in Eq. (44), namely −µ(1 − β)eR/r2. The evolution depicted by the dash-dotted line is for the P-R effect alone (γH = 0, η = 0 in Eq. 44). The evolution depicted by the dotted line is for the P-R effect with the flow of interstellar gas (η = 0 in Eq. 44). The evolution depicted by the dashed line is (cid:28) deβ dt (cid:29) = − eβ a2 β(1 − e2 β)1/2 c 5 2 η ¯Q′ β 1 + pr! µ + cD γH vH r pβ µ(1 − β) × " 3Iβ β)(1 − e2 β − S 2 β) vH e3 β β!# , × 1 − 2 −q1 − e2 σ(I 2 e2 β + 2 (47) if also Eq. (23) is used. The subscript β denotes that the central acceleration −µ(1 − β)eR/r2 is used for calcula- tion of the osculating orbital elements. We remind that the transformation µ → µ(1 − β) has to be done on the RHS sides of Eqs. (20)-(26). If we use definition of the oscu- lating orbital elements, then the physical central accelera- tion is given by the gravitational acceleration of the Sun, −µeR/r2. In this case, the secular evolution of eccentricity is given by Eq. (103) in Klacka (2004), assuming that eβ is calculated from Eq. (47). If optical properties of the dust particle are constant, then the secular time derivative of the argument of perihe- lion equals to zero for the P-R effect and the radial solar wind (Klacka et al. 2007, 2009b). If the flow of interstellar gas is included into the equation of motion, then, even in the planar case, the secular time derivative of the argument of perihelion may not be equal to zero, in general (see Fig. 5). The evolutions of eccentricity and argument of perihe- lion shown in Fig. 5 are significantly affected by the flow of interstellar gas. 4.3. Dust ring in the Edgeworth-Kuiper belt zone Real situation in the Edgeworth-Kuiper belt zone may be much more complicated than the situation discussed in Secs. 4.1 and 4.2. In particular, gravitation of planets may have an important influence on dynamics of dust in the zone. For this reason we included gravitation of four ma- jor planets into the final equation of motion. Observations from Helios 2 during its first solar mission in 1976 (Bruno et. al. 2003) show that the angle between the radial direc- tion and the direction of the solar wind velocity does not significantly depend on heliocentric distance. If the value of P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 9 t = 25 x 104 years 100 50 t = 0 years 100 50 z [ A U ] 0 -50 -100 -100 -50 z [ A U ] 100 50 y [ AU ] 0 -50 50 100 -100 0 -50 -100 -100 -50 x [ A 0 U ] x [ A 0 U ] t = 50 x 104 years t = 75 x 104 years 100 50 z [ A U ] 0 -50 -100 -100 -50 x [ A 0 U ] 100 50 z [ A U ] 100 50 y [ AU ] 0 -50 50 100 -100 0 -50 -100 -100 -50 x [ A 0 U ] 100 50 y [ AU ] 0 -50 50 100 -100 100 50 y [ AU ] 0 -50 50 100 -100 Fig. 6. Time evolution of ring of dust particles with R = 2 µm, = 1 g/cm3 and ¯Q′ pr = 0.75 in the zone of the Edgeworth- Kuiper belt. The ring becomes eccentric in less than 106 years due to the interstellar neutral gas. Orbits of the particles are shown with black color and orbits of the planets are shown in gray. this angle is approximately constant, then the non-radial solar wind can also have an important influence on dynam- ics of dust in outer parts of the Solar System. We took into account the non-radial solar wind with constant value of the angle. Influence of precession of the rotational axis of the Sun on the non-radial solar wind was also considered. Hence, equation of motion of the dust particle has the form dv dt = − µ r2 eR µ v + β β η pr c u ¯Q′ r2 h(cid:16)1 − v · eR ci c (cid:17) eR − µ − r2 v − u (v − u) − cD γH v − vH (v − vH ) Xi=1 − r − ri3 (r − ri) GMi 4 − 4 Xi=1 GMi ri3 ri , (48) where u is solar wind velocity vector, Mi are masses of the planets and ri are position vectors of the planets with re- spect to the Sun. The non-radial solar wind velocity vector was calculated from equation u = u(cid:18)eR cos ε + k × eR k × eR sin ε(cid:19) , (49) where ε is the angle between the radial direction and the direction of the solar wind velocity and k is unit vec- tor with direction/orientation corresponding to the direc- tion/orientation of solar rotation angular velocity vector. Vector k for a given time can be calculated from k = (sin Ωs sin is,− cos Ωs sin is, cos is) , is = 7◦15′, Ωs = 73◦40′ + 50.25′′ (t[years] − 1850) . (50) 10 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 90 80 70 60 50 40 30 ] U A [ a ] ° [ 540 360 180 0 -180 -360 0.8 0.6 0.4 0.2 0.0 ] - [ e 0.0 0.2 0.4 0.6 t [ 106 years ] 0.8 1.0 0.0 0.2 0.8 1.0 0.4 0.6 t [ 106 years ] ] ° [ 180 90 0 -90 -180 -270 0.0 0.2 0.6 0.4 t [ 106 years ] 25 0.8 1.0 0.0 0.2 0.4 0.6 t [ 106 years ] 0.8 1.0 ] ° [ i 20 15 10 5 0 -5 0.0 0.2 0.4 0.6 t [ 106 years ] 0.8 1.0 Fig. 7. Evolution of orbital elements of dust particles in the Edgeworth-Kuiper belt zone during 96 numerical solutions depicted in Fig. 6. Initial values of orbital elements are: ain = 60 AU, ein = 0.1, ωin ∈ {0◦, 45◦, 90◦, ..., 270◦, 315◦}, Ωin ∈ {0◦, 90◦, 180◦, 270◦}, iin ∈ {5◦, 10◦, 15◦} and fin = 0◦. Evolution during the first 750 000 years is influenced mainly by interstellar gas, and, later on, mainly by gravitation of planets (see text). While Eqs. (49)-(50) are consistent with Klacka (1994) and Abalakin (1986), the value of ε, ε = 2.9◦, used in our nu- merical calculations, is in accordance with Bruno et al. (2003) and Klacka et al. (2007). The observed neutral hy- drogen gas velocity vector in the ecliptic coordinates with x-axis aligned toward the equinox is vH = − 26 km/s (cos(259◦) cos(8◦), sin(259◦) cos(8◦), sin(8◦)). As for the initial conditions of dust particles we did not use random positions and velocities. We assumed that putative dust ring in the Edgeworth-Kuiper belt has approximate circu- lar shape and contains lot of particles with approximately equal optical properties. We assumed that the ring contains such large amount of particles that one can choose, approx- imately, a given value of semi-major axis and a given radius of the particles. Accelerations caused by the P-R effect, the solar wind and the interstellar neutral hydrogen gas are in- versely proportional to particle's radius and mass density. Therefore, a large particle is less influenced by the non- gravitational effects than a small particle of the same mass density. The evolution of the large particle under action of the non-gravitational effects is slower than the evolution of the small dust grain. We used uniformly distributed initial P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow 11 values of the argument of perihelion and longitude of the ascending node. Furthermore we assumed that particles in the ring orbit prograde in low inclination orbits. We used particles with R = 2 µm, = 1 g/cm3 and ¯Q′ pr = 0.75. Exact initial values of orbital elements were ain = 60 AU, ein = 0.1, ωin ∈ {0◦, 45◦, 90◦, ..., 270◦, 315◦}, Ωin ∈ {0◦, 90◦, 180◦, 270◦}, iin ∈ {5◦, 10◦, 15◦} and fin = 0◦. Hence, we obtained 8 × 4 × 3 = 96 individual orbits. Results of numerical solutions of Eq. (48) are depicted in Figs. 6 and 7. Fig. 6 depicts evolution of the dust ring viewed from per- spective. Orbits of the planets are also shown. Time span between various pictures in Fig. 6 is 250 000 years. As we can see, the ring becomes more and more eccentric because of a fast increase of eccentricity caused by the interstellar gas flow (see also eccentricity evolution in Fig. 7). Perihelia of orbits are shifted in accordance with the behavior dis- cussed in Fig. 3. This is caused by the facts that influence of interstellar gas is dominant and the solved problem is almost coplanar. The term multiplied by C 2 in Eq. (39) does not have large influence on the first term in the curly braces in Eq. (39), in almost coplanar case. Therefore, the lines connecting the Sun with the perihelia of particles or- bits are approaching the direction perpendicular to the in- terstellar gas velocity vector. Time evolution of the orbital elements of the dust particles in the ring, considered in Fig. 6, is depicted in Fig. 7. Due to the approach of the peri- helia to one direction, the evolutions of the argument of perihelion ω, beginning with a given initial value ωin, are divided into four branches. Each of them corresponds to one initial value of the ascending node. If the time is less than 750 000 years, then: i) the concentration of the parti- cles in the ring is smallest in the direction (from the Sun) into which perihelia of the orbits are approaching, and, ii) the concentration of the particles is greatest in exactly op- posite direction. If the time is greater than 750 000 years, then the orbits of particles in the dust ring are getting close to the orbits of the planets due to the increase of particles eccentricities. The situation after 750 000 years can be seen in Fig. 7. The dust particles with R = 2 µm, = 1 g/cm3 and ¯Q′ pr = 0.75 are characterized by the value β ≈ 0.216 (see Eq. 45). For this value of β, one obtains a ≈ 57.7 AU for the location of the exterior mean motion 3/1 resonance with Neptune; it follows from a = aP (1 − β)1/3 (3/1)2/3, where aP is semi-major axis of Neptune. We can see, from the evolution of semi-major axis in Fig. 7, that the secular semi-major axis is a decreasing function of time during the first 750 000 years. Thus the semi-major axis can evolve from an initial value of 60 AU to the location close to the mean-motion 3/1 resonance. Particles are influenced both by the vicinity of Neptune orbit and the exterior mean mo- tion 3/1 resonance with Neptune. Evolution during the first 750 000 years is influenced mainly by the neutral interstel- lar hydrogen gas and, later on, mainly by the gravitation of planets. Inclusion of the P-R effect, the non-radial solar wind and the interstellar gas into the equation of motion of the dust particle without planets can stabilize the particle's or- bit. The stabilization is characterized by stable values of orbital elements. This stabilization is discussed in Klacka et al. (2009a) for sin ε = 0.05. The process of stabilization requires about 1 × 108 years for the dust particle with β = 0.01. This time is not very sensitive to the efficiency fac- tor for radiation pressure ¯Q′ pr. However, for lower values of ¯Q′ pr the stabilization occur with larger probability, because stabilization effect of the non-radial solar wind is stronger (see Eq. 48). If also planets are considered in the equation of motion, then the stabilizing value of dust particle eccen- tricity is usually sufficiently high to get the particle close to one of the planets during a long time span. As a con- sequence, gravitation of the planet can change orbit of the particle and cancel the stabilization process. 5. Conclusion We investigated orbital evolution of a dust grain under the action of interstellar gas flow. We presented secular time derivatives of the grain's orbital elements for arbitrary ori- entation of the orbit with respect to the velocity vector of the interstellar gas, which is a generalization of several re- sults presented in Klacka et al. (2009a). The secular time derivatives were derived using assumptions that the accel- eration caused by the interstellar gas flow is small in com- parison with gravitation of a central object (the Sun), ec- centricity of the orbit is not close to 1 and the speed of the dust particle is small in comparison with the speed of the interstellar gas. These assumptions lead to secular decrease of semi-major axis a of the particle. The secular time derivative of the semi-major axis is negative and pro- portional to a. This result is not in accordance with Scherer (2000) who has stated that the semi-major of the particle increases exponentially. Scherer's statement is incorrect and our analytical result is confirmed by our detailed numerical integration of equation of motion (see also Fig. 4). If the hydrogen gas velocity vector vH lies in the parti- cle's orbital plane and the major axis of the orbit is not per- pendicular to vH , then the product of (secular) eccentricity and magnitude of the radial component of vH measured in perihelion is, approximately, constant during orbital evolu- tion. We considered simultaneous action of the P-R effect, the radial solar wind and the interstellar gas flow. Numerical integrations showed that the action of the flow of inter- stellar gas can be more important than the action of the electromagnetic and corpuscular radiation of the Sun, as for the motion of dust particles orbiting the Sun in outer parts of the Solar System (see Fig. 5). Physical decrease of semi-major axis can be more than 2-times greater than the value produced by the P-R effect and radial solar wind. The evolution of eccentricity can also be an increasing function of time when we consider the P-R effect and the radial solar wind together with the flow of the neutral interstellar gas. This is also relevant difference from the action of the P-R effect and the radial solar wind when secular decrease of eccentricity occurs. Simultaneous action of all three effects yields that the secular time derivative of the argument of perihelion may not be equal to zero, in general. Gravitation of four major planets was also directly added into the equation of motion, see Eq. (48). This access correctly describes capture of dust grains into mean motion resonances with the planets. Our physical approach differs from the Scherer's approach (Scherer 2000), who has used some kind of secular access to gravitational influence of the planets. Assumption on an existence of dust ring in the zone of the Edgeworth-Kuiper belt is in contradiction with rapid increase of eccentricity of the ring due to an acceleration caused by the interstellar gas flow. Speed of the eccentricity 12 P. P´astor et al.: Orbital evolution under action of fast interstellar gas flow increase (time derivative of eccentricity) is roughly inversely proportional to the particle's size and mass density. As the eccentricity of the particles increases, the particles approach the planets. The particles in the ring are under the gravi- tational influence of the planets. The particles evolve also in semi-major axis and they can be temporarily captured into a mean motion resonance. The particles can remain in chaotic orbits between orbits of the planets, or, the particles are ejected to high eccentric orbits due to close encounters with one of the planets. Only particles with greater size and mass density should survive in the dust ring for a long time. A relevant result of the paper is that equation of motion in the form of Eq. (44) (or, Eq. 48) and Eqs. (45)-(46) have to be used in modeling of orbital evolution of dust grains in the Solar System. The influence of the fast interstellar neutral gas flow might not be ignored in general investi- gations on evolution of dust particles in the zone of the Edgeworth-Kuiper belt. Acknowledgements. The paper was supported by the Scientific Grant Agency VEGA grant No. 2/0016/09. References Abalakin, V. K., 1986, Astronomicheskij jezhegodnik SSSR, (Nauka, Leningrad), 662 Bahcall, J. 2002, Phys. Rev. C., 65, 025801 Baines, M. J., Williams, I. P., & Asebiomo, A. S. 1965, MNRAS, 130, 63 Banaszkiewicz, M., Fahr, H. J., & Scherer, K. 1994, Icarus, 107, 358 Betelsky, V. 1964, Cosmic Research, Engl. Transl., 2 (3), 346 Bruno, R., Carbone, V., Sorriso-Valvo, L., & Bavassano, B. 2003, J. Geophys. Res., 108 (A3), 1130 Fahr, H. J. 1996, Space Sci. Rev., 78, 199 Klacka, J., 1994, Earth, Moon, and Planets, 64, 125 Klacka, J., 2004, Celest. Mech. and Dynam. Astron., 89, 1 Klacka, J., Kocifaj, M., P´astor, P., & Petrzala, J. 2007, A&A, 464, 127 Klacka, J., K´omar, L., P´astor, P., & Petrzala, J. 2009a, Solar wind and motion of interplanetary dust grains, In Handbook on Solar Wind: Effects, Dynamics and Interactions, ed. H. E. Johannson (NOVA Science Publishers, Inc., New York), 227 Klacka, J., Petrzala, J., P´astor, P., & K´omar, L. 2009b, arXiv: astro- ph/0904.2673. Kuchner, M. J., & Holman, M. J. 2003, Astrophys. J., 588, 1110 Kunitsyn, A., 1966, Cosmic Research, Engl. Transl., 4 (2), 295 Landgraf, M. 2000, J. Geophys. Res., 105, 10303 Landgraf, M., Augustsson, K., Grun, E., & Gustafson, B. S. 1999, Science, 286, 2319 Lee, M. A., Fahr, H. J., Kucharek, H., et al. 2009, Space Sci. Rev., 146, 275 Liou, J.-Ch., & Zook, H. A. 1997, Icarus, 128, 354 Liou, J.-Ch., & Zook, H. A. 1999, Astron. J., 118, 580 Mobius, E., Bochsler, P., Bzowski, M., et al. 2009, Science, 326, 969 P´astor, P. 2009, arXiv: astro-ph/0907.4005 Richardson, J. D., Kasper, J. C., Wang, C., Belcher, J. W., & Lazarus, A. J. 2008, Nature, 454, 63 Scherer, K. 2000, J. Geophys. Res., 105, A5, 10329 Sidlichovsk´y, M., & Nesvorn´y, D. 1994, A&A, 289, 972 Wyatt, S. P., & Whipple, F. L. 1950, Astrophys. J., 111, 134
1108.4635
1
1108
2011-08-23T15:45:07
Periodic orbits in the gravity field of a fixed homogeneous cube
[ "astro-ph.EP", "math-ph", "math-ph", "physics.space-ph" ]
In the current study, the existence of periodic orbits around a fixed homogeneous cube is investigated, and the results have powerful implications for examining periodic orbits around non-spherical celestial bodies. In the two different types of symmetry planes of the fixed cube, periodic orbits are obtained using the method of the Poincar\'e surface of section. While in general positions, periodic orbits are found by the homotopy method. The results show that periodic orbits exist extensively in symmetry planes of the fixed cube, and also exist near asymmetry planes that contain the regular Hex cross section. The stability of these periodic orbits is determined on the basis of the eigenvalues of the monodromy matrix. This paper proves that the homotopy method is effective to find periodic orbits in the gravity field of the cube, which provides a new thought of searching for periodic orbits around non-spherical celestial bodies. The investigation of orbits around the cube could be considered as the first step of the complicated cases, and helps to understand the dynamics of orbits around bodies with complicated shapes. The work is an extension of the previous research work about the dynamics of orbits around some simple shaped bodies, including a straight segment, a circular ring, an annulus disk, and simple planar plates.
astro-ph.EP
astro-ph
Periodic orbits in the gravity field of a fixed homogeneous cube Xiaodong Liu1, Hexi Baoyin2, and Xingrui Ma3 School of Aerospace, Tsinghua University, Beijing 100084, China Email: [email protected]; [email protected]; [email protected] Abstract In the current study, the existence of periodic orbits around a fixed homogeneous cube is investigated, and the results have powerful implications for examining periodic orbits around non-spherical celestial bodies. In the two different types of symmetry planes of the fixed cube, periodic orbits are obtained using the method of the Poincaré surface of section. While in general positions, periodic orbits are found by the homotopy method. The results show that periodic orbits exist extensively in symmetry planes of the fixed cube, and also exist near asymmetry planes that contain the regular Hex cross section. The stability of these periodic orbits is determined on the basis of the eigenvalues of the monodromy matrix. This paper proves that the homotopy method is effective to find periodic orbits in the gravity field of the cube, which provides a new thought of searching for periodic orbits around non-spherical celestial bodies. The investigation of orbits around the cube could be considered as the first step of the complicated cases, and helps to understand the dynamics of orbits around bodies 1 PhD candidate, School of Aerospace, Tsinghua University 2 Associate Professor, School of Aerospace, Tsinghua University 3 Professor, School of Aerospace, Tsinghua University 1 with complicated shapes. The work is an extension of the previous research work about the dynamics of orbits around some simple shaped bodies, including a straight segment, a circular ring, an annulus disk, and simple planar plates. Keywords Cube · Periodic orbits · Homotopy method · Poincaré surface of section·Stability·Gravity 1. Introduction Previous literature has mentioned that to study orbits around the Platonic solids might be fruitful (Werner 1994). The investigation of orbits around these simple shaped bodies could be considered as the first step of the complicated cases, and helps to understand the dynamics of orbits around bodies with complicated shapes. The cube, which is a simple one of the five Platonic solids, is the subject of this paper. The dynamics of a particle orbiting around a fixed homogeneous cube are complex despite the cube’s simple shape, and several types of periodic orbits are found. Some relevant research was made in the past. It was once proved that there exist ring-type bounded planar motions in an isolated system consisting of a homogeneous cube and a massive point particle (Michalodimitrakis and Bozis 1985). In our previous research, equilibria of motion around a rotating cube were derived, and invariant manifolds connecting periodic orbits around the equilibria were calculated (Liu et al. 2011). 2 A new thought of searching for periodic orbits around non-spherical bodies is provided in this paper. The cube, as a simple case of non-spherical bodies, is taken as an example to prove the effectiveness of this new thought, i.e. the homotopy method. In artificial satellite theory, periodic orbits around non-spherical celestial bodies are of special interest. Periodic orbit families about asteroid 4769 Castalia were computed using the Poincaré maps and Newton–Raphson iteration (Scheeres et al. 1996). For asteroid 4179 Toutatis in a non-principal-axis rotation state, periodic orbits were found by using the averaged equations of motion to generate the initial conditions and computing the state transition matrix to adjust them (Scheeres et al. 1998). The computation of periodic orbits around asteroid 433 Eros was made with the application of the Poincaré maps and the monodromy matrix (Scheeres et al. 2000). In this paper, the homotopy method, which does not make use of symmetrical characteristic, is proposed to find periodic orbits. It is proved effective for the case of the cube in Sect. 5, and is also applicable to search for periodic orbits around other non-spherical bodies. Although the homotopy method has been introduced for many years, so far there is little literature about the application of the homotopy method to search for periodic orbits around celestial bodies. This study can be considered as an extension of the previous researches about the cases of some simple shaped massive bodies with backgrounds in celestial mechanics. For fixed massive bodies, periodic orbits and other dynamics characteristics around a straight segment (Riaguas et al. 1999; Arribas and Elipe 2001), a solid ring (Broucke and Elipe 2005; Azevêdo et al. 2007; Azevêdo and Ontaneda 2007; Fukushima 2010), 3 a homogeneous annulus disk (Alberti and Vidal 2007; Fukushima 2010), and simple planar plates including square and triangular plates (Blesa 2006) were investigated. The dynamics around a rotating segment were extensively studied (Riaguas et al. 2001; Elipe and Riaguas 2003; Gutiérrez–Romero et al. 2004; Palacián et al. 2006). In these previous studies, the simple shaped bodies are limited to one- or two-dimensional cases. In contrast, the current study extends the simple shaped bodies to three dimensions. Moreover, this study contributes to the investigation of irregular shaped celestial bodies that can be divided into a set of the cubes. In this way, the potential of the celestial bodies can be obtained by summing the potentials of all cubes. To get a clear understanding of the dynamics of orbits around the cube, which can be taken as the basic unit, provides valuable information and help to investigate the irregular shaped celestial bodies. The traditional method to represent the gravity field of celestial bodies is to expand the potential into series of spherical harmonics, which is effective for spheroid-like bodies. However, for irregular shaped celestial bodies, such as asteroids and comet nuclei, this method may fail to converge when within the circumscribing sphere centered at the harmonic expansion center and surrounding the shape. To remedy drawbacks of the harmonic expansion, the mascon approximation (Geissler et al. 1996; Scheeres et al. 1998) and the polyhedral approach (Werner 1994; Werner and Scheeres 1997) were proposed. Based on the polyhedral approach, irregular shaped celestial bodies can also be divided into a finite number of simple-shaped basic units, such as cubes or tetrahedral (Werner and Scheeres 1997). 4 The gravity field of 433 Eros was once modeled as the summation of gravity fields of a set of tetrahedra (Scheeres et al. 2000). In this study, the dynamics of orbits around a fixed homogeneous cube are considered. The potential of the fixed cube is calculated by the polyhedral approach. Periodic orbits in the symmetry planes are derived by computing the Poincaré surface of section. While in General Positions of the cube, periodic orbits are found via the homotopy method. It is proved that periodic orbits exist extensively in symmetry planes of the fixed cube, and also exist near asymmetry planes that contain the regular Hex cross section. The stability of these periodic orbits is computed. 2. The gravitational potential of the homogeneous cube Given a particle P with infinitesimal mass located outside of the fixed homogeneous cube with edge length 2a and constant mass density σ, the potential at a , certain point P can be written as an integral over the volume of the body 1 d  U G V  r V where G is the gravitational constant, and r is the distance from the point P to the (1) differential mass of the cube. Establish a right-handed Cartesian coordinate system Oxyz with the origin O located at the center of the fixed cube, and the three coordinate axes coinciding with the symmetry axes of the cube. The polyhedral approach (Werner and Scheeres 1997) is used to calculate the potential of the cube. The result is provided in the Appendix. Compared to the finite 5 straight segment’s potential (Riaguas et al. 1999), the solid circular ring’ potential (Broucke and Elipe 2005), and the homogeneous annulus disk’ potential (Azevêdo and Ontaneda 2007), the cubic potential is much more complex. It is evident that the potential is symmetrical with respect to the planes passing through the center, and parallel to the face as well as to the diagonal planes of the cube. In this study, scaling is performed such that the half-length of the edge a is the unit of length, and it is assumed that G in order to simplify the computation. 1 The motion equation of the particle in the gravity field of the fixed cube can be obtained U r . (2) 3. Periodic orbits in the symmetry planes parallel to the face of the fixed cube It can be seen in Fig. 1 that xy-plane is one symmetry plane parallel to the face of the cube. Given the particle P that is initially in an orbit tangent to the xy-plane, then the motion of P will stay in the xy-plane due to the symmetrical characteristic of the potential. It is evident that the conservative dynamics system only has two degrees of freedom, and the motion of the particle in z direction is zero all the time. In this section, the method of the Poincaré surface of section is used to find periodic orbits. By integrating many orbits with the same value of the Energy constant, the Poincaré surface of section can be obtained by plotting   ,x x whenever the particle crosses the x-axis. 6 The Poincaré surfaces of section for two different energy levels E=-1.2 and E=-1.0 are shown in Fig. 2. The case of E=-1.2 is used in this study. One central point that corresponds to the stable periodic orbit in the xy-plane can be noticed. z 1 0 -1 1 0 y -1 -1 0 x 1 Fig. 1 One symmetry plane parallel to the face of the cube: xy-plane. x y t i c o l e V 4 2 0 -2 -4 0 2 4 Coordinate x 6 8 (a) 7 x y t i c o l e V 1 0.5 0 -0.5 -1 2 3 4 Coordinate x 5 6 (b) Fig. 2 Poincaré surface of section for orbits in the xy-plane. (a) E=-1.2. (b) E=-1.0. From the Poincaré surface of section shown in Fig. 2(a), the good initial condition for the periodic orbit can be obtained x 0  3.34,  y 0 1.543321002314535 . (3) The orbit with this initial condition of about 100 revolutions is shown in Fig. 3. It can be seen that the orbit is periodic, and symmetric with respect to both the x-axis and the y-axis. y 4 2 0 -2 -4 -4 -2 0 x 2 4 Fig. 3 Periodic orbit in the xy-plane for E=-1.2. 8 4. Periodic orbits in the diagonal plane of the fixed cube In the diagonal plane of the fixed cubic central body, the existence of the periodic orbits can also be proved. It is evident in Fig. 4 that the diagonal plane is also the symmetry plane of the fixed cube. Another right-handed Cartesian coordinate system Ox1y1z1 is established with three coordinate axes i1, j1, k1 parallel to i-k, j, i+k, respectively. Given the particle P that is initially in an orbit tangent to the diagonal plane, then the motion of P will stay in this diagonal plane due to the symmetric characteristic of the potential. It is evident that the conservative dynamics system also only has two degrees of freedom, and the motion in z1 direction is zero all the time. Following the same process as Sect. 3, periodic orbits in the diagonal plane can also be found. The Poincaré surface of section for energy E=-1.2 is shown in Fig. 5. One central point that corresponds to the stable periodic orbit in the diagonal plane can be noticed. 9 z 1 0 -1 1 0 -1 y -1 1 0 x Fig. 4 One diagonal plane of the cube. 1 0.5 0 -0.5 1 x y t i c o l e V -1 2.5 3 3.5 Coordinate x1 4 Fig. 5 Poincaré surface of section for orbits in the x1y1-plane with E=-1.2. From the Poincaré surface of section shown in Fig. 5, the good initial condition for the periodic orbit can be obtained x 10  3.337544007200505,  y 10 1.547808464792009 . (4) 10 The orbit with this initial condition of about 100 revolutions is shown in Fig. 6. It shows that this orbit is periodic, and is symmetric with respect to both the x1-axis and y1-axis. 1 y 4 2 0 -2 -4 -4 -2 0 x1 2 4 Fig. 6 Periodic orbit in the x1y1-plane for E=-1.2. 5. Searching of periodic orbits in general positions of the fixed cube In the asymmetry planes of the fixed cube, searching of periodic orbits is much more difficult than the symmetry cases because of the complexity of the potential of the fixed cube. The main difficulties come from the fact that the shape of the cube is not as smooth as that of the sphere. In this section, the process involving with the homotopy method and Gauss-Newton algorithms is applied to overcome these difficulties. The homotopy method is a type of topology approach. A certain homotopy owns the right properties connecting the two problems, and the process of the homotopy method is to solve a more difficult one by starting from an easier one. 11 Let f and g be two mathematical objects from a space X to a space Y. If there exists a continuous map from X   0, 1 Y  such that  H x , 0    f x  and  H x  , 1   g x  , then the two mathematical are said to be homotopic. In this study, the searching of periodic orbits of the fixed cube leads to a difficult problem, while searching for periodic orbits of the fixed sphere is much easier. It is obvious that the shape of the cube and the shape of the sphere are homotopic, so are the potential of the fixed cube and the potential of the fixed sphere. Therefore, a homotopy can be defined to connect the potentials of the fixed cube and the fixed sphere H   U    cube  1     U sphere , (5) where cubeU is the potential of the fixed cube; sphereU is the potential of the fixed sphere. It can be easily derived that H   0  U , H   1  U cube . sphere (6) So the homotopy links the sphere problem (when 0 ) and the cube problem (when 1 ). The homotopy process is to start at 0 , and increase  up to 1. The condition for 0 , corresponding to the sphere problem, is just Kepler problem, and much more regular. The initial guess for periodic orbits in this case is chosen as Kepler orbit elements. In this problem, ε is increased from 0 up to 1 by the step size 0.01. For each step, the initial value of the iteration is taken as the periodic condition at the preceding step, and is integrated over one period T; the periodic condition for this step is obtained by Gauss-Newton algorithms and the process is iterated until the 12 mismatch between the position and velocity at the initial time t0 and the position and velocity at the final time t0+T is less than 10-8. Finally, when the value of ε reaches 1, the condition for periodic orbits of the fixed cube can be obtained. The stability of these periodic orbits can be determined by the eigenvalues of the monodromy matrix. The system is stable only if all eigenvalues of the monodromy matrix are located on the unit circle. The states of stability of periodic orbits appear in the eleventh column of Table 1, and each orbit is denoted by U (unstable) or S (stable). Several initial conditions in the sphere problem are selected in order to obtain periodic orbits in the cube problem. The inclinations of these initial conditions are free parameters different from 0 to 90 , and the initial semimajor axes a are 4 or 5; while the other four Keplerian orbits elements of the these conditions are the same, with eccentricity e  , right ascension of the ascending node 0 20   , argument of pericentre 0  , and true anomaly f  30  . Via the homotopy method, each initial condition in the sphere problem leads to a periodic orbit around the cube. These periodic orbits around the cube can be divided into three different families: orbits in the planes parallel to the face of the cube, orbits in the diagonal plane, and orbits in the asymmetry planes that contain the regular Hex cross section (One regular Hex cross section of the cube is shown in Fig. 7). These families of about 100 revolutions are shown in Figs. 8, 9 and 10. All the necessary data of these periodic orbits are given in Table 1. 13 Table 1 Data of periodic orbits exposed in Sect. 5. a 4 5 4 5 4 5 Orbit Family A A B B C C 1 2 3 4 5 6 T (s) Stability 24.8498127188 24.8953775322 S S S S S S x0 y0 3.2367087394 3.8120772276 3.3076964041 3.7588046209 z0 0 0 vx0 vy0 -0.9640378049 0.8184399380 -0.9488195578 0.8348743804 vz0 0 0 2.9799883678 3.9159049740 0.9352934196 -0.7070086935 0.2981079514 1.0045476967 24.8975943174 2.9571756627 3.9308880115 0.9730898238 -0.7135375880 0.2888042290 1.0017890027 24.9381286473 3.7283776591 2.3647182241 2.3647182236 -0.8436335399 0.6657159417 0.6657159418 24.8967459026 3.6627045131 2.4102817833 2.4102817829 -0.8612089381 0.6549877885 0.6549877885 24.8589057477 14 1 0.5 0 -0.5 z -1 1 y 0 -1 -1 0 x 1 Fig. 7 One regular Hex cross section of the cube. 1 0 -1 z 5 0 y -5 1 0.5 z 0 -0.5 -1 -10 -5 10 5 0 -5 y Orbit 1 Orbit 2 5 0 -10 -10 0 x -5 x (a) (b) Orbit 1 Orbit 2 1 0.5 z 0 -0.5 0 x 5 10 (c) (d) -1 -10 -5 0 y Orbit 1 Orbit 2 10 Orbit 1 Orbit 2 5 10 15 Fig. 8 Periodic orbits of family A. (a) Orbits in the three-dimensional space; (b) Projection of orbits onto the xy-plane; (c) Projection of orbits onto the xz-plane; (d) Projection of orbits onto the yz-plane. x 10-3 2 0 2 z -2 10 0 y2 -10 2 x 10-3 2 z 1 0 -1 -2 -10 -5 Orbit 3 Orbit 4 10 Orbit 3 Orbit 4 5 10 0 x2 0 y2 Orbit 3 Orbit 4 5 2 y 10 5 0 -5 -10 -10 -5 0 x2 (a) (b) 2 x 10-3 Orbit 3 Orbit 4 2 z 1 0 -1 0 x2 5 10 -2 -10 -5 (c) (d) Fig. 9 Periodic orbits of family B. The Cartesian coordinate system Ox2y2z2 is established with the origin O located at the center of the fixed cube and x2y2-plane coinciding with the plane that contains the regular Hex cross section. (a) Orbits in the three-dimensional space; (b) Projection of orbits onto the x2y2-plane; (c) Projection of orbits onto the x2z2-plane; (d) Projection of orbits onto the y2z2-plane. 16 z 5 0 -5 5 0 y -5 z 10 5 0 -5 -10 -10 Orbit 5 Orbit 6 10 Orbit 5 Orbit 6 3 6 0 x 0 y Orbit 5 Orbit 6 10 0 10 5 0 -5 y -10 -10 Orbit 5 Orbit 6 6 3 0 -3 z -10 x (a) (b) 0 x 10 (c) (d) -6 -6 -3 Fig. 10 Periodic orbits of family C. (a) Orbits in the three-dimensional space; (b) Projection of orbits onto the xy-plane; (c) Projection of orbits onto the xz-plane; (d) Projection of orbits onto the yz-plane. Family A (Fig. 8). These orbits are located in the xy-plane (or other symmetry planes parallel to the face of the cube), and are symmetric with respect to both the x-axis and the y-axis. It can be seen that these orbits are all quasi-circular. Three pairs of eigenvalues of the monodromy matrix are all located on the unit circle, so these orbits are stable. Family B (Fig. 9). 17 These orbits are located near the asymmetry plane that contains the regular Hex cross section of the cube, and are not planar orbits. The regular Hex cross section divides the cube into two equal parts; however, the two parts are asymmetry. It can be seen that the amplitude of motion in z2 direction is three orders of magnitude lower than that in x2 and y2 directions. Three pairs of eigenvalues of the monodromy matrix are all located on the unit circle, so these orbits are stable. Family C (Fig. 10). These orbits lie in one diagonal plane that is perpendicular to the yz-plane (or other diagonal planes), and are symmetric with respect to the x-axis. Three pairs of eigenvalues of the monodromy matrix are all located on the unit circle, so these orbits are stable. The cube owns three symmetry planes parallel to the face of itself, six symmetry planes in the diagonal plane, and four asymmetry planes that contain the regular Hex cross section. It can be seen that these periodic orbits are located either in the symmetry planes, or near the asymmetry planes that contain the regular Hex cross section. Periodic orbits of these three families are all stable. 6. Conclusions This paper investigates the dynamics of orbits in the gravity field of a fixed homogeneous cube, and finds several types of periodic orbits. Based on the method of the Poincaré surface of section, periodic orbits in the two different types of symmetry 18 planes of the fixed cube are easily found. While in general positions, periodic orbits can be found by the homotopy method. Simulations show that periodic orbits around the cube can be divided into three different families: orbits in the planes parallel to the face of the cube, orbits in the diagonal plane, and orbits in the asymmetry planes that contain the regular Hex cross section. The stability of these periodic orbits is determined based on the eigenvalues of the monodromy matrix. The results demonstrate that periodic orbits of these three families are all stable. The work about dividing irregular shaped celestial bodies into a set of the cubes is in progress. 19 Appendix: The potential of the cube Based on the polyhedral method (Werner and Scheeres 1997), the potential of the fixed cube at a certain point P (x, y, z) in space is calculated as  2   U G      x a   2   x a z a     2  x a    y a     y a   ln  2 x a    y a   ln 2 2 a a   ln r r   4 1 r r   4 1 a r 2   6 a r 2   6 a r 2   2 a r 2   2 a r 2   4 a r 2   4 a r 2   6 r a 2   6 r a 2   8 a r 2   8    r 2 r 2 r 1 r 1 r 3 r 3 r 5 r 5 r 7 r 7 2    x a z a    ln 2  x a    y a   ln 2    x a z a    ln 2    x a z a    ln 2 2 a a ln r r   5 1 r r   5 1 a r 2   7 r a 2   7 a r 2   3 a r 2   3 a r 2   8 r a 2   8 a r 2   7 r a 2   7 r a 2   8 a r 2   8 r 3 r 3 r 2 r 2 r 4 r 4 r 6 r 6 r 5 r 5  2    y a z a    ln  2    y a z a    ln  2    y a z a    ln  2    y a z a    ln arctan arctan arctan arctan  2  a z  2   2  a z  2   2  a x  2   2  a x  2   2  a y  2   2  a  2 y                    r r r 6 7 8  r 6  r r r 2 3 6  r 2  r r r 1 4 5  r 1   r r 4 5 arctan r r r 3 4 7  r 3   r r r  3 2 1    r r r  3 1 2  r r r  6 7 8     r r r r r  8 6 7 8 7  r r r  3 6 2     r r r r r  3 2 6 6 3  r r r  5 4 1    r r r  1 4 5  r r r  3 4 7     r r r r r  3 4 7 7 4  r r r 5 6 1   r r r  6 1 5    arctan r r r 1 5 6  r 1   r r 5 6  r 6   r r 5 1   arctan r r r 1 2 6  r 1   r r 2 6 where the Cartesian coordinate system Oxyz is established with the origin O located at the center of the fixed cube and the three coordinate axes coinciding with the symmetrical axes of the cube; G is the gravitational constant; σ is the constant mass density of the cube; a is the half-length of the cubic edge; ri (i=1,2,…,8) is the vector from the origin O to one of the eight vertices of the cube, and ri is the norm of ri. 20 r r r 1 2 3  r 1   r r 2 3  r 3   r r 2 1   arctan r r r 1 3 4  r 1    r r 3 4   r 8   r r 6 7   arctan r r r 5 6 8  r 5    r 3   r r 2 6   arctan r r r 3 6 7  r 6    r 5   r r 4 1   arctan r r r 4 5 8  r 4    r 7   r r 3 4   arctan r r r 4 7 8  r 4      r r 6 8  r r r 3 1 4   r r r  3 1 4  r r r  5 6 8    r r r  6 8 5  r r r  3 7 6     r r r r r  7 3 6 7 3  r r r  8 4 5    r r r  4 8 5  r r r  8 7 4     r r r r  4 7 8 8  r r r 6 2 1   r r r  6 1 2   r 7    r r 5 8          r 4   r r 1 3   r 8   r r 6 5      r 3   r r 6 7  r 5   r r 4 8          r r  4 7          ,   r 8   r 6   r r 1 2 Acknowledgements This research was supported by the National Natural Science Foundation of China (No. 10832004 and No. 11072122). References Alberti, A., Vidal, C.: Dynamics of a particle in a gravitational field of a homogeneous annulus disk. Celest. Mech. Dyn. Astron. 98(2), 75–93 (2007) doi: 10.1007/s10569-007-9071-z Arribas, M., Elipe, A.: Non-integrability of the motion of a particle around a massive straight segment. Phys. Lett. A. 281(2/3), 142–148 (2001) Azevêdo, C., Cabral, H. E., Ontaneda, P.: On the fixed homogeneous circle problem. Adv. Nonlinear Stud. 7(1), 47–75 (2007) Azevêdo, C., Ontaneda, P.: On the existence of periodic orbits for the fixed homogeneous circle problem. J. Differ. Equ. 235(2), 341–365 (2007) Blesa, F.: Periodic orbits around simple shaped bodies. Monografías del Seminario Matemático García de Galdeano. 33, 67–74 (2006) Broucke, R. A., Elipe, A.: The dynamics of orbits in a potential field of a solid circular ring. Regul. Chaotic. Dyn. 10(2), 129–143 (2005) Elipe, A., Riaguas, A.: Nonlinear stability under a logarithmic gravity field. Int. Math. J. 3, 435–453 (2003) 21 Fukushima, T.: Precise computation of acceleration due to uniform ring or disk. Celest. Mech. Dyn. Astron. 108(4), 339–356 (2010). doi: 10.1007/s10569-010-9304-4 Geissler, P., Petit, J.-M., Durda, D., Greenberg, R., Bottke, W., Nolan, M., Moore, J.: Erosion and ejecta reaccretion on 243 Ida and its moon. Icarus. 120(42), 140–157 (1996) Gutiérrez–Romero, S., Palacián, J. F., Yanguas, P.: The invariant manifolds of a finite straight segment. Monografías de la Real Academia de Ciencias de Zaragoza. 25, 137–148 (2004) Liu, X., Baoyin, H., Ma, X.: Equilibria, periodic orbits around equilibria, and heteroclinic connections in the gravity field of a rotating homogeneous cube. Astrophys. Space. Sci. (2011). doi: 10.1007/s10509-011-0669-y Michalodimitrakis, M., Bozis, G.: Bounded motion in a generalized two-body problem. Astrophys. Space Sci. 117(2), 217–225 (1985) Palacián, J. F., Yanguas, P., Gutiérrez–Romero, S.: Approximating the invariant sets of a finite straight segment near its collinear equilibria. SIAM J. Appl. Dyn. Syst. 5(1), 12–29 (2006) Riaguas, A., Elipe, A., Lara, M.: Periodic orbits around a massive straight segment. Celest. Mech. Dyn. Astron. 73(1/4), 169–178 (1999) Riaguas, A., Elipe, A., López-Moratalla, T.: Non-linear stability of the equilibria in the gravity field of a finite straight segment. Celest. Mech. Dyn. Astron. 81(3), 235–248 (2001) Scheeres, D. J., Marzari, F., Tomasella, L., Vanzani, V.: ROSETTA mission: satellite orbits around a cometary nucleus. Planet. Space Sci. 46(6/7), 649–671 (1998) 22 Scheeres, D. J., Ostro, S. J., Hudson, R. S., Dejong, E. M., Suzuki, S.: Dynamics of orbits close ateroid 4179 Toutatis. Icarus. 132(1), 53–79 (1998). doi: 10.1006/icar.1997.5870 Scheeres, D. J., Ostro, S. J., Hudson, R. S., Werner, R. A.: Orbits close to asteroid 4769 Castalia. Icarus. 121, 67–87 (1996). doi: 10.1006/icar.1996.0072 Scheeres, D. J., Williams, B. G., Miller, J. K.: Evaluation of the dynamic environment of an asteroid: applications to 433 Eros. J. Guid. Control. Dyn. 23(3), 466–475 (2000) Werner, R. A.: The gravitational potential of a homogeneous polyhedron or don’t cut corners. Celest. Mech. Dyn. Astron. 59(3), 253–278 (1994) Werner, R. A. Scheeres, D. J.: Exterior gravitation of a polyhedron derived and compared with harmonic and mascon gravitation representations of asteroid 4769 Castalia. Celest. Mech. Dyn. Astron. 65(3), 313–344 (1997) 23
0909.2287
1
0909
2009-09-14T17:51:21
High inclination orbits in the secular quadrupolar three-body problem
[ "astro-ph.EP" ]
The Lidov-Kozai mechanism allows a body to periodically exchange its eccentricity with inclination. It was first discussed in the framework of the quadrupolar secular restricted three-body problem, where the massless particle is the inner body, and later extended to the quadrupolar secular nonrestricted three body problem. In this paper, we propose a different point of view on the problem by looking first at the restricted problem where the massless particle is the outer body. In this situation, equilibria at high mutual inclination appear, which correspond to the population of stable particles that Verrier & Evans (2008,2009) find in stable, high inclination circumbinary orbits around one of the components of the quadruple star HD 98800. We provide a simple analytical framework using a vectorial formalism for these situations. We also look at the evolution of these high inclination equilibria in the non restricted case.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 21 November 2018 (MN LATEX style file v2.2) High inclination orbits in the secular quadrupolar three-body problem F. Farago⋆ and J. Laskar⋆ ASD, IMCCE-CNRS UMR8028, Observatoire de Paris, UPMC, 77 avenue Denfert-Rochereau, 75014 Paris, France 21 November 2018 ABSTRACT The Lidov-Kozai mechanism (Kozai 1962; Lidov 1962) allows a body to periodically exchange its eccentricity with inclination. It was first discussed in the framework of the quadrupolar secular restricted three-body problem, where the massless particle is the inner body, and later extended to the quadrupolar secular nonrestricted three body problem (Harrington 1969; Lidov & Ziglin 1976; Ferrer & Osacar 1994). In this paper, we propose a different point of view on the problem by looking first at the restricted problem where the massless particle is the outer body. In this situation, equilibria at high mutual inclination appear (Palaci´an et al. 2006), which correspond to the population of stable particles that Verrier & Evans (2008, 2009) find in stable, high inclination circumbinary orbits around one of the components of the quadruple star HD 98800. We provide a simple analytical framework using a vectorial formalism for these situations. We also look at the evolution of these high inclination equilibria in the non restricted case. Key words: celestial mechanics -- planetary systems -- methods: analytical -- methods: N -body simulations. 1 INTRODUCTION As it is known, the secular three-body problem after node reduction has two degrees of freedom (e. g. Poincar´e (1905); Malige et al. (2002)). However, due to what Lidov & Ziglin (1976) called a happy coincidence, this problem is integrable when it is expanded up to order two in the ratio of semi- major axes, i.e. at the quadrupolar approximation. Indeed, the argument of perihelion of the outer body does not ex- plicitly appear in the quadrupolar expansion of the secular problem, thus giving one more integral of motion linked to the eccentricity of the outer body. The limiting case where the inner body has no mass has been extensively studied (Kozai 1962; Lidov 1962; Kinoshita & Nakai 2007). We will call this problem the in- ner restricted problem, while the converse case where the two inner bodies are massive and the outer body is massless will be called the outer restricted problem. In the inner re- stricted case, the conservation of the normal component of the angular momentum enables the inner particle to periodi- cally exchange its eccentricity with inclination (the so-called Lidov-Kozai mechanism). The inner restricted model is well suited when the inner body has a small mass with respect to the other two. However, when looking at higher mass ratios, for example in triple star systems, this is no longer justified. ⋆ E-mail: [email protected]; [email protected] Since the Hamiltonian of the quadrupolar problem of three masses is very similar to that of the inner restricted problem when it is written in elliptic variables, the study of the massive problem has mainly focused on the dynamics of the two inner bodies (Harrington 1969; Lidov & Ziglin 1976; Ferrer & Osacar 1994). These previous works com- pletely classified the different dynamical regimes and bifur- cations, using the equations of motion of the inner binary. There is however another limit-case to the massive prob- lem, which is the outer restricted problem. Palaci´an et al. (2006) have studied this case and discussed the existence and stability of equilibria in the non-averaged system using the framework of KAM theory. We give here a very sim- ple model of the outer restricted case which provides an alternate formulation of these previous results and which is directly usable in an astronomical context. We also fully describe the possible motions of the bodies and give an an- alytical expression of their frequencies. We use this model to explain the results of Verrier & Evans (2008, 2009), who find populations of particles at very high inclinations around one of the components of the double-binary star HD 98800, which are stable even under the perturbation of the other component. We then look at the quadrupolar problem of three masses from the perspective of the outer restricted problem and show how the inner and outer restricted cases are related to the general case. Vectorial methods as devel- 2 F. Farago and J. Laskar oped by Bou´e & Laskar (2006, 2009); Tremaine et al. (2009) are extremely well suited for this approach. 2 SECULAR OUTER RESTRICTED PROBLEM 2.1 Derivation of the Hamiltonian We consider here the case of a massless particle orbit- ing a central binary object. We do not restrict ourselves with respect to inclinations or eccentricities. The compo- nents of the binary have masses m0 and m1, the binary's total mass is M01 = m0 + m1 and its reduced mass is β1 = m0m1/(m0 + m1). The two massive components have barycentric positions u0 and u1. We also denote δ = m0/M01 and r1 = u1 − u0, and r2 is the position of the outer particle relatively to the barycentre of the inner bi- nary. Using these notations, the particle has the following Hamiltonian: H = r2 2 2 − G„ m0 r2 − u0 + m1 r2 − u1« , (2.1) where r2 = r2 is the canonical momentum associated to the position of the massless particle, r2. Since u0 = −(1 − δ)r1 and u1 = δr1, we can rewrite the Hamiltonian as: H = r2 2 2 − G„ m0 r2 + (1 − δ)r1 + m1 r2 − δr1« . (2.2) We now suppose that r1 ≪ r2 and expand the Hamil- tonian to order 2 in r1/r2: 2 H =„ r2 2 − r2 « − GM01 Gβ1 2r3 2 „3 (r2.r1)2 r2 2 1« . − r2 (2.3) The first term is the Keplerian energy of the particle interacting with the binary, seen as a point mass M01. It is equal to −GM01/2a2, where a2 is the semi major axis of the particle. Since we are interested in the secular behaviour of the particle, we average this quadrupolar Hamiltonian over the mean anomalies of the binary (M1) and of the particle (M2). In order to do this, we first introduce four unit vectors: (i, j, k) are bound to the orbit of the binary, remain constant, and will provide a natural reference frame; w is bound to the orbit of the particle and will vary. More precisely, i points in the direction of the perihelion of the binary, k is colinear to the angular momentum of the binary, and j = k ∧ i; the last vector w is colinear to the angular momentum of the massless particle. We can then compute the following averaged quantities, where quantities indexed with 1 relate to the binary, quanti- ties with index 2 relate to the particle, and u is an arbitrary fixed vector (see for instance the appendix of Bou´e & Laskar (2006)): r2 1¸M1 (r1.r2)2¸M1 r3 2flM2 fi 1 fi (r2.u)2 2 flM2 r5 = = = = a2 + e2 3 2 a2 1 2 1„1 + 1« ; 2 − (k.r2)2) (r2 1e2 a2 (4(i.r2)2 − (j.r2)2) ; 1 2 1 2)3/2 ; a3 2(1 − e2 u2 − (w.u)2 2(1 − e2 2a3 2)3/2 . (2.4) (2.5) (2.6) (2.7) The substitution of these expressions in (2.3) yields: hHiM1,M2 »„e2 1 − 3 8 GM01 = − 2a2 − 3« + (k.w)2 − e2 1 Gβ1a2 1 2(1 − e2 a3 1(4(i.w)2 − (j.w)2) -- 2)3/2 × (2.8) Since the particle has no mass, the only variable ele- ment of the binary is its mean anomaly M1. After averag- ing over this angle, it is no longer present in the Hamilto- nian. After averaging over the mean anomaly of the par- ticle, its semi major axis a2 becomes constant. Moreover, w = sin i2 sin Ω2 i − sin i2 cos Ω2 j + cos i2 k, so the argu- ment of pericentre ω2 of the particle does not appear in the averaged Hamiltonian. Hence, at the quadrupolar order, the conjugate momentum associated to ω2, i.e. the norm of the 2), is constant. Therefore the eccentricity e2 of the particle is constant. This fact is a feature of the quadrupolar expan- sion, not a property of the restricted problem. As such it remains true when the outer body has a non-zero mass (see section 3). This is the happy coincidence that Lidov & Ziglin (1976) noted. Finally, only one degree of freedom remains, related to the couple (i2, Ω2). angular momentum of the particle G2 =pGM01a2(1 − e2 If we drop the constant terms in (2.8), and introduce the mean motion n1 of the binary into the Hamiltonian (n2 1 = GM01), we get the following expression1 (see also eq. 10 in (Palaci´an et al. 2006)): 1a3 hHi = − where αG2 2 (k.w)2 − e2 1(4(i.w)2 − (j.w)2) , α = 3 4 n1„ a1 a2«7/2 β1 M01 1 (1 − e2 2)2 . (2.9) (2.10) This Hamiltonian can be rewritten in a very compact form as: 1 2 hHi = − T = αG20 @ tw.T.w , where: (2.11) 1 −4e2 0 0 0 e2 1 0 0 0 1 1 A . (2.12) 1 We will from now write hHi for the averaged Hamiltonian, omit- ting the subscripts M1, M2. High inclination orbits in the secular quadrupolar three-body problem 3 We also give the expression of the Hamiltonian using the inclination and the node of the particle: hHi = − αG2 4 2 cos2 i2 − e2 1 sin2 i2 (3 − 5 cos 2Ω2) . (2.13) 2.2 Equations of motion As discussed in (Bou´e & Laskar 2006), the equations of mo- tion for w are simply obtained by: w = 1 G2 ∇w hHi ∧ w . (2.14) After computing the gradient, we find: w = −α(k.w)(k ∧ w) − e2 (2.15) If we note x = (i.w), y = (j.w), and z = (k.w), we get 1(4(i.w)(i ∧ w) − (j.w)(j ∧ w)) . the following system for (x, y, z): x = α(1 − e2 1)yz ; y = −α(1 + 4e2 z = 5αe2 1xy . 1)zx ; (2.16) (2.17) (2.18) In these variables, the fact that w is a unit vector and the energy integral translate into the following equalities: z2 − e2 x2 + y2 + z2 = 1 ; 1(4x2 − y2) = h = Cst . The system of three differential equations (2.16) -- (2.18) has thus two independent first integrals and is as such in- tegrable. It is also straightforward from these two relations that (2.20) (2.19) Figure 1. Intersections of the energy surfaces and the unit angu- lar momentum sphere (a) and its projection in the (x, y) plane (b) for e1 = 0.5. Intersections of the energy surfaces and the unit an- gular momentum sphere (c) and its projection in the (x, y) plane (d) for e1 = 0.2. This precession is equivalent to the precession generated by the quadrupolar potential of a circular and homogeneous ring of mass β1 and of radius a1 following an idea which can be traced back to Gauss (see (Touma et al. 2009) and references therein). 2.4 Motion of a massless body around an elliptic − 4e2 1 6 h 6 1 . (2.21) binary 2.4.1 Qualitative overview 2.3 Motion of a massless body around a circular binary In the case of a circular binary, the Hamiltonian and the equations of motion greatly simplify2: hHi = − αG2 2 (k.w)2 , w = −α(k.w)(k ∧ w) (2.22) (2.23) The scalar product (k.w) = cos i2 remains constant, and the nodes of the orbit of the particle simply precess around the angular momentum of the binary, with a con- stant precession rate: Ω = −α cos i2 = − 3 4 n1„ a1 a2«7/2 β1 M01 cos i2 (1 − e2 2)2 (2.24) 2 The next non-zero term of the Hamiltonian which is the fourth order in (a1/a2) plays an important part in the circular case as has been discussed in detail by Palaci´an & Yanguas (2006). When the binary is elliptic, the situation changes and can- not be explained any longer by the quadrupolar torque of a circular ring. If we substitute z2 in (2.20) using (2.19), we get: (1 + 4e2 1)x2 + (1 − e2 1)y2 = 1 − h > 0 , x2 + y2 + z2 = 1 . (2.25) (2.26) The intersections of the energy surfaces and the nor- malized angular momentum sphere of the particle can thus be seen as the intersections of elliptic cylinders with the unit sphere. For a given value of the energy h, the extremity of the unit angular momentum vector of the particle w will move on the intersection of the corresponding cylinder with the unit sphere. Figures 1.a and c show these intersections for different values of the energy as dotted lines drawn on the unit sphere, in two situations where the binary has an eccentricity of 0.5 and 0.2 respectively. The three axes cor- respond to the scalar products x, y and z that are defined in section 2.2. There are four visible kinds of trajectories: closed tra- jectories around the two poles of the sphere (x, y, z) = 4 F. Farago and J. Laskar (0, 0, ±1), and closed trajectories around the points (x, y, z) = (±1, 0, 0). When the extremity of the angular momentum of the particle w follows a trajectory around the north pole, it means that it precesses around the angular momentum of the binary k with an inclination that is strictly inferior to 90◦: in this case, the orbital motion of the particle is pro- grade relatively to the orbital motion of the binary. When the extremity of the angular momentum of the particle w follows a trajectory around the south pole, it means that it precesses around the opposite of the angular momentum of the binary, −k, with an inclination that is strictly superior to 90◦: in this case, the orbital motion of the particle is retrograde relatively to the orbital motion of the binary. When the extremity of the angular momentum of the particle w follows a trajectory around one of the two points (x, y, z) = (±1, 0, 0), it precesses around the direction of the perihelion of the binary or the opposite of this direction. In this case, the inclination oscillates around ±90◦. 2.4.2 Frequencies The frequencies of these motions can be found analytically. Indeed, using equation (2.25), we see that x and y are on ellipses or arcs of ellipses bounded by the unit circle (figures 1.b and d show respectively the cases where e1 = 0.5 and 0.2). Thus, there is an angle φ such that: 1 + 4e2 1 x = s 1 − h y = s 1 − h 1 − e2 1 cos φ , sin φ . (2.27) (2.28) Using (2.19), we can then express z2 as: z2 = h + 4e2 1 1 + 4e2 1 − 5e2 1 1 + 4e2 1 1 − h 1 − e2 1 sin2 φ . (2.29) There are two opposite values of z for each φ, reflecting the symmetry of the system with respect to the orbital plane of the binary. If we use expression (2.29) in combination with equation (2.18), we obtain the following equation for φ: φ = ∓αq(1 − e2 1 − h h + 4e2 1 5e2 1 1 − e2 k2 = 1 . 1)(h + 4e2 1)q1 − k2 sin2 φ , (2.30) (2.31) The constant k2 is positive because of relation (2.21). The value k2 = 1 defines a limit between two dynamical φ never van- regimes. If k2 < 1, or equivalently if h > e2 1, ishes and the projection of w on the orbital plane of the binary moves along the full ellipse (2.25). In this case, w precesses around the angular momentum of the binary, k. If z > 0 the mutual inclination of the two orbits is always less than 90◦ so the orbital motion of the particle is prograde; conversely, if z < 0 the mutual inclination of the two orbits is always superior to 90◦ so the orbital motion of the particle is retrograde. If k2 > 1 (or h < e2 1), then φ vanishes for φ0 = ± arcsin(1/k), changes its sign (which is accompanied by a change of sign in the z variable), and the angle φ librates between −φ0 and +φ0. Thus, the projection of w on the orbital plane of the binary is bounded by the unit circle to stay on an arc of ellipse (2.25). In this case, w precesses around the direction of perihelion of the binary, so that both the inclination and the node of the particle librate around ±90◦. with a simple quadrature using equation (2.30): In both cases, the period of the motion can be calculated T = 16 3n1 β1 „ a2 M01 a1«7/2 K(k2)(1 − e2 2)2 1)(h + 4e2 1) , (2.32) where K(k2) is the elliptic integral of the first kind defined by: p(1 − e2 K(k2) =8< : 0 R π/2 R φ0 0 dφ√1−k2 sin2 φ dφ√1−k2 sin2 φ if k2 < 1 if k2 > 1 . (2.33) The last case where k2 = 1 (or h = e2 1) corresponds to the trajectories that separate the previous two types. They link the points (x, y, z) = (0,±1, 0), and the associated pe- riod is infinite. In the projection on the (x, y) plane, these separatrices form the ellipse which is tangent to the unit circle. Since all trajectories that are inside this ellipse corre- spond to the precession of w around k, the width ∆xsep of the separating ellipse in the (x, y) plane gives an indication on the proportion of such trajectories. Using equation (2.27) and the fact that h = e2 1 on the separatrix, we get: ∆xsep = 2s 1 − e2 1 1 + 4e2 1 (2.34) Therefore, when the inner binary is circular, this width is equal to 2, the full width of the unit circle, and the only possible motion is precession of w around ±k. When the eccentricity of the binary increases, the width of the sepa- ratrix decreases to zero, which is a limit case since it can only be reached for a value of the binary's eccentricity equal to 1. The precession motions of w around ±i thus become predominant when the eccentricity of the binary grows. 2.5 Comparison with numerical studies In (Verrier & Evans 2009), the authors investigate a fam- ily of particles at high inclinations around the binary HD 98800 Ba-Bb, which remain stable even under the pertur- bation of an outer third stellar companion. They isolate a nodal precession imposed by the inner binary as the stabiliz- ing mechanism working against the destabilizing Kozai per- turbations of the outer companion. They run simulations of test particles orbiting the binary HD 98800 Ba-Bb using non secular equations. They observe the libration islands around i2 = ±90◦ and Ω2 = ±90◦ that we discussed in the previous section. As they show their results in the (i2 cos Ω2, i2 sin Ω2) plane, we plotted the energy levels of the outer restricted Hamiltonian using these same coordinates for an easier com- parison. Figure 2 shows these levels for different values of the eccentricity. The c. panel in particular uses the same value High inclination orbits in the secular quadrupolar three-body problem 5 Figure 2. Energy levels of the Hamiltonian (2.8) in the (i2 cos Ω2, i2 sin Ω2) plane for values of the eccentricity of the bi- nary e1 = 0 (a), e1 = 0.1 (b), e1 = 0.79 (c), e1 = 0.9 (d). for the eccentricity of the binary (e1 = 0.79) as figures 4 and 5 of (Verrier & Evans 2009). Verrier and Evans notice no apparent structure in the dynamics of the (e2, ω2) couple apart from the circulation of the perihelion. This is in agreement with the fact that the particle's eccentricity is constant at the quadrupolar ap- proximation. The authors also suggest that the projection of the an- gular momentum of test particles along the line of apses of the binary may be an integral of motion. From the re- sults of the previous section, it is straightforward to see that the projection x of the angular momentum of test particles along the line of apses of the binary is not constant. It varies with an amplitude that decreases to 0 when the inclination of the particle approaches ±90◦, which can be misleading when looking at numerical results for highly inclined par- ticles. However, the norm of the angular momentum of the test particles is an integral of the secular motion. Finally, the authors give a power-law fit of the period of the libration of the node with respect to three parameters: the eccentricity of the binary, the ratio of the semimajor axes a2/a1, and the mass ratio of the binary, δ. Their power-law is fitted using particles with fixed inclinations (85◦). They give in their equation (5): T ∝ e−1.1 1 δ−0.8„ a2 a1«3.37 . (2.35) By rewriting the mass dependences of equation (2.32), we get the following analytical dependence with respect to the mass ratio and the semi-major axis of the binary: Figure 3. Dependence of the period (2.32) with respect to the eccentricity of the binary, in normalized units. The full line corre- sponds to the calculated period, while the dashed line corresponds to the power-law fit given by Verrier & Evans (2009). We used a least squares method to fit the relative position of the two curves. els. The dependency with respect to e1 is rather complex in equation (2.32), and it is best compared in figure 3. The grid of initial conditions for the particles in (Verrier & Evans 2009) extends however from 3 to 10 AU for a binary separation of 1 AU, so the quadrupolar approx- imation may not be sufficient to fully describe the motion of the particles with the lowest semi major axes. In par- ticular, Verrier and Evans state that some low inclination particles show large eccentricity variations and even insta- bility. This could be due to a low initial semi-major axis and to resonances that are eliminated in our secular model by the averaging over the mean anomalies. 3 PROBLEM OF THREE MASSIVE BODIES As we already stated, the quadrupolar secular three-body problem is still integrable when all the bodies have positive masses. As such, it is possible to show how the outer re- stricted problem we discussed in the previous section relates to the general case, and to the inner restricted case studied by Kozai (1962) and Lidov (1962). We will first express the Hamiltonian of the secular quadrupolar problem using the same vectorial method as in the previous section in order to focus on the relative movements of the orbits. In their studies of the secular quadrupolar problem, Lidov & Ziglin (1976) and Ferrer & Osacar (1994) have shown that this problem depends on two parameters. We will then point out which regions of parameter space are topologically equivalent to the outer restricted case, and which regions correspond to the inner restricted case, in order to show the continuity that exists between both situations. T ∝„ a2 a1«3.5 (δ(1 − δ))−1 . (2.36) 3.1 Hamiltonian These two exponents compare very well with the fitted power law, in spite of the differences between the two mod- Let us consider three masses m0, m1 and m2, this time with m2 6= 0. We note the barycentric coordinates and impulsions 2 r1 « +„ r2 2β2 − 2 „3 β1m2 2r3 − G (r1.r2)2 r2 2 1« . − r2 (3.5) 6 F. Farago and J. Laskar (ui, ui)i=0,1,2. As in the previous section, we suppose that the two bodies of indices 0 and 1 form a binary and that the distance of the third body to this binary is much larger than the separation of the binary. We still note δ = m0/(m0+m1). We first perform a canonical change of variables to Jacobi coordinates, r0 = u0 r1 = u1 − u0 r2 = u2 − δu0 − (1 − δ)u1 r0 = u0 + u1 + u2 = 0 (3.1) r1 = u1 + (1 − δ)u2 r2 = u2 (3.2) (3.3) In these coordinates, the Hamiltonian of the three bod- ies is (Laskar 1989): r1 « + 1 µ1β1 H =„ r2 2β1 − − Gm2„ r2 2 2β2 m0 r2 + (1 − δ)r1 + m1 r2 − δr1« , where β−1 µ1 = G(m0 + m1) and µ2 = G(m0 + m1 + m2). 1 = m−1 0 + m−1 1 , β−1 2 = (m0 + m1)−1 + m−1 2 , to order two in r1/r2 as in the previous section: Using the fact that r1 ≪ r2, we expand the Hamiltonian H =„ r2 2β1 − r2 « µ1β1 µ2β2 1 The first two terms are Keplerian energies and are equal respectively to −µ1β1/2a1 and −µ2β2/2a2, where a1 and a2 are the semi major axes of the inner and the outer body in our system of coordinates. We now average over the two mean anomalies M1 and M2 in order to get the secular part of the Hamiltonian. We will define 4 unit vectors which are analogous to the 4 vectors we used in the first section: (i1, j1, k1) are tied to the orbit of the inner binary; i1 points in the direction of the perihelion of the inner binary, k1 points in the direction of its angular momentum, and j1 = k1 ∧ i1. The last vector k2 is colinear to the angular momentum of the outer body. In this section, the vectors tied to the orbit of the inner binary will no longer have fixed directions. Using the same averaging formulae as in the previ- ous section and using the fact that (i1.k2)2 + (j1.k2)2 + (k1.k2)2 = k2 2 = 1, we can write: hHiM1,M2 = − ×»− + 2e2 1 3 3 8 µ1β1 2a1 − 1 + (1 − e2 Gm2β1 µ2β2 2a2 − (1 − e2 1)(k1.k2)2 − 5e2 2)3/2 a2 1 a3 2 1(i1.k2)2 -- . (3.6) After averaging over the two mean anomalies, the semi- major axes are constant. There are thus four degrees of free- dom in the system, associated to the two eccentricities and the two inclinations. As we explained in the previous section, the argument of perihelion of the outer body does not appear in the quadrupolar expansion, and thus the norm of the an- 2), is constant. This implies that its eccentricity e2 is constant. Using the reduction of the nodes would leave only one de- gree of freedom in the reduced Hamiltonian, associated to gular momentum of the outer body G2 = β2pµ2a2(1 − e2 use the above vectors, we get: (3.4) the couple (e1, ω1). The full reduction of the Hamiltonian and its expression in elliptical variables is the approach that has been used widely, since it yields a very similar Hamilto- nian function as in the inner restricted problem (Harrington 1969; Lidov & Ziglin 1976; Ferrer & Osacar 1994). We want however to look at the motion of the nodes, or equivalently the motion of the vector k2 in the moving frame (i1, j1, k1) of the orbit of the second body. In order to easily compute the equations of motion, we introduce two vectors associated to the orbit of the binary that are colinear to i1 and k1, and include in their norm the eccentricity of the binary, as in (Tremaine et al. 2009): K1 =q1 − e2 1k1 , I1 = e1i1 . (3.7) If we drop all the constant terms in equation 3.6 and α′G2 = − hHiM1,M2 where α′ = 3 4 n1„ a1 2 2I 2 1 + (K1.k2)2 − 5(I1.k2)2 , (3.8) a2«7/2 β1 M01 2)2r1 + 1 (1 − e2 m2 M01 (3.9) , and M01, n1 are defined as in section 2. 3.2 Equations of motion The components of K1, I1 and k2 have the following Poisson brackets3 (Borisov & Mamaev 2005; Bou´e & Laskar 2006): {K1i,K1j} = − {I1i,I1j} = − ǫijk Λ1 K1k , ǫijk Λ1 K1k , {k2i, k2j} = − {K1i, I1j} = − k2k , ǫijk G2 ǫijk Λ1 I1k , (3.10) (3.11) where Λ1 = β1√µ1a1 and ǫijk is the Levi-Civita symbol4. The equations of motion for the three vectors are thus: K1 = − I1 = − k2 = − 1 Λ1 1 Λ1 1 G2 (K1 ∧ ∇K1 H + I1 ∧ ∇I1 H) , (I1 ∧ ∇K1 H + K1 ∧ ∇I1 H) , k2 ∧ ∇k2 H . (3.12) (3.13) (3.14) In order to look at the motion of the vector k2 in the moving frame (i1, j1, k1) of the orbit of the second body, we use as Bou´e & Laskar (2006) the above system to derive equations for x = (k2.i1), y = (k2.j1), z = (k2.k1) and e1. Indeed, x = (k2.I1)/I1, z = (k2.K1)/K1, e1 = I1, and y is obtained using the identity x2 + y2 + z2 = 1: 3 We use the following convention, where pi are momenta and qi positions: {f, g} = Pi " ∂f 4 ǫijk = +1 if (i, j, k) is an even permutation of (1, 2, 3), ǫijk = −1 is the permutation is odd, and ǫijk = 0 in all other cases. ∂pi ". − ∂f ∂qi ∂g ∂qi ∂pi ∂g High inclination orbits in the secular quadrupolar three-body problem 7 Λ1q1 − e2 1y(2 − 5x2) (3.15) 1)(2 − 5x2) + 5e2 1z2] (3.16) 1)yz + α′ G2 1)xz x x = α′(1 − e2 y = −α′(1 + 4e2 − α′ G2 p1 − e2 1xy + α′ G2 Λ1 5e1q1 − e2 e1 = α′ G2 Λ1 z = 5α′e2 Λ1 1 [(1 − e2 5e2 1 xyz p1 − e2 1 1xy (3.17) (3.18) The equations for x, y and z contain two terms: the first one is identical to the outer restricted system, and the second one includes the motion of the reference frame (i1, j1, k1) induced by the interaction with the third body. Note that when Λ1 is very large compared to G2 so that we can assume that G2/Λ1 is equal to zero, which corresponds to the case where m2 ≪ m0 and m1, the above system is identical to the outer restricted system (2.16) -- (2.18). The conservation of the total angular momentum C = G1+G2 introduces the two main parameters of the problem. Indeed, Λ2 1(1 − e2 1) + G2 2 + 2Λ1q1 − e2 1G2z = C 2 . (3.19) We note γ = C/Λ1, γ2 = G2/Λ1. The above expression of the norm of the total angular momentum can be rewritten 1 as a function of as a second degree equation giving p1 − e2 z using the two parameters γ and γ2: (1 − e1)2 + 2γ2zq1 − e2 2 − γ2 = 0 . The Hamiltonian can then be rewritten as: 1 + γ2 (3.20) 1 2 α′Λ1γ2[z2 + e2 hHi = − The inequalities −1 6 z 6 1 and 0 6 e1 < 1 give the boundaries of the parameter space and the range of possible values of e1 for any given couple of parameters5 (γ, γ2): 1(2 − z2 − 5x2)] . (3.21) γ − γ2 6 1 , γ − γ2 6 q1 − e2 1 6 min[γ + γ2, 1] . (3.22) (3.23) With these notations, the outer restricted problem of section 2 corresponds to the limit where γ2 = 0, and in this case e1 = p1 − γ2 is constant as we saw. Note that when γ2 > γ, we have G2 > C, so this part of the parameter space contains only retrograde motions. Our aim in this paper is to show the continuity between the outer restricted case we studied in section 2, and the inner restricted case that was investigated by Kozai (1962) and Lidov (1962). Both these problems lie in the region of parameter space where γ > γ2 so we will restrict our study to this case6. 5 The left part of the second inequality is strict if γ = γ2. 6 The other half of the parameter space (γ 6 γ2) corresponds to retrograde motions which are of less physical interest and much more technical to study using our approach, in particular because equation (3.20) does not have a unique solution in this case. The interested reader will find a complete discussion of this case in (Lidov & Ziglin 1976; Ferrer & Osacar 1994). In our case where γ > γ2, there is only one acceptable root to equation (3.20), which is: q1 − e2 1 = −γ2z +q(γ2z)2 + γ2 − γ2 2 . This relation implies that e1 is a growing function of z. Note that z = cos i2, where i2 is the inclination of the outer body in the reference frame of the inner binary. As such, coplanar prograde motions (z = 1) will always occur for the maximal value of the eccentricity of the inner binary: (3.24) e1,max =p1 − (γ − γ2)2 . (3.25) Conversely, low eccentricities for the binary will be as- sociated to lower values of z, and thus higher inclinations. Relation (3.23) implies that the inner binary can only have a circular motion if γ + γ2 > 1. In this case, coplanar retro- grade motion (z = −1) is not allowed, and the lowest value of z is: z0 = cos i2,max = γ2 − γ2 2γ2 2 − 1 . (3.26) When γ + γ2 < 1 however, coplanar retrograde motion (z = −1) is possible and the associated value of the eccen- tricity of the binary is: e1,min =p1 − (γ + γ2)2 . (3.27) 3.3 Dynamical regimes In appendix A, we briefly derive in the framework of the present study the fixed points of the system and the bound- aries of the dynamical regimes in parameter space that are given in (Ferrer & Osacar 1994). The fixed points are named as follow: the north pole is called N , and the south pole S; linearly stable fixed points are named E, as elliptic, and lin- early unstable points are named H, as hyperbolic; finally, signs are placed as indices to refer to the symmetry of the problem with respect to the two planes x = 0 and y = 0. There are three dynamical regimes in the region of param- eter space we study. In region O of figure 4, the parameter γ2 = G2/Λ1 is small (less than 1/2). This is the case in particular when the mass ratio m2/m1 is small. Moreover, γ2 + 3γ2 2 < 1. The phase space is topologically equivalent to the outer restricted problem of section 2. The north pole, which corresponds to coplanar prograde motion with maximal eccentricity for the binary is linearly stable. There are two additional stable fixed points E± in the plane y = 0 (see section A4). They be- long to the same family as the fixed points y = z = 0, x = ±1 of the outer restricted problem that are responsible for the stable high inclination orbits observed by Verrier & Evans (2009). When γ + γ2 6 1, the south pole which corresponds to coplanar retrograde motion with minimal eccentricity for the binary, is also linearly stable. The a panels of figures 5 and 6 provide a visualisation of the topology of this case.7 When γ + γ2 > 1, the south pole is no longer accessible as stated in the previous section. It is however replaced by a 7 Note that the south pole in figure 5 a corresponds to the out-most trajectory; this is an artifact of the coordinate map (i2 cos Ω2, i2 sin Ω2) which sends the south pole of the sphere onto the circle i2 = 180◦. 8 F. Farago and J. Laskar Figure 4. Parameter Space. The dark gray areas are excluded by equation (3.22), the light gray area correspond to the part of parameter space corresponding to γ2 > γ which we do not study. The dotted line γ + γ2 = 1 separates the zone where there can be coplanar retrograde motion associated to a minimum eccentricity for the inner binary that is strictly higher than 0 (below the dot- ted line) and the zone where the inner binary can be circular but the inclination is bounded (see section A2). In zone O, the prob- lem is topologically equivalent to the outer restricted problem. In zones I and I' it is topologically equivalent to the inner restricted problem, with zone I being equivalent to situations above the crit- ical inclination and zone I' being equivalent to situations under the critical inclination. The letters a -- f correspond to the values of the parameters used to plot the corresponding panels in figures 5 and 6. stable trajectory at a maximal inclination given by equation (3.26), as can be seen on panel b of figures 5 and 6. This trajectory corresponds to a circular inner binary (see sec- tion A2). Finally, there are two unstable points H± in the x = 0 plane that belong to the same family as the unstable points of the outer restricted problem x = z = 0, y = ±1 (see section A3). Panels c of figures 5 and 6 show the limiting case be- tween regions O and I. On this boundary, γ2 + 3γ2 2 = 1. The two unstable points H± are now located on the boundary of the accessible part of the sphere. Regions I and I' of figure 4 are both in the part of the parameter space defined by γ2 + 3γ2 2 > 1. In this zone, the problem becomes topologically equivalent to the inner re- stricted problem studied by Lidov (1962) and Kozai (1962). In the inner restricted case, there is a critical value of the inclination (cos i2 =p3/5) under which a circular inner bi- nary is always linearly stable, and above which a circular inner binary is always linearly unstable, giving rise to Kozai cycles. In region I of figure 4, the dynamical regime is topolog- ically equivalent to the inner restricted problem in the case where the inclination is superior to the critical value. The limit trajectory z = z0 which corresponds to a circular inner Figure 5. Trajectories in the (i2 cos Ω2, i2 sin Ω2) plane for dif- ferent values of the parameters. See section 3.3 for a detailed discussion and appendix A for calculations. a: (γ, γ2) = (0.8, 0) outer restricted case with e1 = 0.6; b: (γ, γ2) = (0.8, 0.25); c: 2 = (1/3)(1 − γ2); d: (γ, γ2) = (0.8, 0.4); e: (γ, γ2) = γ = 0.8, γ2 (1.08, 0.4); f: (γ, γ2) = (1.28, 0.4). binary becomes linearly unstable. However, the north pole and the two fixed points E± are still stable. In the inner restricted phase space, when the inclination is superior to the critical value, there are two possible behaviours for the periastron of the inner particle: it can either circulate, or librate around ±90◦. In our representation, the circulation case corresponds to trajectories around the north pole, and the libration islands correspond to the two fixed points E±. This is shown in panels d and e in figures 5 and 6. In region I' of figure 4, the dynamical regime is topo- logically equivalent to the inner restricted problem in the case where the inclination is inferior to the critical value. Only one stable fixed point remains, on the north pole of the sphere, associated to prograde coplanar motion. This is shown in panel f in figures 5 and 6. In both regions I and I', the parameter γ2 = G2/Λ1 can take higher values. This is in particular true when the mass ratio m2/m1 increases. The curve between regions I and I' is linked to the crit- ical inclination that is defined in the inner restricted case. Indeed, along that curve, given by equation A10, we have High inclination orbits in the secular quadrupolar three-body problem 9 zq1 − e2 1 ≈r 3 5 . (3.34) Recall that z = cos i2, where i2 is the inclination of the outer orbit in the reference frame of the inner orbit. Thus, the inclination of the inner orbit relatively to the outer orbit is i1 = −i2, and the above equation becomes: cos i1q1 − e2 1 ≈r 3 5 . (3.35) This relation is precisely the one giving the critical value of the normal component of the angular momentum of the inner body in the inner restricted problem. CONCLUSION We first studied the case of a massless particle orbiting a binary at a long distance, and, in the secular and quadrupo- lar approximations, gave a full analytical description of the motion along with the expression of the period of the sec- ular motion. When the inner binary is circular, only nodal precession takes place. However, when the binary is elliptic, libration islands appear at high inclinations, and these is- lands grow bigger when the eccentricity of the binary rises. Verrier & Evans (2008, 2009) observe a similar nodal libra- tion in their study of the stability of particle populations in the quadruple stellar system HD 98800, and we showed that the analytical framework that we derived for the outer restricted problem is well suited to explain the results of Verrier and Evans. The quadrupolar secular three-body problem is still integrable when all the bodies have positive masses (Harrington 1969; Lidov & Ziglin 1976; Ferrer & Osacar 1994). Using a vectorial formalism as (Bou´e & Laskar 2006, 2009; Tremaine et al. 2009), we looked at this problem from the point of view of the outer restricted case. We showed how the outer restricted problem relates to the general case, and to the inner restricted case studied by Kozai (1962) and Lidov (1962): when the mass of the outer body is small enough compared to the mass of the inner body, the gen- eral case behaves similarly to the outer restricted problem. When the mass of the outer body increases enough, the gen- eral case behaves like the inner restricted problem. We gave an expression of the boundary between these two regimes. The outer restricted problem and its generalization to the non restricted case provide an interesting starting point in the study of circumbinary planetary systems, such as the one discovered recently around the eclipsing sdB+M sys- tem HW Virginis (Lee et al. 2009). In this system, the in- ner binary is very tight with a period of 2.8 hr, while the proposed planetary companions have periods of 9.1 yr and 15.8 yr, so the quadrupolar expansion is fully justified. An- other field of application of the outer restricted problem is the study of the motion of stars orbiting around binary black holes (Mikkola & Merritt 2008; Gillessen et al. 2009; Merritt et al. 2009). Figure 6. Trajectories on the unit angular momentum sphere for different values of the parameters. See section 3.3 for a detailed discussion and appendix A for calculations. a: (γ, γ2) = (0.8, 0) outer restricted case with e1 = 0.6; b: (γ, γ2) = (0.8, 0.25); c: γ = 0.8, γ2 2 = (1/3)(1 − γ2); d: (γ, γ2) = (0.8, 0.4); e: (γ, γ2) = (1.08, 0.4); f: (γ, γ2) = (1.28, 0.4). the following limits when γ → ∞: γ2 γ → 1 , γ2 − γ → −r 3 5 . (3.28) When G2 is very large compared to G1, we can make the following first order expansion: γ2 − γ = = ≈ 1 + 2(G2.G1) Λ1 G2 − C G2 −pG2 G2 − G2(1 + (G2.G1)/G2 2) (G2.G1) 2 + G2 Λ1 Λ1 (3.29) (3.30) (3.31) (3.32) (3.33) Λ1G2 ≈ − ≈ −zq1 − e2 1 . As such, we see that along the border between regions I and I', when γ and γ2 both tend to infinity, we have the relation 10 F. Farago and J. Laskar APPENDIX A: FIXED POINTS AND BIFURCATIONS The right hand sides of equations (A4) and (A5) vanish for a certain value of x equal to: The fixed points and boundaries presented in section 3.3 have already been studied by Lidov & Ziglin (1976) and Ferrer & Osacar (1994). We briefly present here their deriva- tion in the framework of the present formalism. With the notations of section 3, we will limit ourselves to γ > γ2. x2 0 = γ2 + 3γ2 10γ2 2 2 − 1 . (A8) The curve γ2 + 3γ2 2 = 1 separates in figure 4 the regions noted O and I. We can distinguish three cases: A1 Poles of the sphere, x = y = 0 case corresponds to case 1 in section 5 of This (Ferrer & Osacar 1994). Note that their sphere is con- structed using the eccentricity and perihelion of the inner binary, and is thus different from our angular momentum sphere. For all values of the parameters in the domain we study, the north pole z = 1, which corresponds to coplanar pro- grade motion, is a fixed point of the system. The associated eccentricity of the inner binary is: e1,max =p1 − (γ − γ2)2 . (A1) It is the maximal value of the eccentricity. This fixed point is always linearly stable. It is noted N in figures 5 and 6. Figure 5 shows the lines of equal energy in the (i2 cos Ω2, i2 sin Ω2) plane, and figure 6 shows these lines plotted on the sphere of unit angular momentum of the outer body k2 2 = 1. When γ + γ2 < 1 (under the dotted line in figure 4), the south pole z = −1 (noted S in the following figures), which corresponds to coplanar retrograde motion, is also a linearly stable fixed point of the system. The eccentricity of the inner binary is minimal and equal to: e1,min =p1 − (γ + γ2)2 . (A2) Note that in this region of parameter space, the inner binary cannot be circular. When γ + γ2 > 1 (above the dotted line in figure 4), the minimal eccentricity of the binary is 0 as deduced from (3.23). The south pole z = −1 does not correspond to a real value of the eccentricity in this case. This limit how- ever is not a bifurcation strictly speaking. When crossing it, the stable south pole of the sphere is replaced by a stable trajectory at maximal inclination. A2 Circular Trajectories for the inner binary In the region of parameter space where circular trajectories exist for the binary (above the dotted line in figure 4), the value of z which corresponds to such trajectories is minimal and equal to: z0 = γ2 − γ2 2γ2 2 − 1 . (A3) The equations of motion on the small circle of the sphere z = z0 are: α′y(z0 + γ2(2 − 5x2)) , x = y = −α′x(z0 + γ2(2 − 5x2)) , z = 0 , e1 = 0 . (A4) (A5) (A6) (A7) (i) γ2 + 3γ2 2 < 1. In region O of figure 4, x2 0 < 0 so there is no fixed point on the circle z = z0. As such, this circle is a trajectory for which the inner binary is circular and the outer orbit precesses at a fixed inclination given by i2,max = ArcCos z0. Moreover, this trajectory is linearly stable. (ii) γ2 + 3γ2 2 = 1. There are two fixed points on the circle z = z0 at the coordinates (x = 0, y = ±p1 − z2 (iii) γ2 + 3γ2 y2 0 = 1 − x2 0 − z2 yields: 2 > 1. In this case, we must also check that 0 > 0. The limit case where there is equality 0 ). 5γ4 2 − (4 + 10γ2)γ2 2 + (5γ4 − 8γ2 + 3) = 0 . (A9) This boundary limits the regions I and I' in figure 4. When solving the above equation for γ2 2 and selecting only the relevant solution satisfying γ > γ2 , γ +γ2 > 1, we obtain a solution that corresponds to equation 44 in section 5.2 of (Ferrer & Osacar 1994) and that can be written using our notations as: γ2 2 = 2 + 5γ2 −p60γ2 − 11 5 , γ + γ2 > 1 . (A10) In region I, y2 0 > 0 so there are four fixed points on the circle z = z0, at the coordinates (±x0,±y0). They are noted H±±, in panels d and e of figures 5 and 6. Moreover, the trajectories that correspond to circular binaries are unstable in this zone. In region I' however, y2 0 < 0 so we are again in a region of parameter space where there are no fixed points on the circle z = z0, and the trajectories associated to circular binaries are again stable. A3 The x = 0 plane When x = 0, the only non trivial equation remaining in system (3.15) -- (3.18) is x = 0. Looking for a fixed point different from x = y = 0, we have to solve zq1 − e2 1 + 2γ2 = 0 , which after using relation (3.20) yields: (A11) (A12) (A13) (A14) (A15) 2 , x = 0 , e1 = q1 − γ2 − 3γ2 y = ± pγ2 − γ2 pγ2 + 3γ2 z = − pγ2 + 3γ2 2γ2 , . 2 2 2 We thus have two symmetric fixed points in the plane x = 0. They are noted H± in figures 5 and 6. For these fixed points to exist, the associated eccentricity must be a real number. As such, their domain of existence is the re- gion noted O in figure 4. This is case 2.1 in section 5 of (Ferrer & Osacar 1994). High inclination orbits in the secular quadrupolar three-body problem 11 Mikkola S., Merritt D., 2008, AJ, 135, 2398 Palaci´an J. F., Yanguas P., 2006, Celest. Mech. Dyn. As- tron., 95, 81 Palaci´an J. F., Yanguas P., Fern´andez S., Nicotra M. A., 2006, Physica D Nonlinear Phenomena, 213, 15 Poincar´e H., 1905, Le¸cons de m´ecanique c´eleste profess´ees `a la Sorbonne Touma J. R., Tremaine S., Kazandjian M. V., 2009, MN- RAS, 394, 1085 Tremaine S., Touma J., Namouni F., 2009, AJ, 137, 3706 Verrier P. E., Evans N. W., 2008, MNRAS, 390, 1377 Verrier P. E., Evans N. W., 2009, MNRAS, 394, 1721 These two fixed points are linearly unstable in their do- main of existence. Note that in the outer restricted problem (γ2 = 0), these fixed points are simply x = z = 0, y = ±1. A4 The y = 0 plane When y = 0, the only non trivial equation we must solve is y = 0. Here again, we look for another fixed point than x = y = 0, thus we have to solve: (1 + 4e2 1)z + [(1− e2 1)(2− 5x2) + 5e2 1z2] = 0 . (A16) γ2 p1 − e2 1 1 in 1)2 + 5 8 5 γ2 2 + 1 2 2 )2 = 0 place of z using (3.20), we get: Substituting 1 − z2 in place of x2 and then p1 − e2 1)3 −„γ2 + (1 − e2 8« (1 − e2 (γ2 − γ2 (A17) This equation is the same as equation number 40 in (Ferrer & Osacar 1994). In our region of parameter space, there is at most one root which satisfies to the constraint (3.23). The curve separating the zone where there is one so- lution and the zone where there is no solution corresponds to the case where the limit value e1 = 0 is a solution, and co- incides with the boundary between regions I and I' in figure 4 which is given by equation (A10). When there is a solution, the value of e1 can be trans- lated into a value of z using (3.20). Since y = 0, we get two values of x = ±√1 − z2, and there are thus two symmetric fixed points on the sphere, which are both linearly stable. They are noted E± in figures 5 and 6. When γ2 = 0, these fixed points become simply y = z = 0, x = ±1, which are re- sponsible of the stable orbits at high inclination as discussed in the previous sections. REFERENCES Borisov A. V., Mamaev I. S., 2005, Dynamics of the Rigid Body (in Russian). R&C Dynamics, Moscow, (http://ics.org.ru/) Bou´e G., Laskar J., 2006, Icarus, 185, 312 Bou´e G., Laskar J., 2009, Icarus, 201, 750 Ferrer S., Osacar C., 1994, Celest. Mech. Dyn. Astron., 58, 245 Gillessen S., Eisenhauer F., Trippe S., Alexander T., Genzel R., Martins F., Ott T., 2009, ApJ, 692, 1075 Harrington R. S., 1969, Celest. Mech., 1, 200 Kinoshita H., Nakai H., 2007, Celest. Mech. Dyn. Astron., 98, 67 Kozai Y., 1962, AJ, 67, 591 Laskar J., 1989, in Benest D., Froeschle C., eds, Modern Methods in Celestial Mechanics, Syst`emes de Variables et El´ements. Editions Fronti`eres, pp 63 -- 87 Lee J. W., Kim S.-L., Kim C.-H., Koch R. H., Lee C.-U., Kim H.-I., Park J.-H., 2009, AJ, 137, 3181 Lidov M. L., 1962, P&SS, 9, 719 Lidov M. L., Ziglin S. L., 1976, Celest. Mech., 13, 471 Malige F., Robutel P., Laskar J., 2002, Celest. Mech. Dyn. Astron., 84, 283 Merritt D., Gualandris A., Mikkola S., 2009, ApJ Lett., 693, L35
1804.01532
1
1804
2018-04-04T18:00:01
Planetary population synthesis
[ "astro-ph.EP" ]
In stellar astrophysics, the technique of population synthesis has been successfully used for several decades. For planets, it is in contrast still a young method which only became important in recent years because of the rapid increase of the number of known extrasolar planets, and the associated growth of statistical observational constraints. With planetary population synthesis, the theory of planet formation and evolution can be put to the test against these constraints. In this review of planetary population synthesis, we first briefly list key observational constraints. Then, the work flow in the method and its two main components are presented, namely global end-to-end models that predict planetary system properties directly from protoplanetary disk properties and probability distributions for these initial conditions. An overview of various population synthesis models in the literature is given. The sub-models for the physical processes considered in global models are described: the evolution of the protoplanetary disk, the planets' accretion of solids and gas, orbital migration, and N-body interactions among concurrently growing protoplanets. Next, typical population synthesis results are illustrated in the form of new syntheses obtained with the latest generation of the Bern model. Planetary formation tracks, the distribution of planets in the mass-distance and radius-distance plane, the planetary mass function, and the distributions of planetary radii, semimajor axes, and luminosities are shown, linked to underlying physical processes, and compared with their observational counterparts. We finish by highlighting the most important predictions made by population synthesis models and discuss the lessons learned from these predictions - both those later observationally confirmed and those rejected.
astro-ph.EP
astro-ph
Planetary population synthesis Christoph Mordasini Abstract In stellar astrophysics, the technique of population synthesis has been suc- cessfully used for several decades. For planets, it is in contrast still a young method which only became important in recent years because of the rapid increase of the number of known extrasolar planets, and the associated growth of statistical obser- vational constraints. With planetary population synthesis, the theory of planet for- mation and evolution can be put to the test against these constraints. In this review of planetary population synthesis, we first briefly list key observational constraints. Then, the work flow in the method and its two main components are presented, namely global end-to-end models that predict planetary system properties directly from protoplanetary disk properties and probability distributions for these initial conditions. An overview of various population synthesis models in the literature is given. The sub-models for the physical processes considered in global models are described: the evolution of the protoplanetary disk, the planets' accretion of solids and gas, orbital migration, and N-body interactions among concurrently growing protoplanets. Next, typical population synthesis results are illustrated in the form of new syntheses obtained with the latest generation of the Bern model. Planetary for- mation tracks, the distribution of planets in the mass-distance and radius-distance plane, the planetary mass function, and the distributions of planetary radii, semi- major axes, and luminosities are shown, linked to underlying physical processes, and compared with their observational counterparts. We finish by highlighting the most important predictions made by population synthesis models and discuss the lessons learned from these predictions - both those later observationally confirmed and those rejected. Christoph Mordasini Physikalisches Institut, University of Bern, Gesellschaftsstrasse 6 CH 3012 Bern, Switzerland. e- mail: [email protected] 1 2 Confronting theory and observation Christoph Mordasini Since the discovery of the first extrasolar planet around a solar-like star by Mayor & Queloz (1995), it has become clear from observations that the population of extraso- lar planets is characterized by extreme diversity. This diversity in terms of planetary masses, orbital distances, system architectures, internal compositions etc. was not anticipated by earlier theoretical models of planet formation (e.g., Boss 1995) that were based on just one planetary system, our own solar system, indicating important shortcomings in the theory. Since 1995, the number of known extrasolar planets has increased rapidly, reach- ing now several thousand (Schneider et al. 2011; Wright et al. 2011). This allows one to study the extrasolar planets as a statistical population instead of single objects only, even though the study of a benchmark individual planetary systems (including the solar system) continues to be key to understand planet formation as well. This planetary population is characterized by a number of statistical distribution (e.g., of the mass or eccentricity), dependencies on host star properties (like the stellar metallicity), and correlations between these quantities. These statistical constraints provide a rich data set with which the theoretical predictions of population synthesis models can be confronted. The basic idea behind the planetary population synthesis method is that the ob- served diversity of extrasolar planets is due to a diversity in the initial conditions, the protoplanetary disks (e.g., Andrews et al. 2010). While it is typically difficult to observe the process of planet formation directly (except for a handful special cases, e.g., Sallum et al. 2015), in a numerical model the link between disk and plane- tary system properties can be established with so-called global model. This class of models directly predicts the final (potentially observable) properties of synthetic extrasolar planets based on the properties of their parent synthetic protoplanetary disk. For this, such global end-to-end models build on simplified results of many different detailed models for individual physical processes of planet formation, like accretion and migration. In global models, these individual processes are linked together, which is a source of considerable complexity, even for relatively simple sub-models. Thanks to this approach, the population-wide, statistical consequences of an individual physical description (like orbital migration, e.g., Masset & Casoli 2010) become clear and can be statistically compared with the observed popula- tion (e.g., the semimajor axis distribution or the frequency of mean motion reso- nances). This means that first, theoretical models of a specific process can be put to the observational test which is otherwise often difficult as we can only observe the combined effect of all acting processes, and second that the full wealth of obser- vational data (the entire statistical information coming from different observational techniques like radial velocities, transits, direct imaging, microlensing, ...) can be used to constrain theoretical planet formation models. This also avoids that models are constructed that can only describe specific types of systems, but fail for many others, as illustrated for the case of the solar system mentioned above. As global models in population syntheses are in the end nothing else than cou- pled agglomerates of other specialized models, it is clear that the predictions of Planetary population synthesis 3 population syntheses directly reflect the state of the field of planet formation theory as a whole, which is exactly their purpose. This means that as our understanding of planet formation changes, so do the population synthesis models. In planet formation, fundamental physical processes governing planet forma- tion are currently still uncertain. An important example for this are the processes that drive accretion in protoplanetary disk (classical MRI-driven viscous accretion, MHD winds, e.g., Bai 2016), which has via different disk structures strong con- sequences for planet formation (Ogihara et al. 2015). Another important example is the relative importance of solid building blocks of various sizes ranging from cm-sized pebbles to classical 100 km-sized planetesimals (Ormel 2017). It could appear that given such large uncertainties, currently no meaningful global models can therefore be constructed - but actually, the argument must be turned around: it is with population syntheses that these different theories can be put to the observa- tional test to identify which ones lead to synthetic populations that agree or disagree with observations, and to improve in this way the understanding of how planets form, which is the final goal. This chapter is organized as follows: we first discuss the most important obser- vational constraints, then describe the method of population synthesis including a short overview of the input physics currently considered in global models. We then turn to the discussion of the most important results including the comparison with observations. We conclude the chapter with the discussion of tests of specific sub- models and predictions for future instruments and surveys. For further information on the method, the reader may consider the reviews of Benz et al. (2014) and Mordasini et al. (2015). Two other relevant publications for the (initial) development of the method are Ida & Lin (2004a) and Mordasini et al. (2009a). Statistical observational constraints The number and type of observational statistical constraints available for compar- isons is in principle very large and multifaceted (for reviews, see Udry & Santos 2007; Winn & Fabrycky 2015). However, there are a number of key constraints the comparison to which population syntheses have traditionally focussed on. These key constraints are usually the results of large observational surveys, both from the ground and space. Important surveys and publications analyzing them are, e.g., the HARPS high precision radial velocity survey (Mayor et al. 2011), the Keck & Lick radial velocity survey (Howard et al. 2010), the CoRoT (Moutou et al. 2013) and Kepler transit surveys (Coughlin et al. 2016), the various direct imaging (Bowler 2016), or the microlensing surveys (Cassan et al. 2012). The high importance of surveys stems from the fact that they have a well known observational bias. This makes it possible to correct for it and to infer the underlying actual distributions that are predicted by the theoretical models. 4 Christoph Mordasini All these different techniques put constraints on different aspects of the global models. Especially when they are combined, they are highly constraining even for global models that often have a significant number of free parameters as the com- bined data carries so much constraining information. The constraints can be grouped into three classes: the frequency of different planet types, the distribution functions of planetary properties, and correlations with stellar properties. We next give a short overview of these observational constraints. et al. 2010; Mayor et al. 2011). Frequencies of planet types • The frequency of hot Jupiters around solar-like stars is about 0.5-1% (Howard • The frequency of giant planets within 5-10 AU is 10-20% for FGK stars (Cum- ming et al. 2008; Mayor et al. 2011). The giant planets have a multiplicity rate of about 50% (Bryan et al. 2016). • There is a high frequency (20-50%) of close-in (fractions of an AU) low-mass (a few Earth masses) respectively small (R (cid:46) 4R⊕) super-Earth and sub-Neptunian planets from high-precision radial velocity (Mayor et al. 2011) and the Kepler survey (Fressin et al. 2013; Petigura et al. 2013). These planets are often found in tightly packed multiple systems. Planetary systems clearly different from the solar system are thus very frequent. • There is a low frequency on the 1% level of detectable (i.e., sufficiently luminous) massive giant planets at distances of tens to hundreds of AU. This means that the frequency of giant planets must somewhere drop with orbital distance by about a factor ten. The occurrence rate is likely positively correlated with the stellar mass Bowler (2016). • There is a high frequency of cold, roughly Neptunian-mass planets around M dwarfs as found by microlensing surveys (Cassan et al. 2012). • There is a very high total fraction of stars with detectable planets of ∼75% as indicated by high-precision radial velocity searches with a ∼ 1 m/s precision (Mayor et al. 2011). At least in the solar neighborhood, stars with planets are thus the rule. Distributions of planetary properties • One of the most important diagrams is the observed two-dimensional distribution of planets in the mass-distance (or radius-distance) plane, revealing a number of pile-ups and deserts (see Fig. 1). For the comparison with the synthetic popula- tions, it is of paramount importance to keep in mind that the observed diagram gives a highly distorted impression of the actual population because of the de- tection biases of the different techniques. Hot Jupiters, for example, appear to be Planetary population synthesis 5 Fig. 1: Mass-distance diagram of confirmed planets in "The Extrasolar Planets En- cyclopedia" (Schneider et al. 2011) as of 2017. Red, cyan, magenta, and green points indicate planets detected by the radial velocity, transits, direct imaging, and mi- crolensing technique, respectively. The planets of the solar system are also shown for comparison. frequent in this plot. But the plot still illustrates the enormous diversity in the outcome of the planet formation process. At the same time, it also indicates that there is some structure. • The mass function is approximately flat in log space in the giant planet regime (Marcy et al. 2005) for masses between 30 M⊕ and about 4 M(cid:88) (where 1 M(cid:88) is the mass of Jupiter). At even higher masses, there is a drop in frequency (Santos et al. 2017). The upper end of the planetary mass function is poorly known, but might lie around 30 M(cid:88) (Sahlmann et al. 2011). Towards the lower masses, at around 30 M⊕, there is a break in the mass function and a strong increase of the frequency towards smaller masses (Mayor et al. 2011). The mass function below a few Earth masses is currently unknown. • The semimajor axis distribution of giant planets consists of a local maximum at a period around 4 days caused by the hot Jupiters, a less populated region further out (the period valley) and finally an upturn at around 1 AU (Udry et al. 2003). The frequency seems to be decreasing beyond 3-10 AU (Bryan et al. 2016). • The eccentricity distribution is, in contrast to the solar system with its very low eccentricities, broad, including some planets with eccentricities that exceed 0.9. 6 Christoph Mordasini The upper part of the distribution follows approximately a Rayleigh distribu- tion, as expected from gravitational planet-planet interactions (Juri´c & Tremaine 2008), indicating together with several other points that in some systems strong dynamical interactions occurred (see the discussion in Winn & Fabrycky 2015). A significant fraction of orbits are however also consistent with being circular. Eccentricities of lower mass planets ((cid:46) 30M⊕) are usually restricted to lower values ≤ 0.5 (Mayor et al. 2011). • The radius distribution of confirmed (Kepler) planets has a local maximum at around 1 Jovian radius as expected from the theoretical mass-radius relation (Mordasini et al. 2012b), followed by a distribution that is approximately flat in log(R) at intermediate radii of 4-10 R⊕. Below this radius, there is strong in- crease in frequency (Fressin et al. 2013; Petigura et al. 2013). At about 1.7 R⊕, there is a local minimum in the radius histogram (Fulton et al. 2017) separat- ing super-Earths from sub-Neptunes. This could be due to atmospheric escape of primordial H/He envelopes (Owen & Wu 2017; Jin & Mordasini 2018). Correlations with stellar properties • The best known correlation of planetary and stellar properties is the increase of the frequency of giant planets with host star metallicity (Gonzalez 1997; Santos et al. 2004; Fischer & Valenti 2005). In the super-solar metallicity domain, the frequency of giant planet increases approximately by a factor ten when going from [Fe/H]=0 to [Fe/H]=0.5. This is often taken as indication that core accretion is the dominant mode of giant planet formation (Ida & Lin 2004b; Mordasini et al. 2012a). The frequency of low-mass planets is in contrast independent of metallicity (Mayor et al. 2011). • The frequency of giant planets is lower for lower mass stars and around 2% for M-dwarfs (Bonfils et al. 2013). For stellar masses higher than 1 M(cid:12), the frequency first increases to reach a maximum at around 2 M(cid:12), followed by a rapid drop for M∗ (cid:38) 2.7M(cid:12) (Reffert et al. 2015). • Statistical correlations with stellar age are not yet well explored, but a number of detections of close-in planets around T-Tauri and young PMS stars have occurred (Mann et al. 2016; David et al. 2016; Donati et al. 2016; Yu et al. 2017). They show that close-in massive planets already exist after a few Myr, likely indicat- ing orbital migration via planet-disk interactions. Hot Jupiters might be more frequent around T Tauri stars than main sequence stars (Yu et al. 2017). At large orbital distances, direct imaging also probes young planets with ages of a few 10 Myr. The PLATO survey will put statistical constraints on the temporal evolution of the population of transiting planets, adding a new temporal dimension to the constraints. Planetary population synthesis Population synthesis method 7 In this section, we review the general workflow in the method, the past development of population syntheses models, the physical processes considered in global for- mation and evolution models, and finally the probability distributions of the initial conditions. Workflow of the population synthesis method The general workflow of the planetary population synthesis method is shown in Fig. 2. There are three main elements: first, and most importantly, the global model that predicts planetary properties directly based on disk properties. Second, the Monte Carlo distributions for the initial conditions of the global models that are derived from disk observations, from reconstructions of the disk properties in an equivalent way as done for the minimum mass solar nebula, or from theoretical arguments. Third, tools to apply the observational detection bias and to conduct the statistical comparison with the observed population. In general, this comparison will reveal differences between the theory and observations, which are then tracked back to assumptions about the governing physical processes as implemented in the model as well as the setting of model parameters. In case that the synthetic population matches the observations at least regarding a certain aspect, the synthetic population can also be used to make predictions about aspects that cannot yet be observed, including the expected yield of future surveys. Overview of population synthesis models in the literature In other fields of astrophysics, (stellar) population synthesis is a well-established technique for several decades (e.g., Bruzual & Charlot 2003), while for planets, it is still a recent approach. The construction of planetary population synthesis models was triggered by the rapidly increasing number of known extrasolar planets. In this section, we review past and present developments of such models by various groups. Early models were all based on the classical core accretion paradigm where the solids are accreted in the form of planetesimals (Perri & Cameron 1974; Mizuno 1980; Bodenheimer & Pollack 1986; Pollack et al. 1996). More recently, models were also based on core accretion with pebbles (Ormel & Klahr 2010; Johansen & Lambrechts 2017), and on planet formation via gravitational instability (Kuiper 1951; Cameron 1978; Boss 1997). 1. The Ida & Lin models. The pioneering population synthesis calculation of Ida & Lin (2004a) contained for the first time all the basic elements of population synthesis shown in Fig. 2, namely a purpose-built - and therefore fast - global 8 Christoph Mordasini Fig. 2: Elements and work flow of a planetary population synthesis framework (up- dated from Mordasini et al. 2015). planet formation model based on the core accretion paradigm, and a variation of the initial conditions in a Monte Carlo way. The effects of planetesimal accretion, parameterized gas accretion, and Type II orbital migration in simple power-law disks were considered. As most first-generation population synthesis models, the one-embryo-per-disk approximation was used. Later works added Type I migra- tion (Ida & Lin 2008a), a density enhancement due to a dead zone at the iceline (Ida & Lin 2008b), and finally a semi-analytical statistical treatment of the dy- namical interactions of several concurrently growing protoplanets (Ida & Lin 2010; Ida et al. 2013). 2. The Bern Model. Building on the Alibert, Mordasini & Benz (2004; 2005) model for giant planet formation in the solar system, Mordasini et al. (2009a,b) pre- sented population syntheses that included quantitative statistical comparisons with observations. Compared to the Ida & Lin models, the Bern model explic- itly solves the (partial) differential for the structure and evolution of the proto- planetary disk and the planets' interior structure, rather then using power-law solutions. This has the implication of substantially higher computational costs. Subsequent improvements addressed the structure of the protoplanetary disk (Fouchet et al. 2012), the solid accretion rate (Fortier et al. 2013), and the type I migration description (Dittkrist et al. 2014). The model was extended to include the planets' post-formation thermodynamic evolution (cooling and contraction) over Gyrs timescales (Mordasini et al. 2012c), as well as atmospheric escape (Jin et al. 2014). This makes it possible to predict directly also radii and luminosities instead of masses only. Also these models originally used the one-embryo-per- evolutionarymodelofMordasinietal.(2012c)is,however,stillsignificantlysimplifiedinseveralaspectsasdiscussedinsubsection'Atmosphereoftheplanet'Inviewoffutureobservationsyieldingverypreciseradii(e.g.bythephoto-metricCHaracterizingExOPlanetSatelliteCHEOPS,Broegetal.2013)itwillprobablybenecessarytofindmoreaccuratephysicaldescriptionsalsoinglobalmodels.Themass–radiusdiagramrepresents,inaprototypicalway,thetransitionofthefocusfrompureexoplanetdetectiontobeginningexoplanetcharacterizationinthepastfewyears.BesidestheM–Rrelationship,therewasrecentobservationalprogresstowardscharacterizationintwootherdomains.DirectimagingThefirsttechniquebesidestransitsthathasrecentlyyieldedimportantnewresultsforplanetcharacterizationisthedirectimagingtechnique.Themethodistechnicallychallengingduetothesmallangularseparationofaveryfaintsource(theplanet)fromamuchbrighterone(thehoststar).Thenumberofplanetsdetectedbydirectimagingiscurrentlystilllow.Butalreadythesediscoveries,liketheplanetsaroundHR8799(Maroisetal.2008)orβPictoris(Lagrangeetal.2010)havetriggerednumeroustheoreticalstudiesregardingtheirforma-tion(e.g.Dodson-Robinsonetal.2009;Kratteretal.2010).Twopointsabouttheseplanetsareinteresting:theirlargesemi-majoraxisandthefactthatwedirectlymeasuretheintrinsicluminosityatyoungagesinseveralinfrared(IR)bands.Bothquantitiesareimportanttounderstandtheformationmech-anism(coreaccretionorgravitationalinstability)andinparticularthephysicsoftheaccretionshockoccurringwhentheaccretinggashitstheplanet'ssurfaceduringformation(e.g.Commerçonetal.2011).Ifthegravitationalpotentialenergyoftheaccretinggasisradiatedaway,lowentropygasisincorporatedintotheplanet,leadingtoafaintluminosityandsmallradius(so-called'coldstart',Marleyetal.2007)whiletheaccretionofhighentropymaterialleadstoa'hotstart'withahighluminosityandlargeradius(e.g.Burrowsetal.1997;Baraffeetal.2003).Recently,Spiegel&Burrows(2012)haveshownthatthedifferentscenariosresultinobservabledifferenceinthemagnitudesoftheyoungplanets.Theglobalmodelmainlydiscussedinthiswork(seeFig.3)calculatestheluminosityduringboththeformationandevolutionphasewithaself-consistentcoupling.Foryounggiantplanets,thisisasignificantdifferencecomparedtopurelyevolutionarymodels,sincethiscouplingisnecessarytoknowtheentropyintheenvelopedirectlyafterformationandtocorrectlypredicttheluminosityatyoungages.Sincemulti-bandphotometrycanbeusedtoestimatethemetalenrichmentofaplanetandbecausenewdirectimaginginstrumentsarecurrentlybecom-ingoperational(SPHEREandGPI),itisimportantthatfutureglobalmodelswillincludebetterdescriptionsofthegasaccretionshockandbetteratmosphericmodels(cf.subsection'Atmosphereoftheplanet').SpectroscopySecond,oneofthemostimportantaspectsoftherecentobservationalprogresstowardscharacterizationarethespectraofanumberofexoplanetstransitingbrightstars(e.g.Richardsonetal.2007).TheatmosphererepresentsawindowintothecompositionofaplanetandcontainsamultitudeofFig.3.Schematicrepresentationoftheworkflowinthepopulationsynthesismethod(Ida&Lin2004a;Mordasinietal.2009a).The11computationalsub-modelsofthecombinedglobalplanetformationandevolutionmodelarebasedonthecoreaccretionparadigm(seeAlibertetal.2005a,2013,Mordasinietal.2012b,c).Globalmodelsofplanetformationandevolution5Tides Planetary population synthesis 9 disk approximation. The concurrent formation of multiple protoplanets interact- ing via an explicit N-body integrator was added in Alibert et al. (2013). 3. The models of Hasegawa & Pudritz (2011, 2012); Hasegawa & Pudritz (2013) combine a planet formation model based initially on the Ida& Lin models with power-law disks with inhomogeneities or the analytical disk model of Chambers (2009) and Cridland et al. (2016). These models emphasize the importance of "planet traps", i.e., special locations in the disk where orbital migration is slowed down or stopped due to transitions in the disk. These transitions are the edge of the MRI-dead zone, icelines, and the transition from the viscously heated to the irradiation-dominated region in the disk. Later updates (Alessi et al. 2017; Cridland et al. 2017) include models for the dust physics, astrochemistry, and radiative transfer. 4. While not used in population syntheses but in parameter studies, the models of Hellary & Nelson (2012); Coleman & Nelson (2014); Coleman & Nelson (2016) are global models that combine an N-body integrator with a 1D model for the disk's structure and evolution and the planets' orbital migration. In contrast to other models, the planetesimals are directly included in the N-body as massless test particles, and not simply represented as a surface density. Early models use fits to the results of Movshovitz et al. (2010) for the planets' gas accretion rate while later models (Coleman et al. 2017) calculate it by solving 1D structure equations. Similar global models were also presented by Thommes et al. (2008). 5. Based on the global model of Bitsch et al. (2015b), Ndugu et al. (2018) presented population syntheses based on the core accretion paradigm where the cores grow by the accretion of pebbles instead of planetesimals. Ndugu et al. (2018) focussed on the effect of the stellar cluster environment. The gas disk structure is obtained from 2D simulations including viscous heating and stellar irradiation assuming a radially constant mass flux (Bitsch et al. 2015a). The planets' gas accretion rate is given by analytical results of Piso & Youdin (2014). The cores grow by the accretion of mm-cm sized drifting pebbles (Lambrechts & Johansen 2012; Lam- brechts & Johansen 2014). The model uses the one-embryo-per-disk approach, such that N-body interactions are neglected, while type I and II migration are included. 6. An increasing number of population synthesis calculations are also based on variants of the gravitational instability model for giant planet formation (e.g., Forgan & Rice 2013; Nayakshin & Fletcher 2015; Muller et al. 2018). Similar to the core accretion models, these global models couple simple semi-analytic sub- models of disc evolution, disk fragmentation, initial embryo mass, gas accretion and loss (for example by tidal downsizing, Nayakshin 2010), orbital migration, grain growth, formation of solid cores by sedimentation, and recently, the N-body interaction of several fragments (Forgan et al. 2018). In the remainder of the article we concentrate on the population synthesis models based on the core accretion paradigm. 10 Global models: simplified but linked Christoph Mordasini The core accretion paradigm states that giant planets form in a two-step process. First a so-called critical core is built (with a mass of about 10 M⊕), which then trig- gers the accretion of the gaseous envelope. This happens in evolving disks of gas and solids in which also other protoplanets grow, leading to dynamical interactions. The gas disk and the protoplanets exchange angular moment, so that orbital migra- tion occurs. All these processes occur on similar timescales, meaning that they need to be considered in a self-consistently coupled fashion. A global planet formation model must thus consider this minimal set of physical processes (Benz et al. 2014): 1. The structure and evolution of the protoplanetary gas disk 2. The structure and evolution of the disk of solids (dust, pebbles, planetesimals) 3. The accretion of solids leading to the growth of the planetary solid core 4. The accretion of H/He leading to the growth of the planetary gaseous envelope 5. Orbital migration resulting from the exchange of angular momentum 6. N-body interaction among (proto)planets Additional sub-models may describe the internal structure of the core and envelope, the structure of the planetary atmosphere, the interaction of infalling planetesimals and pebbles with the protoplanets' envelope, the evolution of the star, or the loss of the gaseous envelope during the evolutionary phase, for example via atmospheric escape (e.g., Jin et al. 2014). Low-dimensional approximation For a statistical approach like population synthesis where hundreds of planetary sys- tems must be simulated over timescales of many millions of years during the forma- tion epoch, and even for billions of years during the evolution phase, it is currently not possible to use detailed multi-dimensional hydrodynamic simulations possibly even including radiative transfer and magnetic fields because of computational time limitations. Instead, the sub-models are either parameterized based on the results of detailed models or solve differential equations describing low-dimensional approximations like 1D spherically symmetric hydrostatic planet interior equations or 1D axisym- metric protoplanetary disk evolution equations. A key challenge of population syn- thesis is thus to "distill" the insights from 2D or 3D detailed models of one specific process (like orbital migration or pebble accretion) into simpler computationally efficient approximations, that however still correctly capture the essence of the gov- erning physical mechanism. Exploring which approximations are possible without losing the essence is an ongoing challenge for population synthesis models. On the other hand, the fact that the different processes are considered in a self- consistent coupled fashion over a long timescale is a strong aspect of global models, Planetary population synthesis 11 as it captures the non-linear interactions of the different processes as they are oc- curring also in nature. This coupling is a source of considerable complexity of the models, even for relatively simple individual sub-models. We next briefly discuss some of these sub-models. More detailed descriptions can be found in Benz et al. (2014) and Mordasini et al. (2017). Disk models The disk model describes the evolution of the surface density of gas and solids. It also gives the gas temperature, pressure, and vertical scale height. These quantities and their radial derivatives enter into the other sub-models in multiple ways, making the disk model a key component of a global model. The gas disk properties for ex- ample determine as outer boundary conditions the gas accretion rate of protoplanets, enter into the migration rates, control the aerodynamic behavior of small particles, or the damping of the random velocities of the planetesimals. The simplest way of setting up a parameterized (gas) disk model is a power law approach inspired by the minimum mass solar nebula MMSN (Weidenschilling 1977; Hayashi 1981), as used in the original models of Ida & Lin (2004a). In the MMSN approach, the present-day positions, masses, and compositions of the solar system planets are used to reconstruct the radial distribution of matter in the solar nebula, assuming in situ growth. In such models, the (initial) surface density of gas Σ as a function of distance from the star r is given as (cid:18) (cid:19)−3/2 Σ(r) = Σ0 r 1AU . (1) In a population synthesis, the normalization constant Σ0 is varied as a Monte Carlo variable to represent disk of different masses (see Sect. Probability distribution of disk initial conditions) with Σ0 ≈ 2400 g/cm2 corresponding for example to the sur- face density in the MMSN. In such simple models, the temperature T is also given as a power law. (Ida & Lin 2004a) for example assumed an optically thin disk, and a (cid:33) main sequence-scaling of the stellar luminosity as L ∝ M4 (cid:63) with stellar mass, so that (cid:19)−1/2(cid:32) M(cid:63) (cid:18) (2) T(r) = 280K r 1AU . M(cid:12) This however neglects (a) that disks are optically thick (both radially and vertically) with opacity transitions at condensation fronts which can act as migration traps, (b) the effect of viscous heating, and (c) it does not include any temporal evolution, including the fact that stars are not yet on the main sequence during the presence of the gas disk. A certain improvement over such simple MMSN-like models are ana- lytical disk models that take these effects into account, as for example the Chambers (2009) disk model that distinguishes between an inner viscously heated part and an outer irradiation-dominated part. 12 Christoph Mordasini A more complex, but still 1D approach is to solve the classical viscous evolution equation (Lust 1952; Lynden-Bell & Pringle 1974) for the surface density of the gas as a function of time t and distance from the star r (cid:34) (cid:16) (cid:17)(cid:35) ∂Σ ∂t = 1 r ∂ ∂r 3r1/2 ∂ ∂r r1/2νΣ − Σphot(r)− Σplanet(r). (3) with a viscosity ν that is written in the α-parameterization as ν = αcsH with cs the sound speed and H the vertical scale height (Shakura & Sunyaev 1973). Besides the viscous evolution term, the effects of mass loss by photoevaporation (e.g., Alexander et al. 2014) represented by Σphot(r) and of gas accretion by the planets giving raise to the Σplanet(r) term are also to be included. As an initial condition for this equation, the gas surface density is assumed to consist of a decrease close to the star due to the stellar magnetospheric cavity, a power-law in the main part, and an exponential decrease outside of a characteristic radius, as found in the analytical solution to the viscous accretion disk problem of Lynden-Bell & Pringle (1974). The initial gas surface density is then (cid:19)pg (cid:18) r 1AU (cid:32) r − Rout (cid:33)2+pg(cid:32) 1− (cid:114) r (cid:33) Rin Σg(t = 0,r) = Σ0 exp . (4) In this equation, Rout is the "characteristic" (outer) disk radius, Rin the inner ra- dius, and pg the power law exponent. Observations indicate pg ≈ −1 (Andrews et al. 2010). The four parameters in this equation may be treated as Monte Carlo random variables in a population synthesis. Under the assumption that dust is converted early in the disk's evolution every- where with full efficiency into planetesimals, the initial surface density of planetes- imals Σp would be given as (Mordasini et al. 2009a) Σp(t = 0,r) = fdgηiceΣg(t = 0,r) (5) where fdg is the dust-to-gas ratio (≈ the heavy element mass fraction Z), which is about 0.0149 in the Sun (Lodders 2003). It is another Monte Carlo variable, rep- resenting the different metallicities of stars (see Fig. 4). Finally, ηice reflects the reduction of the solid surface density at iceline(s). However, observations (e.g., Pani´c et al. 2009) and theoretical results (e.g., Birn- stiel et al. 2012) indicate that a significant radial redistribution of solids in the form of pebbles occurs. This may lead to more concentrated and steeper distributions of the solids (Kornet et al. 2001; Birnstiel & Andrews 2014) than predicted by Eq 5. In pebble-based models, the pebble surface density is calculated from the radial flux of pebbles, which is in turn controlled by the production rate of pebbles from dust at the pebble production line (Bitsch et al. 2015b). The temporal evolution of the gas disk is found by solving the aforementioned equation describing a viscous accretion disk including photoevaporation. In param- eterized models like in Ida & Lin (2004a), one uses instead an equation of the form Planetary population synthesis Σg(r) = − Σg(r) τdisk + Σphot. 13 (6) The first term on the right hand side leads to an exponential self-similar decay, while the second mimics the effects of photoevaporation. The characteristic disk timescale τdisk can again be treated as a Monte Carlo variable (Sect. Probability distribution of disk initial conditions). The surface density of solids decreases within the planet's feeding zone accord- ing to the amount of mass that the planet accretes, assuming that the surface density is uniform within the feeding zone (Thommes et al. 2003), i.e., Σp = − (3M∗)1/3 pBLM1/3 p 6πa2 Mc (7) where BL is the width of the feeding zone in Hill spheres, Mp the planet's mass, ap its semimajor axis, and Mc the planet's planetesimals accretion rate. Accretion of solids In planetesimal based models, the growth of the solid core with mass Mc is assumed to occur in the classical picture via the accretion of small background planetesimals. For this, a version of the Safronov (Safronov 1969) equation which gives the core accretion rate Mc is used: Mc = ΩΣpR2 captureFG (8) where Ω is the Keplerian frequency, Σp the mean surface density of planetesimals in the planet's feeding zone, Rcapture the capture radius which is in general larger than the core radius because of gas drag (Podolak et al. 1988; Mordasini et al. 2006), and FG is the gravitational focussing factor (Nakazawa et al. 1989; Greenzweig & Lis- sauer 1992). It depends among other quantities on the random velocities of the plan- etesimals vpls and would be given in the (idealized) two-body case as 1 +(vesc/vpls)2, where vesc is the escape velocity from the protoplanet (Safronov 1969). The ran- dom velocities of smaller planetesimals are more strongly damped by nebular gas drag leading to a higher focussing factor in Eq. 8. Furthermore, the drag-enhanced capture radii of the protoplanets in Eq. 8 is increased as well for smaller bodies, approaching very large radii for small particles as exemplified by pebble accretion. An insight into the dependencies of the core accretion rate on parameters can be obtained by considering the core accretion timescale τc in Mc = Mc τc . (9) Based on the work of Kokubo & Ida (2002), Ida & Lin (2004a) derive an approxi- mate expression for the accretion timescale in the oligarchic growth regime. In the oligarchic regime, the random velocities of the planetesimals is raised by viscous stirring by the protoplanet, while it is damped by gas drag during the presence of 14 Christoph Mordasini the gas disk (Ida & Makino 1993). This regimes occurs after an initial runaway planetesimal accretion phase as soon as the protoplanets have grown to a size of 100-1000 km depending on orbital distance (Ormel et al. 2010). The accretion timescale in this regime is τc = 1.2× 105 yr (cid:32) Σp 10 g cm−2 (cid:32) (cid:33)−1(cid:18) ap 1 AU Σg 2400 g cm−2 (cid:19)1/2(cid:32) Mc (cid:33)1/3(cid:32) M(cid:63) (cid:33)−1/5(cid:18) ap (cid:33)−1/6× (cid:19)1/20(cid:32) m M⊕ M(cid:12) 1 AU 1018 g (cid:33)1/152 . (10) In this equation Σg is the gas surface density at the planet's position at ap, and m is the mass of a planetesimal. We see from Eq. 10 that the growth is faster at smaller distances as the collisional growth scales with the orbital frequency leading to growth wave propagating out- ward. The timescale also increases as M1/3 as typical for the oligarchic regime, and decreases inversely proportional to Σp. This faster growth, and the fact that cores can also become more massive at higher Σp (e.g., Kokubo & Ida 2012) explains why the core accretion theory predicts a higher number of giant planets at higher [Fe/H] (see Sect. Correlations with disk properties). c In pebble based models (see, e.g., Ormel 2017), the accretion of pebble sets in once the protoplanets have reached a size where the encounter with the pebbles transitions from the ballistic to the settling regime. In the former, gas-drag effects are not relevant, whereas in the settling regime the encounter time is sufficiently long to allow the incoming particles to couple aerodynamically to the gas, and to sediment towards the protoplanet during the encounter. This leads to an efficient, gas-drag-aided accretion. This transition occurs at one AU when the protoplanets have reached a size of several hundred km, increasing with orbital distance (Visser & Ormel 2016). Accretion of gas Two approaches are used in global models to calculate a protoplanet's gas accretion rate. The first more complex approach taken for example in the Bern model is to calculate the interior structure of the (proto)planets (Alibert et al. 2005; Mordasini et al. 2012c). The planets' interior is modeled by integrating numerically the 1D spherically symmetric structure equations which are the mass conservation, hydro- static, energy conservation, and energy transport equations (Bodenheimer & Pollack 1986): (cid:32) = 4πr2ρ = 4πr2ρ ∂m ∂r ∂l ∂r ε− P ∂V ∂t − ∂u ∂t (cid:33) ∂P ∂r ∂T ∂r = −Gm r2 ρ ∂P ∂r T P = ∇(T, P) (11) (12) Planetary population synthesis 15 where r is the radius measured from the planet's center, m the enclosed mass, P the pressure, ρ the density, and G the gravitational constant. The gradient ∇ depends on the process by which the energy is transported (radiation or convection). The energy equation is the only equation that is time t dependent and controls the temporal evolution. In this equation, V = 1/ρ is the specific volume, u the specific internal energy, ε an energy source like impact or radiogenic heating, and l is the intrinsic luminosity. These structure equations are solved with different outer boundary conditions depending on the phase a protoplanet is in (Bodenheimer et al. 2000; Mordasini et al. 2012c). In the first so-called attached (or nebular) phase, the envelope transitions smoothly into the background nebula, such that the outer pressure and temperature are approximately equal to the local disk pressure and temperature. In this phase, the outer radius is given as the minimum of a fraction of the Hills sphere and the Bondi radius, and can thus be found if the planet's mass is known. Radiative cooling allows the gas in the protoplanetary envelope to contract. This (formally) results in an empty shell between the planet's outer edge of the envelope and the surrounding nebula. This is filled in by new nebular gas, allowing the envelope mass to increase. This means that during the attached phase, the gas accretion rate is regulated by the envelope's cooling (Kelvin-Helmholtz) timescale as found by solving the structure equations. When the core reaches a mass of about 10 M⊕, the contraction of the envelope becomes so rapid that the protoplanetary disk can no longer supply gas at a rate sufficient to keep the envelope and disk in contact (runaway gas accretion leading to giant planet formation). The planet's outer radius now detaches from the nebula and contracts rapidly, but still quasi-statically (Bodenheimer & Pollack 1986) to a radius that is much smaller than the Hills sphere (about 1.5 - 5 R(cid:88) depending on the entropy, Mordasini et al. 2012c; Mordasini et al. 2017). In this second so-called detached (or transition) phase, the radius is free and found by solving the structure equations, while the gas accretion rate is given by processes in the protoplanetary disk and no longer by the envelope's contraction. This disk-limited rate may be given by the Bondi accretion rate (D'Angelo & Lubow 2008; Mordasini et al. 2012c) (cid:18) RH (cid:19)3 3 Me,Bondi ≈ Σg H Ω (13) where Σg, H, RH, and Ω are the mean gas surface density in the planet's feeding zone, the disk's vertical scale height, the planet's Hill sphere radius, and the orbital frequency at the planet's position. It is also possible to calculate the planet's disk- limited gas accretion rate as a fraction flub of the local viscous accretion rate in the disk, which is in equilibrium given as Me,visc = flub3πνΣg. (14) Hydrodynamical simulations (Lubow et al. 1999) indicate that flub may be as high as 0.9 meaning that the planet accretes 90% of the local gas flow through the disk. At higher masses, gap formation starts to reduce the gas accretion rate, leading to a 16 Christoph Mordasini reduction which can be fitted as (Veras & Armitage 2004) (cid:33)1/3 (cid:32) Mp MJup (cid:33) (cid:32) − Mp 1.5MJup fva04 = 1.668 exp + 0.04. (15) The second simpler approach used by other global models (e.g., Ida & Lin 2004a; Hasegawa & Pudritz 2012; Ndugu et al. 2018) to calculate the gas accretion rate is based on fits for the KH-timescale. The gas accretion rate due to the contraction of the envelope is approximated as Me,KH = Mp τKH (16) where the Kelvin-Helmholtz cooling timescale of the envelope is parameterized as (Ikoma et al. 2000) (cid:33)qKH(cid:32) (cid:32) Mp M⊕ κ 1 g cm−2 (cid:33) τKH = 10pKH yr (17) where pKH and qKH are parameters that are obtained by fitting the accretion rate found with internal structure calculations like Bodenheimer & Pollack (1986); Ikoma et al. (2000); Mordasini et al. (2014). For example, Ida & Lin (2004a) used pKH = 9 and qKH = −3 and neglected the influence of κ. Mordasini et al. (2014) found pKH = 10.4, qKH = −1.5, and κ = 10−2 g/cm2. Once can see from Eq. 17 that the accretion rate is a rapidly increasing function of mass, and that gas accretion becomes important once τKH becomes comparable to, or shorter than, the disk life- time. Compared to the method of solving directly the internal structure, the KH- method is computationally much simpler and more robust. But it cannot take into account how the gas accretion rate depends on (a) the (variable) luminosity of the core because of solid accretion, and (b) the varying outer boundary conditions. This means that it cannot easily recover the spread in associated envelope masses for a fixed core mass visible in Fig. 3. Furthermore it does not yield the planets' internal structure and thus radius and luminosity. But also the solution of the 1D hydrostatic structure equations is only an approximation as it neglects that protoplanetary en- velopes are not closed strictly hydrostatic 1D systems, but that they can exchange gas with the surrounding disk in a hydrodynamic multi-dimensional manner (Ormel et al. 2015). This can delay gas accretion through the advection of high entropy ma- terial (Cimerman et al. 2017). As it is typical for global end-to-end models to rely on low-dimensional approaches (1D or 1+1D) because of computational efficiency, these effects were not considered so far in population syntheses. Figure 3 shows the envelope mass as a function of core mass at the end of the for- mation phase in the population around 1 M(cid:12) stars presented in Sect. Results. These envelope masses were found by solving the aforementioned 1D internal structure equations assuming a grain opacity in the protoplanetary atmospheres that is 0.003 times as large as the ISM grain opacities (Mordasini et al. 2014), and by limiting the Planetary population synthesis 17 Fig. 3: H/He envelope mass as a function of core mass found from solving the internal structure equations for the synthetic population discussed in Sect. Results. The relation is shown at the end of the formation phase when the gaseous disks disperse. The colors show the planets' semimajor axis as log(a/AU). The black line scales as M2.5 core. gas accretion rate in the disk-limited phase similarly to Eq. 13 (see Mordasini et al. (i.e., M2.5 2012c for details). At low masses, the envelope mass scales as M c in the simulation here, indicated by the black line, see also Mordasini et al. 2014). Then, at a core mass of about 5-20 M⊕, gas accretion becomes rapid (runaway ac- cretion), so planets move upwards nearly vertically to higher Me and become giant planets. −qKH+1 c There are fewer planets in the intermediate mass range between about 10 to 100 M⊕. This is because the timescale to accrete this gas mass in runaway accretion is shorter than the disk lifetime, so that it is unlikely that the disk disappears exactly at an intermediate moment/mass. This is the origin of the so called "planetary desert" (Ida & Lin 2004a). It is weaker in the population here compared to the original Ida & Lin (2004a) predictions due to the larger effective qKH = −1.5 instead of -3 as used by Ida & Lin (2004a), meaning that the gas accretion rate does not increase as rapidly with mass, and due to the limits given by the Bondi rate. Regarding the termination of gas accretion, in this simulation, the disk-limited gas accretion de- creases in time just because Σg in the feeding zone (Eq. 13) decreases because of disk evolution. Gas accretion is thus terminated when the gas disk disappears. 10-210-1100101102103104100101102Envelope mass [ME]Core mass [ME]-2-1.5-1-0.5 0 0.5 1 1.5 2 18 Orbital migration Christoph Mordasini The gravitational interaction of the gaseous disk and the embedded protoplanets results in the exchange of angular momentum (for recent reviews see Kley & Nelson 2012; Baruteau et al. 2016), which means that the planets change their semimajor axis, i.e., the undergo orbital migration (Goldreich & Tremaine 1979; Ward 1986; Lin & Papaloizou 1986a). The angular momentum transfer between disk gas and planets via torques leads in most cases to a loss of angular momentum for the planet which means inward migration. The angular momentum J of a planet of mass Mp orbiting a star of mass M∗ at a semimajor axis ap, and the migration rate da/dt given a total torque Γtot = dJ/dt are (cid:113) J = Mp GM∗ap da dt = 2ap Γtot J . (18) Other effects like planetesimal-driven migration (e.g., Levison et al. 2010; Ormel et al. 2012) or Kozai migration due to an external perturber (Kozai 1962; Fabrycky & Tremaine 2007) can also modify the orbits, but were up to now not considered in population synthesis models. Disk-driven migration occurs in two types, Type I and Type II migration. Type I migration occurs if the planet's Hill sphere radius is smaller than the disk's vertical scale height and if the viscous torques are dominant compared to the gravity torques induced by the planet (Crida et al. 2006), meaning that Type I migration applies to low-mass planets. Various descriptions of Type I migration have been derived in the literature. Early derivations (Tanaka et al. 2002) assumed that the disk behaves (locally) isothermal, and predicted rapid inward migration. Later, more realistic cal- culations (e.g., Baruteau & Masset 2008; Casoli & Masset 2009; Paardekooper et al. 2010; Kley et al. 2009) directly modeled the cooling behavior of the disk gas. They showed that there are several sub-types of Type I migration (locally isothermal, adiabatic, (un)saturated corotation torque) that can be identified by considering an number of timescales (Dittkrist et al. 2014). An important finding is that for non- isothermal Type I migration, in some parts of the disk outward migration can occur. Therefore, there are special locations like condensation fronts where the torque van- ishes because of the associated opacity transitions. Such locations can act as traps for migrating planets (Lyra et al. 2010; Hasegawa & Pudritz 2011; S´andor et al. 2011; Kretke & Lin 2012), and can serve as locations of efficient planetary growth (e.g., Horn et al. 2012; Hasegawa & Pudritz 2012). To calculate the torque causing a planet's migration, one needs among other quantities like the gas surface density the local power law exponent of the disk tem- perature pT and of the gas surface density pΣ, which are yielded by the disk model, underlining its importance. The migration timescale in the isothermal approxima- tion used by Ida & Lin (2008a) is given as τtypeI = 1 2.728 + 1.082pΣ apΩ Mp −1 Ω M∗ a2 pΣg (19) (cid:32) cs (cid:33)2 M∗ Planetary population synthesis and the migration rate is then ap = − ap τtypeI 19 (20) which shows that migration speeds up a planet grow more massive. In a more recent analysis, Paardekooper et al. (2010) found that the total torque Γtot resulting from summing up the contributions from the inner and outer Lindblad torques plus the corotation torque can be expressed in an equation of the form Γtot = 1 γ (C0 +C1 pΣ +C2 pT)Γ0 with Γ0 = Σga4 pΩ2 (cid:18) q (cid:19)2 h (21) (22) where γ is the ratio of the heat capacities, h = H/ap the local disk aspect ratio, q = Mp/M∗, Σg the gas surface density at the planet's position, and Ω its Keple- rian frequency. The constants Ci depend on the Type I sub-regime. Their numerical values are listed for example in Dittkrist et al. (2014). Type II migration occurs if the angular momentum injection rate of the planet into the disk is so large that it carves a gap into the gas disk around its location. For global models, several different descriptions of Type II migration were considered in the literature: Ida & Lin (2004a) consider the angular momentum transfer rate in a viscous accretion disk without planets (the viscous torque or "couple" in the terminology of Lynden-Bell & Pringle 1974) and assume that planets in the type II migration regime act as relays that transmit angular momentum also at this rate across their gap via tidal torques. Inserting the viscous torque into Eq. 18, the Type II migration rate is given as ap = 3 sign(ap − Rm)α Σg,mR2 m Mp Ωm Ω Ωm (23) where quantities with the subscript m are evaluated at the radius of maximum vis- cous couple (or velocity reversal, i.e., where the disk changes from accreting to decreting, see Lynden-Bell & Pringle 1974). The position of Rm can either be esti- mated as in Ida & Lin (2004a) (cid:32) Hm (cid:33)2 ap (cid:32) 2t (cid:33) 5τdisk Rm = 10 AUexp . (24) or results automatically from solving the evolutionary equation for the gas surface density (Eq. 3). The Type II migration description of Alibert et al. (2005) assumes that a planet follows the motion of the gas except for the case that the planet is massive compared to the local disk mass, when the planet is assumed to slow down because of its inertia (Alexander & Armitage 2009). The migration rate is thus given as 20 1,  2Σga2 p Mp ap = ur min Christoph Mordasini (25) where ur is the local radial velocity of the accreting gas which is in equilibrium given as 3ν/(2ap). A more realistic yet computationally still feasible approach for a population synthesis (i.e., a 1D approach) is to employ the impuls approximation (Lin & Papaloizou 1986b) to estimate the Type II migration, as for example done in Coleman & Nelson (2014). Given recent results (e.g., Duffell et al. 2014; Durmann & Kley 2015) questioning the classical conception that planets in Type II migration simply follow the viscous evolution of the disk, but that their migration rate is entirely given by the torques, make it likely that migration models will undergo significant modifications in the future. The same is true for Type I migration, where new effects like an additional "heating" torque resulting from the protoplanet's accretional luminosity counteracts inward migration (Ben´ıtez-Llambay et al. 2015), or a "dynamic" corotation torque (Paardekooper 2014; Pierens 2015) that results from the fact that the relative motion of gas and a migrating planet can lead to a feedback (usually, torques are measured for planets at fixed positions). This can lead to outward migration as well. N-body interactions The concurrent formation of several protoplanets in a protoplanetary disk affects the growth history of the protoplanets in multiple ways: the protoplanets compete for the accretion of gas and solids, increase the velocity dispersion of the planetesimals potentially reducing the solid accretion rate of neighboring protoplanets, alter the surface density of planetesimals and for giant planets of the gas, and reduce the ra- dial flux of pebbles. The gravitational interaction between the protoplanetes leads in the case of insufficient damping by the gas disk to the excitation of the eccentricities, resulting in the alteration of the orbits, collisions, and ejections. Orbital migration is affected as well, since the planets can capture into mean motion resonances and migrate together as resonant convoys. For specific parameters, this can even invert the direction of migration (Masset & Snellgrove 2001) and lead to the outward mi- gration of two giant planets, as invoked for the "grand tack" scenario in the solar system (Walsh et al. 2011). As discussed above, all early population synthesis models used the one-embryo- per-disk approach, which was one of the most important limitation of the first gen- eration of the models, in particular for low-mass planets as they usually occur in multiple systems, often in compact configurations (e.g., Mayor et al. 2011). This means that they likely influenced each other during formation. This limitation was addressed in Ida & Lin (2010); Ida et al. (2013) and Alibert et al. (2013). In the Bern model an explicit N-body integrator was added "on top" of the existing sub-models that calculates the N-body interactions and collisions of the concurrently forming protoplanets. In order to keep the computational time sufficiently low for population syntheses, about 20-50 low-mass embryos (0.01-0.1 Planetary population synthesis 21 M⊕) are put into each disk. They interact via the usual Newtonian N-body forces written in the heliocentric system, n(cid:88) ri = −G (M∗ + mi) −G ri r3 i m j j=1, j(cid:44)i  ri − rj (cid:12)(cid:12)(cid:12)ri − rj (cid:12)(cid:12)(cid:12)3  + rj r3 j (26) with i = 1,2,3 . . . N the planet index, ri and mi the heliocentric position and mass of planet i, and M∗ the mass of the central star. The consequences of the gravitational interaction with the gas disk (orbital migration and damping of eccentricities and inclinations) are entered as additional forces into the integrator (Cresswell & Nelson 2008). A different approach was taken by Ida & Lin (2010) and Ida et al. (2013), who developed a new semi-analytical approach to describe the gravitational interactions of several protoplanets in a statistical way based on orbit crossing timescales, in- cluding the effect of resonant capture for migrating planets. The advantage of this approach is a computational cost that is orders of magnitude lower than the direct N-body integration, while still yielding distributions of the eccentricities and semi- major axes of interacting planets that agree well with the direct N-body simulations. Probability distribution of disk initial conditions The second central ingredient for a population synthesis calculation are sets of ini- tial conditions (see Fig. 2). These sets of initial conditions are drawn in a Monte Carlo way from probability distributions. These probability distributions represent the varying properties of protoplanetary disks and are derived as closely as possible from results of disk observations, or, if the quantities are not observable, from theo- retical arguments. Typically, there are at least four Monte Carlo variables employed (Ida & Lin 2004a; Mordasini et al. 2009a): 1. The metallicity and dust-to-gas ratio It is usually assumed that the bulk metallic- ity is identical in the star and its protoplanetary disk. Then, the disk metallicity [M/H] can be modeled as a normal distribution as observed spectroscopically in the photosphere of solar-like stars in the solar neighborhood, with µ = −0.02 and σ=0.22 (Santos et al. 2005). The [M/H] is converted into a disk dust-to-gas ratio (Eq. 5) via fdg = fdg,(cid:12)10[M/H], with a solar fdg,(cid:12) of about 0.01 to 0.02 (Lodders 2003). Together with the initial disk gas mass and the locations of icelines, fdg sets the amount of solids (dust, pebbles, planetesimals) available in the disk for planet formation. 2. The initial disk gas mass The concept of an "initial" disk mass is of course ques- tionable as it results from the dynamical collapse of a molecular cloud core (Shu 1977; Hueso & Guillot 2005), but it could be associated with the disk's mass at the moment when the main infall phase has ended, and no self-gravitational instabilities occur any more. Stability arguments (Shu et al. 1990), the inferred 22 Christoph Mordasini Fig. 4: Distributions of initial conditions for disk around 1 M(cid:12) stars. Top left: Metal- licity. Top right: initial disk gas mass. Bottom left: initial content of planetesimals. Bottom right: lifetime of synthetic disk (blue). The black solid and dotted lines show observationally determined lifetimes by Haisch et al. (2001) and Mamajek (2009), respectively. The horizontal bars shows the typical observational age uncertainty. mass of the MMSN (Weidenschilling 1977; Hayashi 1981), and observations of protoplanetary disk (Andrews et al. 2010; Manara et al. 2016) point towards disk masses of about 0.1 to 10% of the star's mass. The disk masses seem to be distributed roughly log-normally with a mean around 1% of the star's mass (Mor- dasini et al. 2009a), but one should note that this distribution is poorly known. 3. The disk lifetime The observations of IR and UV excesses of young stars indicate that the fraction of stars with protoplanetary disks decreases on a timescale of 1-10 Myr, with a mean lifetime of about 3 Myr (Haisch et al. 2001; Mamajek 2009). In a global model, this timescale can either be set directly in Eq. 6, or is used to find a distribution of photoevaporation rates (Eq. 3) that lead together Planetary population synthesis 23 with viscous accretion to a distribution of lifetimes of the synthetic disks that agrees with observations. 4. The initial starting positions of the embryos Based on the finding of N-body simulations that oligarchs emerge with relative spacings of a few Hill spheres (Kokubo & Ida 2000), a distribution of the starting embryos that is uniform in the log of the semimajor axis is usually used. It is also possible to arrange the embryos such that they "fill" the disk taking into account the asymptotic plan- etesimal isolation mass (Ida & Lin 2010). In the trapped evolution models of Hasegawa & Pudritz (2011); Cridland et al. (2016) embryos rapidly move into traps, so that it is the locus and movement of the traps that effectively gives the formation locations. Other quantities that may also be treated as Monte Carlo variables are for exam- ple the quantities describing the initial radial distribution of the gas and solids in Eq. 4. Other important parameters of the global models like the stellar mass, the plan- etesimals size, or -for viscous accretion disks- the α viscosity parameter (Shakura & Sunyaev 1973) are usually kept constant for one synthetic population, but are var- ied across different populations to understand their statistical impact in parameter studies (e.g., Mordasini et al. 2009b). Figure 4 shows the distributions of the disk (and stellar) metallicities, initial gas disk masses, the mass of planetesimals initially contained in the disks obtained with fdg,(cid:12) = 0.02, and the disk lifetimes. These are the initial conditions for the popu- lation synthesis described below in Sect. Results, containing 504 stars with 1 M(cid:12). Note that in this model, the disk lifetime is not directly set, but results from the combined action of viscous accretion and an appropriately chosen distribution of photoevaporation rates. Results To illustrate what can be obtained from modern population synthesis calculations, we present in the following sections results from the latest generation of the Bern model. The underlying global formation and evolution model used here is very sim- ilar to the model published in Alibert et al. (2013) regarding the N-body interactions of the protoplanets, and to the model published in Mordasini et al. (2012b) regarding the internal structure and long-term thermodynamic evolution (cooling, contraction, envelope evaporation) of the planets. In the new simulations presented here, these two aspects are now combined. This makes it possible to predict not only the or- bital elements and masses, but also the radii and luminosities of planets in multi- planet synthetic systems. As differences to the two previously published models, the evolution of the star is now also considered via the Pisa stellar evolution tracks (Dell'Omodarme et al. 2012), and the Mercury N-body integrator (Chambers 1999) is now employed. As in previous models (e.g., Mordasini et al. 2016), the location of the icelines is calculated with the initial disk structure and remains static in time under the assumption that efficient planetesimal formation happens early, and that 24 Christoph Mordasini for the 300-m planetesimals, drift is not very important, which should be the case at least in the outer nebula (e.g., Piso et al. 2015). Clearly, this is a strong assumption. Initial conditions and parameters Each system initially contains 20 planetary embryos with a starting mass of 0.1 M⊕. These planetary seeds are distributed randomly according to a log-uniform distri- bution between 0.05 and 40 AU. Because of the influence of the initial condition, results concerning synthetic planets that are not clearly more massive than 0.1 M⊕ should be regarded with caution. The stellar mass is in all cases 1 M(cid:12), and 504 star-disk systems are simulated. The formation phase of the systems was simulated during 10 Myr, during which the disks of gas and solid evolve, while the planets ac- crete mass, migrate, and interact and collide via the N-body integrator. Afterwards, the thermodynamic long-term evolution was calculated for 10 Gyr. During this later phase, the planets' mass is constant except for atmospheric escape and no dynamical interactions occur. The initial gas surface density follows the profile in Eq. 4, while the initial plan- etesimal follows a steeper profile ∝ r−1.5 as in the MMSN and a outer exponential radius that is half as large as the one for the gas (in Eq. 4) to account for the in- ward drift of dust (Kornet et al. 2001; Birnstiel & Andrews 2014) and the more concentrated distributions resulting from planetesimal formation (Dra¸zkowska et al. 2016). The planetesimal size is 300 meters, and the α viscosity parameter is 0.002. The grain opacity in the protoplanetary atmospheres during formation is reduced to 0.003 the ISM grain opacity (Mordasini et al. 2014). During evolution, atmospheric opacities of a condensate-free solar-composition gas are assumed (Freedman et al. 2014). Formation tracks Before discussing the statistical results, we present simulations obtained with the global model for one specific system, as the phenomena found in one system often help to understand the statistical results. Figure 5 illustrates the effect of planetary growth, N-body interaction, and orbital migration (Type I and II) for a system taken from the population described above. The effects of general inward migration, resonant capture, collisions, eccentricity excitation by planet-planet interaction, as well as eccentricity damping because of the gas disk can be seen. Starting from 20 embryos that are interacting via the N- body integrator, the system in the end contains 2 giant planets, a hot Neptunian planet, and 1 inner and 2 outer low-mass planets. Orbital migration reduces the orbital distance for several planets by up to a factor 10. The existence of Type I migration traps at opacity transitions (Sect. Orbital mi- Planetary population synthesis 25 Fig. 5: Inward migration, growth, and dynamical interaction in a synthetic system containing initially 20 planetary embryos of 0.1 M⊕ The system is taken from the population synthesis described at the beginning of this section. The tracks of the planets in the time versus orbital distance plane are shown. The black-blue-red lines show the planets' semimajor axes, with the color code representing the planets' mass in Earth masses. The grey lines show the apocenter and pericenter. Lines end when the corresponding protoplanet was either accreted by another body, or ejected because of dynamical interactions. gration), the slowing down of Type II migration because of the giant planets' inertia (Eq. 25), and the finite disk lifetime still prevent the planets from falling into the star. Figure 6 shows growth tracks in the distance-mass plane in the same synthetic system as in Fig. 5. Several effects can be seen: at the beginning, the accretion timescale of solids is much shorter than the migration timescale, leading to nearly vertically rising tracks. With increasing mass, the solid accretion timescale (Eq. 10) becomes longer, while the migration timescale becomes shorter (Eq. 19), so the the planets start to migrate inwards at nearly constant mass once they have grown to about 5-10 M⊕. Some very low-mass planets are also captured into MMRs and pushed inwards by more massive protoplanets. Three protoplanets grow so massive that they trigger runaway gas accretion oc- curring when Mcore ≈ Menv ≈ 10M⊕ as visible from the color code. During gas run- away, the growth tracks are again nearly vertical. Finally, the N-body interaction 0.01 0.1 1 10 0 0.5 1 1.5 2 2.5 3Distance [AU]Time [Myr] 0.1 1 10 100 1000 26 Christoph Mordasini Fig. 6: Growth tracks in the distance-mass plane in the same synthetic system as in Fig. 5. The colored lines show the semimajor axis, color coding the ratio of the H/He envelope mass relative to the to core mass Menv/Mcore. Gray lines show the apocen- ter and pericenter distances. Open circles show the final position of the remaining planets. between the three giant planets increases their eccentricities until their orbits over- lap, as visible from the gray lines in Fig. 5 and 6. At about 1.9 Myr, this leads to the ejection of one giant planet that was located at about 0.5 AU between the two surviving ones. Diversity of planetary system architectures The specific outcome in the simulation shown in Fig. 5 and 6 depends obviously on the initial conditions, and only shows one possible realization. To illustrate the ar- chitecture of planetary systems resulting from the global model and the distributions of initial conditions described in Sect. Probability distribution of disk initial condi- tions, we show in Fig. 7 the final mass-distance diagram of 23 selected synthetic planetary systems. The systems are ordered from bottom left to top right according to increasing metallicity [M/H]. Note that the systems were selected by hand to dis- play the diversity of architectures, and give the incorrect impression that systems with giant planets are common. This is not the case: only about 18 % of all systems 0.1 1 10 100 1000 0.01 0.1 1 10Mass [ME]Orbital distance [AU] 0.01 0.1 1 10 100 Planetary population synthesis 27 have a giant planet (see Sect. Comparison with observations: planet frequencies). To first order, the systems can be split in three classes: Fig. 7: 23 selected synthetic planetary systems in the mass-semimajor axis plane for 1 M(cid:12) stars. The system is taken from the population synthesis described at the beginning of this section. Systems are ordered according to increasing [M/H], the number in each panel. Red points are giant planets with Menv/Mcore > 1. Blue sym- bols are planets that have (partially) accreted volatile material (ices) outside of the iceline(s), while green symbols have only accreted refractory solids. Open green and blue circles have 0.1≤ Menv/Mcore ≤ 1. Filled green points and blue crosses have Menv/Mcore ≤ 0.1. The black horizontal bars go from a− e to a + e. The top right panel is the solar system for comparison. (1) The large majority of the systems are similar to those visible in the bottom left corner: they only contain low-mass planets, with masses of 0.1-10 M⊕. These are systems where the disk properties are such (low surface densities of gas and solids, short disk lifetime) that only little growth occurs during the first 10 Myr. Not much 28 Christoph Mordasini orbital migration and dynamical interaction has occurred because of the low plane- tary masses. Note that further growth on long timescales after 10 Myr is neglected in the model. This could first lead to further accretion, and second it could reduce the number of planets by giant impacts (in these systems the number of final planets is close to the initial number of embryos). The systems have a simple compositional transition from rocky to icy with increasing distance. The more massive planets (∼ 5− 10M⊕) have accreted gas envelopes comparable to Uranus and Neptune, oth- erwise little gas accretion has occurred (and was partially lost after formation by atmospheric escape which is modeled as described in Jin et al. 2014). This type of system is preferentially forming at subsolar [M/H], but note that they also exist at high [M/H] (for example the [M/H]=0.15 and 0.39 systems). This makes clear that all four Monte Carlo variables play an important role in determining the outcome of the formation process. (2) Already much less common are systems with giant planets and low-mass planets. Some contain only rocky low-mass planets inside of the giant planets (the [M/H]=-0.02 system), some only icy planets outside of them (the [M/H]=0.06 sys- tem). Some are also reminiscent of the solar system and contain both inner terrestrial planets and outer icy planets (see the [M/H]=-0.00 system). Such a small-large- small arrangement is the classical outcome for collisional growth from planetesi- mals with little orbital migration (or redistribution of solids in general): inside, the low availability of solids prevents much growth, while outside it is the long growth timescale that keeps planet masses low. The sweet spot for giant plant growth is a region outside of the water iceline. The architecture of the giant planets varies significantly: the number of giant planets in a system varies from 1 to 5; in some systems the giants' mass increases with distance, in others it increases, and in some it is fairly constant; the eccentricities also range from near-zero value in many cases to higher values of about 0.2. However, despite the diversity, there is one common characteristic that distin- guishes almost all synthetic systems from the solar system: the innermost giant planet is clearly closer-in than Jupiter, namely at about 1 AU or even less. This is an intriguing result. For the solar system, the "grand tack" model (Walsh et al. 2011) suggests that Jupiter (and Saturn) also migrated to about 1.5 AU, to then migrate outward because of the Masset-Snellgroove effect occurring for resonantly coupled giants (Masset & Snellgrove 2001). The way Type II migration is calculated in the model here from the gas' radial velocity (Eq. 25) does not allow such outward mi- gration. It will be interesting to see whether the inclusion of more realistic migration models (see Sect. Orbital migration) will also lead to more distant synthetic giant planets. (3) In a small fraction of systems, only one massive giant planet remains at the end. Such systems form in metal-rich and massive disk where several giant plan- ets form in vicinity, leading to violent planet-planet scattering. In Plot 7, in the [M/H]=0.10 and 0.40 systems such giant planets with masses exceeding 30 M(cid:88) can be seen with semimajor axes of 20-50 AU and high eccentricities. This gives them pericenter distances of less than 10 AU where the scattering occurred and apocen- ter distances approaching in one case of about 100 AU. They may be detectable by Planetary population synthesis 29 direct imaging. In such systems, all other planets were either accreted, sent into the star, or ejected. The a− M distribution Fig. 8: Synthetic mass-distance diagram. Points shows the semimajor axis, while gray horizontal bar go from a− e to a + e. Ejected planets are shown at 100 AU. As in Fig. 7, the colors and symbols show the planets' bulk composition. The black crosses represent the solar system planets. To finally get a statistical overview of the diversity of planetary systems, Figure 8 shows the superposition of all 504 individual systems in the mass-distance plane. This diagram is of similar importance for (extrasolar) planets as the Hertzsprung- Russell diagram for stars. 30 Christoph Mordasini The first, and most fundamental results is that the variation of the initial condi- tions over a range indicated by observations of protoplanetary disks leads to a high diversity of planetary systems which covers a large part (but not all) of the param- eter space that was found to be covered by the observation of extrasolar planets (compare with Fig. 1). The plot also visually highlights the prevalence of low-mass planets that was already a key result of the first population syntheses (Ida & Lin 2004a). The prevalence is further quantified with the planetary mass function (Sect. The planetary mass function and the distributions of a, R and L) and the planet frequencies discussed in Sect. Comparison with observations: planet frequencies. Considering the solar system, we see that the terrestrial planets, Jupiter, and Uranus are in regions that are well populated with synthetic planets, whereas Sat- urn and Neptune are rather on the outer edge of the populated envelope, at least for the 504 synthetic system shown here. As mentioned, this could be linked to a more compact early configuration of the system as predicted by the Nice Model, where Saturn and Neptune were at about 8 and 14 AU, respectively (Gomes et al. 2005). For the giant planets, a certain pile-up is see around 1 AU, similar as in the ob- served distribution (Sect. Distributions of planetary properties). It is a consequence of the existence of a preferred formation location for giant planets outside of the water iceline, and a typical inward migration of several (1-10) AU. The finite ex- tent of migration is due to the non-negligible type II migration timescale (≈ viscous timescale), the slowing down in the inner system (Eq. 25), and the typical finite disk lifetimes that are comparable to the formation timescale of the giant planets. This mean that giants often migrate in evolved disk with a mass that has already significantly decreased, slowing down orbital migration (Mordasini et al. 2012a). In terms of the bulk composition of the core, we see that close-in low-mass plan- ets have an Earth-like composition (green symbols in Fig. 8) as they did not accrete outside of the iceline. But there are also close-in, more massive (sub-)Neptunian planets that have started to form outside of the iceline giving them significant ice mass fractions indicated by blue symbols. They then migrated in through a "hori- zontal branch" (Mordasini et al. 2009a), as the positive corrotation torque saturates at such masses, leaving only the negative Lindblad torques, which drives fast inward Type I migration. Starting with about 50% ice in mass in the core while accreting outside of the water iceline, they eventually have an ice mass fraction in the core of about 10-20%, as they accrete rocky planetesimals during their migration through the inner system, such that composition-wise, they are not really Neptune-like. This phenomenon was previously seen in simulations for GJ 436 b (Figueira et al. 2009). We also see that planets with masses of about 10− 30M⊕ have a H/He mass fraction of Menv/Mcore=0.1-1, and more massive planets are giants where Menv/Mcore > 1. A quite populated planetary desert Compared to early population syntheses in particular from the Ida & Lin model, there is no strong "planetary desert" (absence of intermediate mass planets) visible in Fig. 8, even though a certain dip in the mass function at intermediate masses Planetary population synthesis 31 of about 30-100 M⊕ can still be seen in the mass function (Fig. 10). As partially discussed before (Sect. Accretion of gas), the reason is mainly three-fold: First, in the models shown here, the heating from planetesimal accretion leads to a τKH that decreases less rapidly with increasing mass compared to the Ida & Lin (2004a) model, as discussed in Sect. Accretion of gas, meaning that planets move less rapidly through the intermediate mass regime. Thus, the probability that the gas disk disappears during this time is higher. Second, the gas accretion rates obtained in the disk-limited phase calculated with Eq. 13 is often only a few 10−4 M⊕/yr. The reason is that cores often only reach a mass sufficient to trigger gas runaway in advanced stages of disk evolution, when the gas surface density has already decreased significantly. Such a timing at first appears unlikely; it is not, if we consider that in most disks, planetary growth is so slow that cores sufficiently massive to trigger gas runaway never form during the gas disk's lifetime (the frequency of stars with giant planets is at most 20%). So a late formation is actually probable. This seems to be a difference to pebble-based models (Bitsch et al. 2015b). The removal of disks before cores have a chance to undergo runaway growth also gives naturally rise to a population of numerous super-Earths (Hasegawa & Pudritz 2012; Alessi et al. 2017). Third, in contrast to earlier simulations, we find a significant multiplicity of giant planets (see Sect. Comparison with observations: planet frequencies), meaning that individual proto-giants compete for gas while growing, reducing further the maxi- mal gas accretion rates, and leading to more intermediate mass planets. The a− R distribution The results of the Kepler mission (Borucki et al. 2011) has given us a unique insight into the statistics of close, mostly small planets (e.g., Howard et al. 2012; Fressin et al. 2013; Petigura et al. 2013; Mulders et al. 2015; Petigura et al. 2018). As a tran- sit mission, it however yields planetary radii, and not masses, and no unique relation exists that links mass and radius (e.g., Wolfgang et al. 2016). Keeping track of the basic material type that a planet accretes (iron, silicates, ices and H/He) combined with the calculation of the evolution of its internal structure (Eq. 11) make it pos- sible to predict radii from a global model for a direct comparison (Mordasini et al. 2012b). Fig. 9 shows the synthetic distance-radius diagram at 5 Gyr. In contrast to the mass-distance relation, there is a still signifiant evolution of the radii also after the dissipation of the gas disk because of contraction and atmospheric escape. The plot shows that giant planets with Menv/Mcore > 1 have radii larger than 6-7 R⊕, while intermediate planets with 0.1≤ Menv/Mcore ≤ 1 have radii larger than 3-4 R⊕. To zero order, the planets of intermediate radii have a frequency that is uniform in log(R). Two prominent feature can be seen: First, the radii of most giant planets fall in a relatively narrow range of about 10- 12.4 R⊕ (0.9 to 1.1 R(cid:88), where R(cid:88) is the radius of Jupiter). This is expected (Mordasini 32 Christoph Mordasini Fig. 9: Synthetic planet radius-semimajor axis diagram at 5 Gyr. The colors and symbols are the same as in Fig. 7. Small black open circles additionally show plan- ets that have lost the entire primordial H/He envelope by atmospheric escape. The yellow line shows the location of the gap determined observationally by Van Eylen et al. (2017). et al. 2012b), as the mass-radius relation of giant planets between about the mass of Saturn and well into the brown dwarf regime is such that the radius is nearly independent of mass, and always around 1 R(cid:88) (Chabrier et al. 2009). The reason is that the interiors become more and more compressible with increasing mass. In the synthetic population here the pile-up is exaggerated as all planets have during evolution the same solar-composition opacity in the atmosphere (Freedman et al. 2014). Varying opacities would cause the radii to vary more (Burrows et al. 2007). Additionally, no bloating effects are included (for an overview, see Baruteau et al. 2016). A second prominent feature is the gap running diagonally downwards with in- creasing semimajor axis at small radii. It separates inside of 1 AU planets that have Planetary population synthesis 33 kept or lost the primordial H/He. Close-in low-mass planets lose their envelope be- cause the binding energy of their H/He envelope is small compared to the incoming stellar XUV radiation that the planets are exposed to. Note that this evaporation valley was theoretically predicted by models (Owen & Wu 2013; Lopez & Fortney 2013) including population synthesis models (Jin et al. 2014) before it was observed (Fulton et al. 2017; Van Eylen et al. 2017). As expected for a simple energy-limited evaporation model with a constant efficiency factor, the model here predicts a slope that is somewhat steeper than observed (Owen & Wu 2017; Van Eylen et al. 2017). The Earth-like rocky composition of the planets in the region of the gap predicted in the synthesis is consistent with the observed location of the gap (Owen & Wu 2017; Jin & Mordasini 2018). The planetary mass function and the distributions of a, R and L Figure 10 shows four distributions of fundamental planetary properties. The top left panel shows the mass distribution (including planets at all semimajor axes) P-MF which is also shown with a linear y-axis in Fig. 11. The prediction of the mass function is a key goal of population synthesis. We see a result that is qualitatively similar to earlier results obtained in the one-embryo-per-disk simplification (Mor- dasini et al. 2009b), and characteristic for the core accretion paradigm: two main regimes exist, one below about 30 M⊕ which corresponds to planets with a com- position dominated by solids, and one consisting of gas-dominated giant planets at higher masses. The break at around 30 M⊕ corresponds to a state when (critical) core and envelope mass are approximately equal, just before gas runaway accretion occurs. The most fundamental aspect of the core accretion paradigm - the existence of a critical core mass - is thus imprinted into the planetary mass function. In the two regimes below and above 30 M⊕ respectively, different physical mech- anisms (the accretion of solids vs. the accretion of gas) control the growth of the planets, leading to two different slopes of the mass function. These two basic regimes exist both in earlier syntheses employing the one-embryo-per-disk simpli- fication and in newer ones with numerous concurrently forming protoplanets (as here), explaining why qualitatively, the mass function is similar. Above 30 M⊕, the mass function is to zero approximation flat in log(M) up to about 5 M(cid:88), i.e., the number N of planets scales with mass M as N ∝ M−1. Below the break at about 30 M⊕, the dependency is steeper, and scales roughly like N ∝ M−2. Finally, towards the upper end of the planetary mass function above about 5 M(cid:88), the decrease follows a similar steep scaling. These scalings are shown in Fig. 10 with dotted red lines. The semimajor axis distribution of intermediate and low-mass planets is charac- terized by a rapid rise in the frequency between 0.01 and 0.1 AU, followed by a large interval between 0.1 and almost 10 AU where the distribution is approximately flat (or slightly decreasing) in log(a). This is similar as indicate by observations (Pe- tigura et al. 2018) for a (cid:46) 1 AU. It indicates that despite orbital (Type I) migration 34 Christoph Mordasini Fig. 10: Distributions of fundamental planetary properties in the synthetic popula- tion. Top left: planetary mass function P-MF. The red dotted lines show scalings discussed in the text. Top right: semimajor axis distribution for three mass intervals (blue > 30M⊕; red 10-30 M⊕; black 1-10 M⊕). Bottom left: bolometric luminosity. Bottom right: radius. The luminosity is shown at 20 Myr, the other panels are for 5 Gyr. that is included in the model without artificial reduction factors, and which leads for many planets to a significant reduction of the semimajor axis by factors of around 4- 10 or even more relative to the starting position, it nevertheless preserves the initial distribution that is uniform in log(a) as well. Giant planets are restricted to smaller semimajor axis range (about 0.1 to 6 AU), with a rapid drop both inside and outside and outside this distance. The distribution peaks a bit inside of 1 AU. The luminosity distribution - shown at 20 Myr - mainly traces the mass distri- bution (Mordasini et al. 2017) because of the power law relation between mass and luminosity approximately given as L ∝ M2 at a fixed time. Compared to the mass Planetary population synthesis 35 distribution, a third local maximum appears at around log(L/L(cid:12)) ≈ −3.5 which is caused by deuterium burning planets (Molli`ere & Mordasini 2012). The distribution of the radii of the planets (including planets at all semimajor axis) finally is also similar as in simulations using the one-embryo-per-disk simpli- fication (Mordasini et al. 2012b), and contains a local maximum at around 1 Jovian radius (for reasons discussed in Sect. The a− R distribution), and a relatively con- tinuous raise towards smaller radii. This is caused by the prevalence of low-mass planets and the fact that their KH-timescale for gas accretion is long, such that they have small H/He mass fractions and thus also small radii (Mordasini et al. 2012b). This KH-timescale effect, and the EOS linking mass, bulk composition, and radius is the same as in the one-embryo-per-disk case, explaining the similarity. Compared to the input distributions of the starting planetary embryos (initial mass of 0.1 M⊕ for all seeds, semimajor axes uniformly distributed in log(a) be- tween 0.05 and 40 AU, all seeds put into the disk at the beginning of the simulation), the final mass and radius distributions are very different, and contain some specific physically explainable sub-structures. This suggest that the specific initial mass of the embryos has not directly influenced these distributions, at least for planets with M (cid:29) 0.1M⊕. On the other hand, the final semimajor axis distribution is still -to zero order- uniform in log(a) for planets with masses between 1 and 30 M⊕. This could indicate that the input semimajor axis distribution influences this result, and that the final distribution would differ if another initial distribution would be used. Such dif- ferent distributions seem quite possible, for example because of preferred formation locations of the planetesimals (e.g., Dra¸zkowska et al. 2016), particle pile-ups out- side of orbits of already existing planets (e.g., Pinilla et al. 2015) or strong migration traps (e.g., Horn et al. 2012; Hasegawa & Pudritz 2012). This shows the necessity to include the earlier stages of planet formation in future global models, namely the dust, pebble and planetesimal formation stages. Comparison with observations: planet frequencies Before comparing observed and synthetic distributions, we address the frequency of three fundamental planet types predicted by the synthesis: first, giant planets with a mass of at least 300 M⊕, second close-in planets (period ≤ 100 d, i.e., a ≤ 0.42 AU, radius ≥ 1R⊕) comparable to the planets probed by Kepler (e.g., Marcy et al. 2014; Petigura et al. 2018), and third planets in the classical habitable zone with mass of 0.3 to 5 M⊕ and a semimajor axis of 0.95 to 1.37 AU (Kasting et al. 1993). We give the overall fraction of stars having such planets, indicating also their multiplicity. For the comparison with observed frequencies (Sect. Frequencies of planet types), one should keep in mind that the synthetically predicted absolute frequencies are less robust than the relative ones. The reason is that the absolute frequencies depend more directly on model parameters like the arbitrary chosen planetesimal size of 300 m which influences the solid accretion rate, as discussed in Sect. Accretion of solids. 36 Christoph Mordasini Assuming a smaller (bigger) size of the planetesimals would increase (decrease) the fraction of disks in which massive planets form. N Giant planets Close-in planets Planets in HZ 1 2 3 4 ≥5 Overall synthetic Overall observed 4.8 7.4 5.4 0.4 0.0 18.0 10-20 8.4 12.8 11.4 10.0 11.4 54.0 50-60 30.7 8.2 1.0 0.0 0.0 39.9 unknown Table 1: Percentage of stars with N planets of the given type in the synthetic popu- lation, and comparison to observations (last line). We see that the overall fraction of stars with giant planets is about 18%. The most frequent number of giant planets per star is 2, occurring for 7.4% of all stars. Single giant planets and stars with 3 giants are both on the 5% level. Only 0.4 % of the stars have 4 giant planets, and none has more than that. Note that these numbers may be upper limits, as the calculation of the N-body interaction was stopped at 10 Myr, such that we do not take into account ejections or collisions at later moments. The overall frequency of giant planets is however similar as observed (10-20%, see Sect. Frequencies of planet types), even if the multiplicity in the synthetic population might be higher than observed (Bryan et al. 2016). The population of close-in planets is dominated by small (or low-mass) planets as visible from Figs. 8 and 9. Even if a quantitative comparison with the HARPS high-precision RV survey or the Kepler transit survey would require a dedicated modeling of the observational biases (e.g., Mayor et al. 2011; Petigura et al. 2018), the frequencies of these close-in planets in Table 1 makes nevertheless clear that such planets are a very common outcome of the formation process, similarly as ob- served (Sect. Frequencies of planet types). About 54% of the synthetic systems have such planets which is similar to the observed frequency (Petigura et al. 2013), and the multiplicity is high, with a mean number of about 3 such planets per star that has this type of planet, which is again at least qualitatively similar as observed. This is high frequency of close-in planets can only be reproduced in the syntheses if a steep, centrally concentrated distribution of the planetesimals is used, as described in Sect. Initial conditions and parameters (see also Chiang & Laughlin 2013). This is an in- teresting constraint for drift and planetesimal formation models (e.g., Dra¸zkowska et al. 2016). The fraction of stars with planets in the classical habitable zone is in contrast lower, but with 39.9% still very significant. The mean number of this type of planet is 1.25 for those stars that have such planets, i.e. typically there is only one planet in the classical habitable zone per such system. Observationally, the frequency of solar- like stars with potentially habitable planets is still now well known, with estimates Planetary population synthesis 37 ranging from 1 to 100% (e.g., Burke et al. 2015). For M-dwarfs, where a direct determination of this frequency is in contrast already possible with radial velocity surveys, the fraction is 0.41+0.54−0.13 (Bonfils et al. 2013). Comparison with observations: distributions Figure 11 compares the synthetic mass and radius distribution with their observa- tional counterparts as found by the HARPS high precision RV survey (Mayor et al. 2011) and an analysis of the Kepler transit survey (Howard et al. 2012). The ob- served distributions are corrected for the observational bias. The synthetic distribu- tions only include the planets in the same mass/radius and orbital distance range as in the observational samples. Fig. 11: Comparison of the synthetic mass and radius distribution (black) of the population introduced at the beginning of Sect. Results with the bias-corrected ob- served distributions (blue) of the HARPS high precision RV survey (Mayor et al. 2011) and the Kepler transit survey (Howard et al. 2012). The two theoretical and observed distributions were normalized to the same value at the bin at 30 M⊕ and 2-3 R⊕, respectively. The basic shape echoes the distributions that include all synthetic planets shown in Fig. 10. The synthetic mass distribution compares quite well with the observed one, in particular regarding the aforementioned break in the mass function at about 30 M⊕ and the associated change of the slope that was already predicted in early population syntheses (Mordasini et al. 2009b). This break is visible also in other bias-corrected high-precision RV surveys like Howard et al. (2010), and even in the biased, directly observed mass distribution (Schneider et al. 2011; Wright et al. 2011). The log-flat distribution between about 30 M⊕ and 5 M(cid:88) agrees with the ob- 38 Christoph Mordasini served distribution (Marcy et al. 2005) as well. The fact that the two theoretically predicted slopes are visible also in the observed distribution, and that the break oc- curs at a similar mass in theory and observations constitutes a major success of the core accretion theory, and planet formation theory in general. This is of an astro- physical importance comparable to the development of a theory for the stellar ini- tial mass function (Chabrier 2003), including the classical Salpeter slope (Salpeter 1955). In the radius distribution of synthetic planets inside of 0.27 AU, a significant difference is seen for the peak at about 1 Jovian radius relative to the distribution that includes all orbital distance in Fig. 10, as already found in Mordasini et al. (2012b): the peak is less pronounced compared to Fig. 10 because of the broader bins, and due to the fact that only planets inside of 0.27 AU are included (as in the observational sample), while synthetic giant planets are mostly further out (see Fig. 9). In these broad bins, the evaporation valley and the associated gap in the radius distribution is not visible as a finer radius resolution is required (Jin & Mordasini 2018). Another difference is that the synthetic distribution is less abruptly increasing towards the small radii compared to the observed one. We note that the population contains many planets with masses still lower than 1 M⊕ and/or radii less than 1 R⊕. They are found in systems in which not much growth and migration occurred during the first 10 Myr (the time during which the systems' formation was simulated as explained in Sect. Initial conditions and parameters). At 10 Myr, there are still around 20 low-mass protoplanets left. Such systems arise for initial conditions with low amounts of solids (low [Fe/H] and/or initial gas disk masses) and/or short disk lifetimes (see Mordasini et al. 2012a for an extensive dis- cussion of the correlations between disk and planetary properties). This dependency is illustrated by the systems forming at low [Fe/H] in Fig. 7. Over longer timescales, these low-mass planets could collide to form more mas- sive planets, such that this result could be an artifact of only modeling the N-body interaction during 10 Myr. This also means that all results concerning planets with masses close to the initial embryo mass should be taken with caution. Correlations with disk properties An important application of population synthesis is to understand how the planetary formation process depends on the properties of the protoplanetary disk. The most important observed correlation in this context is that the probability of observing giant planets increases with the host star metallicity (e.g., Gonzalez 1997; Santos et al. 2004; Fischer & Valenti 2005), the so-called "metallicity effect", as discussed in Sect Correlations with stellar properties. Syntheses by Ida & Lin (2004b) and later Mordasini et al. (2009b, 2012a) have shown quantitatively that this metallicity effect is a natural outcome of the core accretion model. This is under the assumption that stellar and disk metallicity are proportional (Sect. Initial conditions and parameters), and further, that the surface density of planetesimals increases with the mass fraction Planetary population synthesis 39 of heavy elements as well (Eq. 5), which is indicated by planetesimal formation models (Brauer et al. 2008). In this case, the growth of a planetary core by accreting planetesimals occurs on a shorter timescale (Eq. 10), and leads to more massive cores as well (Kokubo & Ida 2012). Thus, there is a higher chance to grow to the critical core mass and trigger rapid gas accretion before the gas disk has dissipated. Fig. 12: Fraction of stars with giant planets (M ≥ 300M⊕) as found in the synthetic population (black line). The blue line shows the fit to the observed frequency from Mortier et al. (2013). Figure 12 shows the fraction of stars with at least one giant planet (mass higher then 300 M⊕) in the present synthesis for 1 M(cid:12) stars as a function of metallicity, compared to a fit to the observed relation by Mortier et al. (2013). While the model over-predicts the number of giant planets in absolute terms, it agrees with the ob- served relative increase quite well. Further correlation of disk properties were studied by Mordasini et al. (2012a) where - not surprisingly - a high number of correlations was found: for example, for the planetary initial mass function, high metallicities lead to a higher frequency of giant planets while higher initial disk (gas) masses lead mainly to giant planets of a higher mass. For long disk lifetimes, giant planets are both more frequent and massive. At low metallicities, very massive giant planets cannot form, but otherwise giant planet mass and metallicity are nearly uncorrelated. In contrast, (maximum) planet masses and disk gas masses are correlated. Testing theoretical sub-models The final goal of population synthesis is to improve our understanding of planet formation and evolution. For this task, specific sub-models are put via syntheses 40 Christoph Mordasini to the observational test as described in Sect. Workflow of the population synthe- sis method. A non-conclusive list of mechanisms that were addressed in this way is (1) orbital migration, mostly (non)isothermal Type I migration. Early population syntheses (Ida & Lin 2008a; Mordasini et al. 2009b) showed that the observed dis- tribution of planetary orbital distances and the fraction of stars with hot Jupiters can only be reproduced if the migration rates predicted by the then existing (isothermal) type I migration models (Tanaka et al. 2002) are strongly reduced. This sparked numerous dedicated studies that led to more realistic non-isothermal type I migra- tion rates (like Masset & Casoli 2010; Kley et al. 2009; Paardekooper et al. 2010). These new models where then in turn included in the population syntheses, leading to orbital distances that are more similar to observations (Dittkrist et al. 2014). This is a prime example of how population synthesis and specialized models advance each other. (2) grain dynamics and opacities in protoplanetary atmospheres influ- encing the bulk composition of planets. Here it was found with population synthe- ses that the observed mass-radius relation (i.e., the bulk composition) of extrasolar planets with H/He can be reproduced only if the grain opacity in protoplanetary at- mospheres is clearly lower than the ISM opacity (Mordasini et al. 2014). This led to the development of specialized models for the grain dynamics (growth, settling, destruction) and resulting opacities (Ormel 2014; Mordasini 2014), which are in- deed low compared to the ISM. (3) disk inhomogeneities and transitions leading to migration traps (Hasegawa & Pudritz 2011; Hasegawa & Pudritz 2013; Coleman & Nelson 2016), (4) stellar cluster environments (Ndugu et al. 2018), (5) the gas ac- cretion shock structure and the luminosity of young giant planets (Mordasini et al. 2017), (6) constraints on formation pathways from the chemical composition and atmospheric spectra (e.g., Marboeuf et al. 2014; Madhusudhan et al. 2014), or fi- nally (7) atmospheric escape of primordial H/He envelopes (Jin et al. 2014; Jin & Mordasini 2018). Predictions: observational confirmations and rejections Another central application of population syntheses are predictions for upcoming instruments and surveys, i.e., quantitative theoretical predictions that can be falsified with more accurate observational methods. Mordasini et al. (2009b) for example studied the consequences for the fraction of stars with detectable planets and the shape of the P-MF if the radial velocity measurement accuracy improves from 10 m/s to 1 m/s and 0.1 m/s. Some of the predictions made by population syntheses were later confirmed by observations, others turned out to be inconsistent. Some of the most important con- firmed predictions are: (1) the prevalence of low-mass and small planets (Ida & Lin 2004b) that was made well before they were found by precise RV surveys and Ke- pler (e.g., Howard et al. 2010; Mayor et al. 2011; Borucki et al. 2011), (2) the break in the planetary mass function at around 30 M⊕ (Mordasini et al. 2009b) that was later detected by Howard et al. (2010); Mayor et al. (2011) through high-precision Planetary population synthesis 41 RV, (3) the pile-up of planetary radii around 1 R(cid:88) (Mordasini et al. 2012b) which was not visible in early polluted Kepler data but which is has become apparent re- cently in cleaned samples (Petigura et al. 2018), or (4) together with other models (Owen & Wu 2013; Lopez & Fortney 2013), the depleted evaporation valley in the distance-radius plane (Jin et al. 2014), which was later confirmed observationally by Fulton et al. (2017) and Van Eylen et al. (2017). Important predictions that turned out to be inconsistent with observations were (1) the existence of a strongly depleted "planetary desert" (Ida & Lin 2004b), i.e., a strong depletion in the frequency of planets with masses between 10 to 100 M⊕ (a weak depletion might actually exist, see Fig. 1), or (2) an absence of close-in planets (a (cid:46) 0.1 AU) with masses less than approximately 10 M⊕ (Mordasini et al. 2009a). These inconsistent predictions are actually of particular interest, as they point at important shortcomings in the theoretical models. In the former case, the gas accre- tion rate in the runaway phase was overestimated (see Sect. Accretion of gas and A quite populated planetary desert for limiting effects), in the latter, a strongly re- duced isothermal type I migration rate and a criterion for the transition into type II migration based only on the thermal criterion caused the discrepancy. These short- comings were then addressed in later generations of the models (Sect. Overview of population synthesis models in the literature), and helped in this way to improve the understanding of the planet formation process, and to avoid oversimplifications. Summary and conclusions The increase in statistical observational constraints on extrasolar planets has been enormous in the last two decades. Both ground and space-based surveys have de- rived distributions of fundamental planetary properties like the frequency of planets in the mass-distance and radius-distance planes, the planetary mass function, the eccentricity distribution, or the planetary mass-radius relation. All these observed distributions put strong statistical constraints on the theory of planet formation and evolution (Sect. Statistical observational constraints). The method of choice to use these constraints in order to improve our understanding of planet formation is populations synthesis, an approach that has been used for many decades in various fields of stellar astrophysics. The underlying idea of population synthesis is that the same physical processes govern the formation of planets in all protoplanetary disk, but that the initial con- ditions for these processes (the properties of the parent protoplanetary disk) and potentially the boundary conditions (like the stellar cluster environment) differ, and that this gives raise to the observed diversity of (extrasolar) planets. Methodically, population syntheses (Sect. Population synthesis method) thus consist of two main components: first, probability distributions of the initial con- ditions (disk properties, Sect. Initial conditions and parameters) and second a global end-to-end model of planet formation and evolution that can predict observable planetary properties based directly on disk properties. Building both on the core 42 Christoph Mordasini accretion and the gravitational instability scenario, the different modern population synthesis models in the literature (Sect. Overview of population synthesis models in the literature) include in a self-consistently coupled way an impressive number of physical processes like the evolution of the protoplanetary disks of solids and gas, planetary accretion of solids (both of planetesimals and pebbles), the accretion of gas, orbital migration, and N-body interactions, and often several more. A de- scription of these sub-models can be found in Sect. Global models: simplified but linked. The sub-models describing these processes are, however, either parameterized or low-dimensional static approximations of 3D dynamical systems (Sect. Low- dimensional approximation). To what extent the dynamical multi-dimensional na- ture of one individual governing process (like orbital migration or pebble accretion) can be "distilled" into a simpler, lower-dimensional approximation that still captures the essence of the physics while allowing population syntheses with acceptable com- putational costs, is a key challenge for any population synthesis approach, and an ongoing development. On the other hand, population synthesis is often the only possibility to obser- vationally test theoretical models of a specific mechanism (Sect. Testing theoreti- cal sub-models) as it produces synthetic data that can be directly statistically com- pared with observations, factoring in the non-linear interaction between the different mechanism concurrently acting during planet formation. This leads to planetary for- mation tracks and synthetic planetary systems of a large diversity (Sect. Diversity of planetary system architectures). Population synthesis therefore also has a high pre- dictive power that is difficult to achieve with theoretical models of just one physical process. In the Section Results, the typical output obtained from a population synthesis model is presented and linked to underlying physical effects. These are the distri- bution of planets in the mass-distance and radius-distance plane, the planetary mass function, and the distribution of radii, orbital distances, and luminosities. These dis- tributions are compared with their observed counterparts (Sect. Comparison with observations: distributions). Among these results, a first key prediction of population synthesis models (Sect. The a − M distribution) was the prevalence of low-mass and small planets that was later observationally confirmed by high-precision radial velocity surveys and the Kepler satellite (Sect. Predictions: observational confirmations and rejections). Clearly this has important implications beyond planet formation theory like for the question about the existence of other habitable planets. A second key prediction was the existence of two regimes in the planetary mass function (Sect. The planetary mass function and the distributions of a, R and L) with a break at around 30 M⊕ when the transition from solid to gas-dominated planet occurs at the critical core mass. The most fundamental aspect of the core accretion paradigm - the existence of a critical core mass - is thus imprinted into the planetary mass function. This prediction was later confirmed by high-precision RV surveys (Sect. Predictions: observational confirmations and rejections). The finding that the two theoretically predicted slopes in the mass function are visible also in Planetary population synthesis 43 the observed distribution, and that the observed break occurs at a similar mass as theoretically predicted represents a major success for the core accretion theory, and of planet formation theory in general. This is in a broader astrophysical context of similar importance as the development of a theory for the stellar initial mass func- tion. A third, maybe even more fundamental insight is that the population syntheses show that the variation of the initial conditions over a range suggested by proto- planetary disk observations lead to an extreme diversity in the resulting synthetic planetary systems, which is probably the single most characteristic property also of the actual extrasolar planet population. It indicates that at least some of the strong non-linearities and feedback mechanisms occurring during planet formation are in- deed captured in the theoretical models. These points do, however, clearly not mean that the current population models describe in a definitive way the actual planet formation and evolution process. They rather reflect the state of the field of planet formation theory where important physi- cal mechanisms governing planet formation are still not well understood (Sect. Con- fronting theory and observation). It is therefore not surprising that other important predictions made by population synthesis models turned out to be inconsistent with observations (Sect. Predictions: observational confirmations and rejections). But it is exactly the quantitative falsifability of population synthesis that is important, as this it is the key to reject some theoretical concepts or to identify missing ones, improving in this way our understanding of how planets form and evolve. Acknowledgements: C.M. thanks Ralph Pudritz for the invitation to write this review and for the editorial guidance. C.M. acknowledges the support from the Swiss National Science Founda- tion under grant BSSGI0 155816 "PlanetsInTime". Parts of this work have been carried out within the frame of the National Center for Competence in Research PlanetS supported by the SNSF. References Alessi, M., Pudritz, R. E., & Cridland, A. J. 2017, MNRAS, 464, 428 Alexander, R., Pascucci, I., Andrews, S., Armitage, P., & Cieza, L. 2014, Protostars and Planets VI, 475 Alexander, R. D. & Armitage, P. J. 2009, ApJ, 704, 989 Alibert, Y., Carron, F., Fortier, A., et al. 2013, A&A, 558, A109 Alibert, Y., Mordasini, C., & Benz, W. 2004, A&A, 417, L25 Alibert, Y., Mordasini, C., Benz, W., & Winisdoerffer, C. 2005, A&A, 434, 343 Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2010, ApJ, 723, 1241 Bai, X.-N. 2016, ApJ, 821, 80 Baruteau, C., Bai, X., Mordasini, C., & Molli`ere, P. 2016, Space Sci.Rev., 205, 77 Baruteau, C. & Masset, F. S. 2008, ApJ, 672, 1054 Ben´ıtez-Llambay, P., Masset, F., Koenigsberger, G., & Szul´agyi, J. 2015, Nature, 520, 63 Benz, W., Ida, S., Alibert, Y., Lin, D., & Mordasini, C. 2014, Protostars and Planets VI, 691 Birnstiel, T. & Andrews, S. M. 2014, ApJ, 780, 153 Birnstiel, T., Klahr, H., & Ercolano, B. 2012, A&A, 539, 148 44 Christoph Mordasini Bitsch, B., Johansen, A., Lambrechts, M., & Morbidelli, A. 2015a, A&A, 575, A28 Bitsch, B., Lambrechts, M., & Johansen, A. 2015b, A&A, 582, A112 Bodenheimer, P. H., Hubickyj, O., & Lissauer, J. J. 2000, Icarus, 143, 2 Bodenheimer, P. H. & Pollack, J. B. 1986, Icarus, 67, 391 Bonfils, X., Delfosse, X., Udry, S., et al. 2013, A&A, 549, A109 Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 728, 117 Boss, A. P. 1995, Science, 267, 360 Boss, A. P. 1997, Science, 276, 1836 Bowler, B. P. 2016, PASP, 128, 102001 Brauer, F., Dullemond, C. P., & Henning, T. 2008, A&A, 480, 859 Bruzual, G. & Charlot, S. 2003, MNRAS, 344, 1000 Bryan, M. L., Knutson, H. A., Howard, A. W., et al. 2016, ApJ, 821, 89 Burke, C. J., Christiansen, J. L., Mullally, F., et al. 2015, ApJ, 809, 8 Burrows, A., Hubeny, I., Budaj, J., & Hubbard, W. B. 2007, ApJ, 661, 502 Cameron, A. G. W. 1978, Moon and Planets, 18, 5 Casoli, J. & Masset, F. S. 2009, ApJ, 703, 845 Cassan, A., Kubas, D., Beaulieu, J., & al. 2012, Nature, 481, 167 Chabrier, G. 2003, The Publications of the Astronomical Society of the Pacific, 115, 763 Chabrier, G., Baraffe, I., Leconte, J., Gallardo, J., & Barman, T. S. 2009, Cool Stars, 1094, 102 Chambers, J. E. 1999, MNRAS, 304, 793 Chambers, J. E. 2009, ApJ, 705, 1206 Chiang, E. & Laughlin, G. 2013, MNRAS accepted Cimerman, N. P., Kuiper, R., & Ormel, C. W. 2017, MNRAS, 471, 4662 Coleman, G. A. L. & Nelson, R. P. 2014, MNRAS, 445, 479 Coleman, G. A. L. & Nelson, R. P. 2016, MNRAS, 457, 2480 Coleman, G. A. L., Papaloizou, J. C. B., & Nelson, R. P. 2017, MNRAS, 470, 3206 Coughlin, J. L., Mullally, F., Thompson, S. E., et al. 2016, ApJS, 224, 12 Cresswell, P. & Nelson, R. P. 2008, A&A, 482, 677 Crida, A., Morbidelli, A., & Masset, F. S. 2006, Icarus, 181, 587 Cridland, A. J., Pudritz, R. E., & Alessi, M. 2016, MNRAS, 461, 3274 Cridland, A. J., Pudritz, R. E., Birnstiel, T., Cleeves, L. I., & Bergin, E. A. 2017, ArXiv e-prints Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP, 120, 531 D'Angelo, G. & Lubow, S. H. 2008, ApJ, 685, 560 David, T. J., Hillenbrand, L. A., Petigura, E. A., et al. 2016, Nature, 534, 658 Dell'Omodarme, M., Valle, G., Degl'Innocenti, S., & Prada Moroni, P. G. 2012, A&A, 540, A26 Dittkrist, K.-M., Mordasini, C., Klahr, H., Alibert, Y., & Henning, T. 2014, A&A, 567, A121 Donati, J. F., Moutou, C., Malo, L., et al. 2016, Nature, 534, 662 Dra¸zkowska, J., Alibert, Y., & Moore, B. 2016, A&A, 594, A105 Duffell, P. C., Haiman, Z., MacFadyen, A. I., D'Orazio, D. J., & Farris, B. D. 2014, ApJ, 792, L10 Durmann, C. & Kley, W. 2015, A&A, 574, A52 Fabrycky, D. C. & Tremaine, S. 2007, ApJ, 669, 1298 Figueira, P., Pont, F., Mordasini, C., et al. 2009, A&A, 493, 671 Fischer, D. A. & Valenti, J. A. 2005, ApJ, 622, 1102 Forgan, D. & Rice, K. 2013, MNRAS, 432, 3168 Forgan, D. H., Hall, C., Meru, F., & Rice, W. K. M. 2018, MNRAS, 474, 5036 Fortier, A., Alibert, Y., Carron, F., Benz, W., & Dittkrist, K.-M. 2013, A&A, 549, 44 Fouchet, L., Alibert, Y., Mordasini, C., & Benz, W. 2012, A&A, 540, 107 Freedman, R. S., Lustig-Yaeger, J., Fortney, J. J., et al. 2014, ApJS, 214, 25 Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, 81 Fulton, B. J., Petigura, E. A., Howard, A. W., et al. 2017, AJ, 154, 109 Goldreich, P. & Tremaine, S. 1979, ApJ, 233, 857 Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005, Nature, 435, 466 Gonzalez, G. 1997, MNRAS, 285, 403 Greenzweig, Y. & Lissauer, J. L. 1992, Icarus, 100, 440 Planetary population synthesis 45 Haisch, K. E., Lada, E. A., & Lada, C. J. 2001, ApJ, 553, L153 Hasegawa, Y. & Pudritz, R. E. 2011, MNRAS, 417, 1236 Hasegawa, Y. & Pudritz, R. E. 2012, ApJ, 760, 117 Hasegawa, Y. & Pudritz, R. E. 2013, ApJ, 778, 78 Hayashi, C. 1981, Progress of Theoretical Physics Supplement, 70, 35 Hellary, P. & Nelson, R. P. 2012, MNRAS, 419, 2737 Horn, B., Lyra, W., Mac Low, M.-M., & S´andor, Z. 2012, ApJ, 750, 34 Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS, 201, 15 Howard, A. W., Marcy, G. W., Johnson, J. A., et al. 2010, Science, 330, 653 Hueso, R. & Guillot, T. 2005, A&A, 442, 703 Ida, S. & Lin, D. N. C. 2004a, ApJ, 604, 388 Ida, S. & Lin, D. N. C. 2004b, ApJ, 616, 567 Ida, S. & Lin, D. N. C. 2008a, ApJ, 673, 487 Ida, S. & Lin, D. N. C. 2008b, ApJ, 685, 584 Ida, S. & Lin, D. N. C. 2010, ApJ, 719, 810 Ida, S., Lin, D. N. C., & Nagasawa, M. 2013, ApJ, 775, 42 Ida, S. & Makino, J. 1993, Icarus, 106, 210 Ikoma, M., Nakazawa, K., & Emori, H. 2000, ApJ, 537, 1013 Jin, S. & Mordasini, C. 2018, ApJ, 853, 163 Jin, S., Mordasini, C., Parmentier, V., et al. 2014, ApJ, 795, 65 Johansen, A. & Lambrechts, M. 2017, Annual Review of Earth and Planetary Sciences, 45, 359 Juri´c, M. & Tremaine, S. 2008, ApJ, 686, 603 Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108 Kley, W., Bitsch, B., & Klahr, H. 2009, A&A, 506, 971 Kley, W. & Nelson, R. P. 2012, ARA&A, 50, 211 Kokubo, E. & Ida, S. 2000, Icarus, 143, 15 Kokubo, E. & Ida, S. 2002, ApJ, 581, 666 Kokubo, E. & Ida, S. 2012, Progress of Theoretical and Experimental Physics, 2012 Kornet, K., Stepinski, T. F., & Rozyczka, M. 2001, A&A, 378, 180 Kozai, Y. 1962, AJ, 67, 591 Kretke, K. A. & Lin, D. N. C. 2012, ApJ, 755, 74 Kuiper, G. P. 1951, Proceedings of the National Academy of Science, 37, 1 Lambrechts, M. & Johansen, A. 2012, A&A, 544, 32 Lambrechts, M. & Johansen, A. 2014, A&A, 572, A107 Levison, H. F., Thommes, E. W., & Duncan, M. J. 2010, ApJ, 139, 1297 Lin, D. N. C. & Papaloizou, J. C. B. 1986a, ApJ, 307, 395 Lin, D. N. C. & Papaloizou, J. C. B. 1986b, ApJ, 309, 846 Lodders, K. 2003, ApJ, 591, 1220 Lopez, E. D. & Fortney, J. J. 2013, ApJ, 776, 2 Lubow, S. H., Seibert, M., & Artymowicz, P. 1999, ApJ, 526, 1001 Lust, R. 1952, Zeitschrift Naturforschung Teil A, 7, 87 Lynden-Bell, D. & Pringle, J. E. 1974, MNRAS, 168, 603 Lyra, W., Paardekooper, S.-J., & Mac Low, M.-M. 2010, ApJ, 715, L68 Madhusudhan, N., Amin, M. A., & Kennedy, G. M. 2014, ApJ, 794, L12 Mamajek, E. E. 2009, in American Institute of Physics Conference Series, ed. T. Usuda, M. Tamura, & M. Ishii, Vol. 1158, 3–10 Manara, C. F., Rosotti, G., Testi, L., et al. 2016, A&A, 591, L3 Mann, A. W., Newton, E. R., Rizzuto, A. C., et al. 2016, AJ, 152, 61 Marboeuf, U., Thiabaud, A., Alibert, Y., Cabral, N., & Benz, W. 2014, A&A, 570, A36 Marcy, G. W., Butler, R. P., Fischer, D. A., et al. 2005, Progress of Theoretical Physics Supplement, 158 Marcy, G. W., Weiss, L. M., Petigura, E. A., et al. 2014, Proceedings of the National Academy of Science, 111, 12655 Masset, F. & Snellgrove, M. 2001, MNRAS, 320, L55 46 Christoph Mordasini Masset, F. S. & Casoli, J. 2010, ApJ, 723, 1393 Mayor, M., Marmier, M., Lovis, C., et al. 2011, arXiv, astro-ph 1109.2497 Mayor, M. & Queloz, D. 1995, Nature, 378, 355 Mizuno, H. 1980, Progress of Theoretical Physics, 64, 544 Molli`ere, P. & Mordasini, C. 2012, A&A, 547, A105 Mordasini, C. 2014, A&A, 572, A118 Mordasini, C., Alibert, Y., & Benz, W. 2006, in Tenth Anniversary of 51 Peg-b: Status of and prospects for hot Jupiter studies, ed. L. Arnold, F. Bouchy, & C. Moutou (Paris: Frontier Group), 84–86 Mordasini, C., Alibert, Y., & Benz, W. 2009a, A&A, 501, 1139 Mordasini, C., Alibert, Y., Benz, W., Klahr, H., & Henning, T. 2012a, A&A, 541, 97 Mordasini, C., Alibert, Y., Benz, W., & Naef, D. 2009b, A&A, 501, 1161 Mordasini, C., Alibert, Y., Georgy, C., et al. 2012b, A&A, 547, 112 Mordasini, C., Alibert, Y., Klahr, H., & Henning, T. 2012c, A&A, 547, 111 Mordasini, C., Klahr, H., Alibert, Y., Miller, N., & Henning, T. 2014, A&A, 566, A141 Mordasini, C., Marleau, G.-D., & Molli`ere, P. 2017, A&A, 608, A72 Mordasini, C., Molli`ere, P., Dittkrist, K.-M., Jin, S., & Alibert, Y. 2015, International Journal of Astrobiology, 14, 201 Mordasini, C., van Boekel, R., Molli`ere, P., Henning, T., & Benneke, B. 2016, ApJ, 832, 41 Mortier, A., Santos, N. C., Sousa, S., et al. 2013, A&A, 551, A112 Moutou, C., Deleuil, M., Guillot, T., et al. 2013, Icarus, 226, 1625 Movshovitz, N., Bodenheimer, P. H., Podolak, M., & Lissauer, J. J. 2010, Icarus, 209, 616 Mulders, G. D., Pascucci, I., & Apai, D. 2015, ApJ, 814, 130 Muller, S., Helled, R., & Mayer, L. 2018, ArXiv e-prints Nakazawa, K., Ida, S., & Nakagawa, Y. 1989, A&A, 220, 293 Nayakshin, S. 2010, MNRAS, 408, L36 Nayakshin, S. & Fletcher, M. 2015, MNRAS, 452, 1654 Ndugu, N., Bitsch, B., & Jurua, E. 2018, MNRAS, 474, 886 Ogihara, M., Morbidelli, A., & Guillot, T. 2015, A&A, 584, L1 Ormel, C. W. 2014, ApJ, 789, L18 Ormel, C. W. 2017, in Astrophysics and Space Science Library, Vol. 445, Astrophysics and Space Science Library, ed. M. Pessah & O. Gressel, 197 Ormel, C. W., Dullemond, C. P., & Spaans, M. 2010, Icarus, 210, 507 Ormel, C. W., Ida, S., & Tanaka, H. 2012, ApJ, 758, 80 Ormel, C. W. & Klahr, H. H. 2010, A&A, 520, A43 Ormel, C. W., Shi, J.-M., & Kuiper, R. 2015, MNRAS, 447, 3512 Owen, J. E. & Wu, Y. 2013, ApJ, 775, 105 Owen, J. E. & Wu, Y. 2017, ApJ, 847, 29 Paardekooper, S.-J. 2014, MNRAS, 444, 2031 Paardekooper, S.-J., Baruteau, C., Crida, A., & Kley, W. 2010, MNRAS, 401, 1950 Pani´c, O., Hogerheijde, M. R., Wilner, D., & Qi, C. 2009, A&A, 501, 269 Perri, F. & Cameron, A. G. W. 1974, Icarus, 22, 416 Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Proceedings of the National Academy of Science, 110, 19273 Petigura, E. A., Marcy, G. W., Winn, J. N., et al. 2018, AJ, 155, 89 Pierens, A. 2015, MNRAS, 454, 2003 Pinilla, P., Birnstiel, T., & Walsh, C. 2015, A&A, 580, A105 Piso, A.-M. A., Oberg, K. I., Birnstiel, T., & Murray-Clay, R. A. 2015, ApJ, 815, 109 Piso, A.-M. A. & Youdin, A. N. 2014, ApJ, 786, 21 Podolak, M., Pollack, J. B., & Reynolds, R. T. 1988, Icarus, 73, 163 Pollack, J. B., Hubickyj, O., Bodenheimer, P. H., et al. 1996, Icarus, 124, 62 Reffert, S., Bergmann, C., Quirrenbach, A., Trifonov, T., & Kunstler, A. 2015, A&A, 574, A116 Safronov, V. S. 1969, Evolution of the Protoplanetary Cloud and Formation of the Earth and the Planets (Nauka, Moscow) Planetary population synthesis 47 Sahlmann, J., Segransan, D., Queloz, D., et al. 2011, A&A, 525, 95 Sallum, S., Follette, K. B., Eisner, J. A., et al. 2015, Nature, 527, 342 Salpeter, E. E. 1955, ApJ, 121, 161 S´andor, Z., Lyra, W., & Dullemond, C. P. 2011, ApJ, 728, L9 Santos, N. C., Adibekyan, V., Figueira, P., et al. 2017, A&A, 603, A30 Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153 Santos, N. C., Israelian, G., Mayor, M., et al. 2005, A&A, 437, 1127 Schneider, J., Dedieu, C., Le Sidaner, P., Savalle, R., & Zolotukhin, I. 2011, A&A, 532, A79 Shakura, N. I. & Sunyaev, R. A. 1973, A&A, 24, 337 Shu, F. H. 1977, ApJ, 214, 488 Shu, F. H., Tremaine, S., Adams, F. C., & Ruden, S. P. 1990, ApJ, 358, 495 Tanaka, H., Takeuchi, T., & Ward, W. R. 2002, ApJ, 565, 1257 Thommes, E. W., Duncan, M. J., & Levison, H. F. 2003, Icarus, 161, 431 Thommes, E. W., Matsumura, S., & Rasio, F. A. 2008, Science, 321, 814 Udry, S., Mayor, M., & Santos, N. C. 2003, A&A, 407, 369 Udry, S. & Santos, N. C. 2007, Annual Review of A&A, 45, 397 Van Eylen, V., Agentoft, C., Lundkvist, M. S., et al. 2017, ArXiv e-prints Veras, D. & Armitage, P. J. 2004, MNRAS, 347, 613 Visser, R. G. & Ormel, C. W. 2016, A&A, 586, A66 Walsh, K. J., Morbidelli, A., Raymond, S. N., O'Brien, D. P., & Mandell, A. M. 2011, Nature, 475, 206 Ward, W. R. 1986, Icarus, 67, 164 Weidenschilling, S. 1977, Astrophysics and Space Science, 51, 153 Winn, J. N. & Fabrycky, D. C. 2015, ARA&A, 53, 409 Wolfgang, A., Rogers, L. A., & Ford, E. B. 2016, ApJ, 825, 19 Wright, J. T., Fakhouri, O., Marcy, G. W., et al. 2011, PASP, 123, 412 Yu, L., Donati, J.-F., H´ebrard, E. M., et al. 2017, MNRAS, 467, 1342
1002.2109
1
1002
2010-02-10T15:15:19
Giant planet formation in the framework of the core instability model
[ "astro-ph.EP" ]
In this Thesis I studied the formation of the four giant planets of the Solar System in the framework of the nucleated instability hypothesis. The model considers that solids and gas accretion are coupled in an interactive fashion, taking into account detailed constitutive physics for the envelope. The accretion rate of the core corresponds to the oligarchic growth regime. I also considered that accreted planetesimals follow a size distribution. One of the main results of this Thesis is that I was able to compute the formation of Jupiter, Saturn, Uranus and Neptune in less than 10 million years, which is considered to be the protoplanetary disk mean lifetime.
astro-ph.EP
astro-ph
Universidad Nacional de La Plata Facultad de Ciencias Astronómicas y Geofísicas Tesis para obtener el grado académico de Doctor en Astronomía Formación de planetas gigantes en el marco del modelo de inestabilidad nucleada Lic. Andrea Fortier Director: Dr. Omar G. Benvenuto Co-Director: Dr. Adrián Brunini Marzo de 2009 Agradecimientos Esta Tesis Doctoral fue posible gracias al apoyo que recibí de muchísima gente, que durante estos cinco años estuvieron presentes en mi vida compartiendo alegrías y desen- cuentros, e impulsándome a seguir adelante. Hoy, el resultado se traduce en este manuscrito. Sin embargo, entre las citas que encontrarán en él generalmente no aparecen mencionadas las personas que, directa o indirectamente, colaboraron para la realización de este traba jo. Es por ésto que considero importante mencionarlas en este apartado. Y si bien segura- mente no podré hacer justicia al intentar transmitir lo que ellos significaron para mí en este tiempo, espero al menos poder dejar en claro que una parte muy importante de este traba jo se las debo a todos ellos. Por empezar, mi más profundo agradecimiento a mis directores, Omar Benvenuto y Adrián Brunini, por haberme dejado participar de este proyecto, ambicioso quizá en un principio, de intentar entender cómo se formaron los planetas gigantes. Me resulta increíble al mirar hacia atrás, cuando en los comienzos del año 2004, empecé a intentar “empapar- me” en el tema, luchando contra un código numérico que nunca hacía lo que yo pretendía, y encontrar que hoy podemos presentar los resultados que conforman esta Tesis. Fueron Omar y Adrián quienes estuvieron presentes siempre, guiándome, enseñándome y, funda- mentalmente, apoyándome para que el traba jo pudiera salir adelante. Pero además, junto a ellos aprendí otras cosas, que se dieron con el tiempo, a partir de charlas y discusiones, y que quizá sean tan importantes como estos resultados. Fundamentalmente, puedo decir que gracias a ellos pude comenzar a formarme un criterio. Un criterio para distinguir las cosas importantes de las que no lo son tanto, y que no solo se aplica al armado de un modelo para estudiar un problema particular, sino que se extiende a las relaciones humanas y a la vida en general. Por todo lo que compartimos y por todo lo que aprendí estando al lado de ellos, gracias. Agradezco la Universidad Nacional de La Plata y, muy en particular, al personal docente y no docente de la Facultad de Ciencias Astronómicas y Geofísicas por la ayuda brindada en este tiempo. También, le agradezco al Instituto de Astrofísica de La Plata (IALP) y al Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET). A la Facultad de Ingeniería y, en particular, al Departamento de Ciencias Básicas les agradezco por permitir mi desarrollo docente y por el reconocimiento que en el transcurso de estos años me brindaron. Mi especial consideración para la Lic. Lilina Carboni (Jefa del Departamento de Ciencias Básicas), la Prof. Nélida Echebest (Coordinadora de la Cátedra i i i de Matemática C) y para mis compañeros de cátedra. Mi agradecimiento también al Prof. Raúl Rossignoli por los años compartidos y la confianza que siempre tuvo en mi persona. A mis compañeros de oficina, Alejandra De Vito y Rubén Martinez, todo mi cariño por haberme sabido acompañar, escuchar y ayudar. Los tres, junto también a Omar Benvenuto, hemos compartido muchas horas de nuestras vidas, que quedarán por siempre en mis recuerdos. Gracias por, mate mediante, haberme permitido conocerlos y crecer a su lado. A los miembros del Grupo de Ciencias Planetarias: Romina Di Sisto, Lorena Dirani, Pablo Santamaría y Yamila Miguel por muchos años más compartiendo inquietudes, al- muerzos y discusiones. Gracias por saber que puedo contar siempre con ustedes. A Héctor Viturro por el soporte técnico. Y a mi gran amigo, Gonzalo de Elía, compañero en las buenas y en las malas, mi “mellizo” de carrera, gracias, muchas, muchas gracias por todo. Este trayecto de mi vida podría haber sido muy distinto sin su amistad. Gracias, Gonza, por haber estado a mi lado. En estos años tuve la suerte de asistir a reuniones científicas las cuales, no solo con- tribuyeron a mi formación profesional, sino que me permitieron conocer gente que, con su ejemplo, me impulsó a seguir en esta carrera. Por el aliento y por las enseñanzas que, probablemente sin proponérselo, me dieron quiero agradecer a: Prof. J. Fernandez, Prof. G. Wuchterl, Prof. M. Kürster, y Prof. S. Ida. Pero, sin duda, lo más provechoso de es- tas reuniones fue conocer a muchos de mis pares, algunos de los cuales se convirtieron en grandes amigos a pesar de las distancias. A Nancy Sosa, Pablo Pais, Christoph Mordasini, Arnaud Cassan, Andrés Carmona, Olga Zakhozhay, Marin Treselj y Francesca De Meo, gracias. Por el compañerismo, por el haber compartido tantas cosas juntos, por el deseo de que esto siga siendo así , gracias a: Javier Vasquez, Nicolás Duronea, Andrea Torres, Marcelo Miller, Ileana Andruchow, Diego Bagú, Jorge Panei, Leandro Althaus, Fernando de la Corte, Ricardo Amorín, Guillermo Hägele, Germán Rubino, Juan Ignacio Sabbione, Luis Guarracino, Sofía Cora y Sixto Gimenez-Benitez. A Mariana Orellana por su amistad. Durante este tiempo, siete amigas pudimos construir una relación que me llena de orgullo y que representa una de las cosas más valiosas que tengo. A Verónica Firpo, Cecilia Fariña, Anahí Granada, Anabella Araudo, Claudia Scoccola y Antonela Monachesi gracias por la amistad que cultivamos. Gracias por ser como son. A Melina Bersten, gracias por todas las cosas compartidas. A pesar de la distancia, nuestra amistad dio sobradas pruebas de ser auténtica y robusta; espero saber honrarla. A Juan Cappi y Lorena Praderio, por todas las cenas compartidas y por todas las que vendrán. Gracias por formar parte de mi vida y por permitirme formar parte de la suya. Y por último, a los verdaderos responsables de que esta Tesis se haya podido escribir: mi mamá Beatriz, mi papá Eduardo, mi hermano Andrés, mi hermana “postiza” Florencia y a mi abuela Claire, gracias. Por el amor, la paciencia y el apoyo, gracias. Por todos los momentos que compartimos juntos, buenos y malos, gracias. Por su presencia incondicional, gracias. Por aceptarme y quererme como soy ... no tengo palabras. iv A mi mamá, a mi papá y a mi hermano. vi Resumen Los planetas del Sistema Solar han captado la atención de la humanidad por siglos. Sin embargo, a pesar de toda la información con la que se cuenta, ya sea teórica o proveniente de datos observacionales, son todavía muchas las incógnitas que involucran la evolución de los sistemas planetarios y de los diversos cuerpos que los componen. El objetivo de esta tesis es estudiar la formación de los objetos de masa subestelar más grandes que pueblan los sistemas planetarios: los planetas gigantes. Existen dos modelos, cualitativamente muy diferentes, que explican el proceso de for- mación de los planetas gigantes: la hipótesis de inestabilidad gravitatoria del disco y la hipótesis de inestabilidad nucleada. Si bien no se puede descartar la plausibilidad de la pri- mera como escenario de formación, es la hipótesis de inestabilidad nucleada la que cuenta con mayor consenso en la comunidad científica debido a que consigue explicar en forma sencilla la actualmente aceptada estructura interna de los planetas gigantes. La hipótesis de inestabilidad nucleada propone dos etapas para la formación de los planetas gigantes: en una primera instancia, por acreción de planetesimales, se forma un núcleo sólido. Cuando éste llega a aproximadamente unas 10 masas terrestres, comienza la segunda etapa, durante la cual el planeta alcanza su masa final a instancias de acretar grandes cantidades de gas en una corta escala de tiempo. El presente traba jo fue desarrollado adoptando la hipótesis de inestabilidad nucleada como modelo de formación. Por su parte, el crecimiento de los embriones sólidos no se produce a tasa constante sino que atraviesa también varios estadíos. Al principio los planetesimales siguen un crecimiento retroalimentado conocido como crecimiento runaway, cuyo producto es la aparición de los primeros embriones planetarios en tan solo unas decenas de miles de años. Cuando estos embriones alcanzan aproximadamente la masa de la Luna perturban gravitatoriamente a los planetesimales vecinos, aumentando las velocidades relativas y disminuyendo así su tasa de crecimiento. Este régimen se conoce como crecimiento oligárquico ya que solo los embriones más grandes son capaces de continuar capturando planetesimales. La tasa de acreción de sólidos puede ser crucial para estimar el tiempo de formación total del planeta. Los datos observacionales provenientes de discos circumestelares limitan la escala de tiempo de formación de los planetas gaseosos, la cual no puede superar los 10 millones de años. Esta es una de las mayores dificultades asociadas a la hipótesis de inestabilidad nucleada, ya que en general los resultados de las simulaciones no consiguen encuadrarse satisfactoriamente dentro de esta cota. Es por esto que la mayoría de los traba jos que se encuentran en la vi i bibliografía adoptan el crecimiento runaway como único régimen de crecimiento para el núcleo durante toda la formación planetaria. Esta Tesis compila los resultados de nuestro estudio de la formación de planetas gi- gantes. El traba jo se realizó ba jo la suposición de que el planeta se encuentra en órbita circular, no perturbada, alrededor del Sol, donde no se han tenido en cuenta la existencia de campos magnéticos, turbulencias o rotación del planeta. Los cálculos fueron realizados con un código de tipo Henyey, el cual resuelve las ecuaciones diferenciales de evolución estelar acopladas en forma autoconsistente a la tasa de de acreción de planetesimales. Se consideró, por primera vez en este tipo de cálculos, al crecimiento oligárquico como el ré- gimen dominante para el crecimiento del núcleo durante toda la formación del planeta, incorporando así un tratamiento más realista para la acreción del material sólido. Como consecuencia de esto, el tiempo de formación resulta mucho más prolongado que lo encon- trado en traba jos previos. Por otra parte, sin embargo, hemos observado una dependencia muy fuerte entre el tiempo de formación y el tamaño de los planetesimales acretados: cuan- to más pequeños son, más rápido se completa el proceso. Consecuentemente, dado que los planetesimales que pueblan el disco no son todos iguales, hemos incluido en el cálculo una función que represente la distribución de tamaños. Las simulaciones realizadas contemplan una población con hasta nueve tamaños, entre los 100 metros y 100 kilómetros de radio, siendo los más pequeños los más abundantes. Ba jo estas hipótesis hemos conseguido mo- delar la formación de Júpiter, Saturno, Urano y Neptuno en menos de 10 millones de años. Además, las masas obtenidas para los núcleos se encuentran dentro de las cotas aceptadas actualmente. Otro de los resultados relevantes de esta tesis, fue la aparición, ba jo ciertas circuns- tancias, de eventos cuasi-periódicos de pérdida de gas cuando el planeta llega a ciertos valores críticos de su masa. Este efecto no había sido observado previamente, y podría tener consecuencias determinantes en el proceso de formación de los planetas gigantes. vi i i Índice general 1. Introducción 1.1. Un poco de historia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. El Sistema Solar: generalidades . . . . . . . . . . . . . . . . . . . . . . . . 1.3. Clasificación de los objetos subestelares . . . . . . . . . . . . . . . . . . . . 1.3.1. Enanas marrones . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.2. Planetas gigantes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.3. Planetas terrestres . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.4. Planetas enanos y cuerpos pequeños 1.4. ¿Por qué estudiar la formación de planetas gigantes? . . . . . . . . . . . . 2. Datos observacionales 2.1. Características de los planetas gigantes del Sistema Solar . . . . . . . . . . 2.2. Planetas extrasolares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1. Métodos de detección . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2. Programas espaciales . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.3. Características de los planetas extrasolares detectados . . . . . . . . 2.3. Teorías de migración . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3. Escenarios de formación 3.1. Formación de un sistema planetario . . . . . . . . . . . . . . . . . . . . . . 3.1.1. Nubes moleculares gigantes: la nursery . . . . . . . . . . . . . . . . 3.1.2. Formación del disco protoplanetario . . . . . . . . . . . . . . . . . . 3.1.3. Del polvo a los planetesimales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.4. Planetas rocosos y planetas gaseosos 3.2. Teorías de formación de planetas gigantes . . . . . . . . . . . . . . . . . . . 1 1 2 5 7 9 10 10 11 13 13 20 20 25 27 29 33 33 33 34 35 38 40 ix 3.2.1. Modelo de inestabilidad nucleada . . . . . . . . . . . . . . . . . . . 3.2.2. Modelo de inestabilidad del disco . . . . . . . . . . . . . . . . . . . 3.2.3. ¿Un modelo híbrido? . . . . . . . . . . . . . . . . . . . . . . . . . . 4. Modelo teórico de formación de planetas gigantes 4.1. El disco protoplanetario . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Sobre el crecimiento de los planetesimales . . . . . . . . . . . . . . . . . . . 4.2.1. Régimen de crecimiento runaway . . . . . . . . . . . . . . . . . . . 4.2.2. Régimen de crecimiento oligárquico . . . . . . . . . . . . . . . . . . 4.3. La aparición de los embriones planetarios . . . . . . . . . . . . . . . . . . . 4.4. La envoltura gaseosa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1. Ecuaciones constitutivas . . . . . . . . . . . . . . . . . . . . . . . . 4.4.2. La ecuación de estado del gas . . . . . . . . . . . . . . . . . . . . . 4.4.3. Opacidades . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.4. Condiciones de borde . . . . . . . . . . . . . . . . . . . . . . . . . . Interacción entre las capas de gas y los planetesimales ingresantes . 4.4.5. 5. Tratamiento numérico del modelo de formación de planetas gigantes 5.1. Llevando las ecuaciones diferenciales a ecuaciones en diferencias . . . . . . 5.2. Resolución de las ecuaciones . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Tratamiento de la acreción de sólidos . . . . . . . . . . . . . . . . . . . . . 5.4. Cálculo de la acreción de gas . . . . . . . . . . . . . . . . . . . . . . . . . . 41 43 45 47 48 49 52 53 56 61 61 64 70 71 72 77 77 82 85 90 6. Resultados 93 6.1. Simulaciones considerando una población de planetesimales de tamaño único 94 6.1.1. Ejemplo y análisis de la formación de un planeta gigante . . . . . . 96 6.1.2. Variación de los resultados frente al cambio de las condiciones de borde externas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 6.2. Simulaciones considerando una población de planetesimales que sigue una distribución de tamaños . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 6.2.1. El modelo de Niza . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6.2.2. Resultados . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 6.3. Resultados inesperados: eventos cuasi periódicos de variación de la masa de la envoltura . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 x 7. Discusión 8. Conclusiones y perspectivas Apéndices Símbolos y unidades 129 141 151 151 xi xi i Capítulo 1 Introducción 1.1. Un poco de historia1 El concepto de “Sistema Solar” surge después de la aceptación de la teoría heliocéntrica, quedando definitivamente incorporado en la comunidad científica hacia fines del siglo XVII. El modelo actual de formación del Sistema Solar (la hipótesis nebular) se remonta al siglo XVIII y tiene sus orígenes en las ideas de E. Swedenborg, I. Kant y P. S. Laplace. Según esta teoría, el Sol y los planetas se habrían formado a partir de una nube molecular rotante en contracción. Asociado a este proceso, tendría lugar la formación de un disco alrededor de la estrella, donde a su vez se originarían los planetas y demás cuerpos menores. Este escenario explica naturalmente la coplanaridad y la cuasi circularidad de las órbitas de los planetas del Sistema Solar. Sin embargo, en el pasado esta teoría fue muy cuestionada debido a la distribución actual de momento angular entre el Sol y los planetas: mientras que el 99,86 % de la masa del Sistema Solar está en el Sol, el 99,5 % del momento angular total del sistema reside en los planetas. De la aplicación directa de la teoría de Swedenborg, Kant y Laplace, se desprende que la mayor parte de la masa y del momento angular del sistema deberían haber permanecido en el Sol. Este hecho hizo que se desestimara a la hipótesis nebular por casi dos siglos. Sin embargo, desde comienzos de 1980 las observaciones muestran que la mayoría de las estrellas jóvenes están rodeadas por un disco de gas y polvo, lo cual concuerda con las predicciones de la hipótesis nebular. Actualmente, la hipótesis nebular se encuentra completamente aceptada como el escenario estándar de formación del Sistema Solar. Por su parte, Descartes, en 1644, propuso que los planetas eran el resultado último de un sistema de vórtices girando alrededor del Sol. Esta idea no incluía la presencia de un disco protoplanetario ni tampoco sugería la manera en que llegaron a formarse dichos vórtices. Tres siglos más tarde, en 1944, C.F. von Weizsäcker retomó esta idea proponiendo que entorno al Sol existió un disco turbulento, de composición solar, en el cual se generaron 1Estas notas están basadas, fundamentalmente, en el artículo Historical notes on planet formation (Bodenheimer 2006). 1 vórtices que permitieron la acreción de pequeñas partículas, lo cual finalmente habría dado lugar a los planetas. El aporte más importante de von Weizsäcker fue mostrar que en un disco turbulento el momento angular se transporta como consecuencia de la viscosidad, donde durante el proceso de acreción la masa fluye hacia el centro y el momento angular es transportado hacia afuera. Esta idea permitía resolver uno de los mayores problemas de la hipótesis nebular, según la cual el Sol debería rotar mucho más rápido de lo que lo hace en realidad. La propuesta de la formación de vórtices fue bastante criticada, sobre todo porque los vórtices tendrían una vida media muy corta, con lo cual no fue tomada en serio hasta 1995 cuando Barge & Sommeria mostraron que las partículas más pequeñas resultaban fácilmente capturadas por estos vórtices, permitiendo la formación de los núcleos planetarios en alrededor de 105 años, acelerando así el proceso de formación y dando, de este modo, una posible solución al problema de las escalas de tiempo involucradas en la formación de los planetas gigantes (ver capítulo 6). Por su parte, Hoyle (1960) invocó la ruptura magnética para explicar la “lenta” rotación del Sol. Según sus cálculos, los efectos hidrodinámicos por sí mismos (como, por ejemplo, la viscosidad) no serían suficientes para explicar la transferencia de momento angular ya que, de haber sido así, el material del disco debería haber estado en contacto con el Sol y, en ese caso, la velocidad de rotación del Sol habría disminuido solo hasta encontrarse en co- rotación con la parte interior del disco. Hoyle propuso entonces que durante el proceso de formación del sistema Sol-disco se debió haber abierto una brecha entre ambos, pronunciada aún más por el campo magnético, lo cual permitió la transferencia de momento angular desde el Sol hacia la parte interior del disco, y desde allí hacia afuera. Entonces, durante su contracción, el Sol habría acelerado su velocidad de rotación pero, a su vez, el momento angular habría continuado siendo transferido a través del campo magnético. Para obtener el acoplamiento magnético, la temperatura debió ser del orden de los 1.000 K, y la intensidad del campo magnético de alrededor de 1 gauss. Más allá del borde interno del disco no se requiere de este acoplamiento ya que la viscosidad del disco permitiría la transferencia de momento angular hacia la región exterior del mismo. Hoy en día la explicación básica para la pérdida de momento angular por parte de la estrella central radica en el transporte magnético y la turbulencia. Sin embargo, los cálculos actuales estiman que los campos magnéticos involucrados deberían ser del orden de los 1000 gauss. 1.2. El Sistema Solar: generalidades La estructura básica del Sistema Solar involucra la presencia de objetos muy diversos, tanto en tamaño como en composición. El Sistema Solar está dominado por el Sol, una estrella enana de tipo G2V según su clasificación espectral, estable, que se encuentra atra- vesando la mitad de su existencia y que es la principal responsable de que en La Tierra se desarrolle la vida. La composición química del Sol es: un 70 % de hidrógeno, un 28 % de helio y un 2 % de elementos pesados. El Sol produce su energía por fusión nuclear en su 2 Tabla 1.1. Características principales de los planetas del Sistema Solara . Planeta Mercurio Venus Tierra Marte Júpiter Saturno Urano Neptuno a [UA] 0,39 0,72 1,00 1,52 5,20 9,54 19,19 30,07 Período orbital [años] 0,24 0,61 1,00 1,88 11,86 29,46 84,07 164,8 e 0,206 0,007 0,017 0,093 0,048 0,054 0,047 0,009 i Masa Radio Densidad Media [M⊕ ] [R⊕ ] [g cm−3 ] 5,43 0,38 0,055 5,24 0,35 0,817 5,52 1,00 1,000 0,107 0,53 3,90 1,30 11,2 317,8 0,70 9,41 94,30 14,60 4,11 1,30 1,50 3,81 17,20 7,0◦ 3,4◦ 0,0◦ 1,8◦ 1,3◦ 2,5◦ 0,8◦ 1,8◦ a El semieje orbital a está dado en Unidades Astronómicas (UA), la masa del planeta en masas de la Tierra (M⊕ ) y el radio del planeta en radios de la Tierra (R⊕ ). La excentricidad se designa con e, y la inclinación del plano orbital respecto de la eclíptica con i. centro, donde los átomos de hidrógeno se combinan para dar átomos de helio. La tempe- ratura central del Sol, donde ocurre la fusión nuclear, es de 15, 7 × 106 K, mientras que la temperatura promedio en la fotosfera es de 5800 K. Entorno al Sol orbitan numerosos cuerpos, siendo los planetas los más destacados. Los planetas orbitan alrededor del Sol en planos muy cercanos a la eclíptica, todos en la misma dirección, que corresponde al sentido contrario al de las agujas del relo j visto desde el polo norte ecliptical. La arquitectura de nuestro sistema planetario puede resumirse, según la distancia orbital creciente en: cuatro planetas terrestres (Mercurio, Venus, La Tierra y Marte), el Cinturón de Asteroides, dos planetas gigantes gaseosos (Júpiter y Saturno), dos planetas gigantes de hielo (Urano y Neptuno), el Cinturón de Kuiper (compuesto por asteroides y cometas) y, finalmente, en los confines del Sistema Solar, la Nube de Oort. En la tabla 1.1 se listan las características principales de los planetas del Sistema Solar. La región comprendida entre Marte y Júpiter está ocupada por numerosos cuerpos irregulares y se la conoce como Cinturón de Asteroides. Más de la mitad de su masa está contenida en los cuatro objetos más grandes: Ceres (planeta enano), Vesta, Pallas e Hygiea. El resto de la población se distribuye entre asteroides más pequeños y partículas de polvo. Los asteroides son planetesimales2 remanentes de la formación del Sistema Solar que no pudieron ser acretados para formar un planeta debido a las perturbaciones gravitatorias a las que se vieron sometidos por parte de los planetas gigantes. Las colisiones entre es- 2La palabra planetesimal proviene del concepto matemático de infinitesimal y literalmente significa la fracción más pequeña que compone a un planeta. La denominación “planetesimal” se utiliza para referirse a los cuerpos rocosos más pequeños presentes en el Sistema Solar durante el proceso de formación planetaria. 3 tos cuerpos habrían sido muy violentas, siendo mayormente destructivas, lo cual habría impedido la acreción en un cuerpo más masivo. El Cinturón de Kuiper está formado por asteroides y cometas, y es similar al Cinturón de Asteroides pero mucho más grande, tanto en tamaño como en masa. Esta región del Sistema Solar, que se extiende desde la órbita de Neptuno (∼ 30 Unidades Astronómicas - UA) hasta aproximadamente unas 55 UA, está compuesta por los remanentes de la for- mación del Sistema Solar. A diferencia de los miembros del Cinturón de Asteroides, los cuerpos del Cinturón de Kuiper están compuestos mayormente por volátiles helados, como metano, amoníaco y agua. El cuerpo más grande del Cinturón de Kuiper es Eris. El primer objeto, descontando a Plutón, que confirmó la existencia del Cinturón de Kuiper se des- cubrió en 1992 (Jewitt & Luu, 1992), aunque su existencia fue propuesta durante décadas por varios investigadores. Actualmente se conocen más de mil objetos pertenecientes a esta región, y se estima que debe haber más de 70.000 con un diámetro mayor a los 100 km. En el Cinturón de Kuiper se distinguen cuatro poblaciones dinámicamente diferentes. El Cin- turón de Kuiper clásico está formado por cometas de ba jas inclinaciones y excentricidades (40 < a < 48 UA, e < 0, 2). Una segunda población de objetos con órbitas excéntricas y más inclinadas, generalmente de mayor semieje, forman el scattered disk (o “disco disper- so”), reservorio de los cometas de corto período. Estos objetos tienen perihelios cercanos a la órbita de Neptuno (30 < q < 39 UA, e > 0, 2), e interactúan gravitatoriamente con él. Se presume que Tritón, el satélite más grande de Neptuno, sería un objeto capturado del Cinturón de Kuiper. Otra región es el Scattered Disk extendido, con q > 39 UA y a > 50 UA. Por último se distingue la Población Resonante, formada por objetos que tienen alguna relación de conmensurabilidad con Neptuno, siendo la resonancia 3:2 la más poblada. Debido a la controversia que genera la denominación “Cinturón de Kuiper”, y las diferentes regiones que se pueden distinguir más allá de la órbita de Neptuno, actualmente los astrónomos se refieren a estos objetos en forma general como objetos transneptunianos (TNOs). La masa del Cinturón de Kuiper contenida hasta 50 UA sería de aproximadamente 0, 5 M⊕ . En función del número de cometas observados, el número de cometas del Cinturón de Kuiper entre 30 y 50 UA de diámetro mayor que 1 km está estimado en 109 . Del mismo orden sería el número de cometas que conforman el Scattered Disk. Los cometas son planetesimales helados de la región externa de la nebulosa solar que orbitan más allá de Júpiter. Estos proto-cometas que originalmente tenían órbitas entre los planetas gigantes fueron eyectados por sus interacciones gravitatorias con ellos, la mayoría hacia el espacio interestelar. Sin embargo, una fracción de ellos quedó ligado gravitatoria- mente al Sol, formando una esfera de aproximadamente 1 pc de radio. Las perturbaciones estelares y galácticas aumentaron los perihelios de los cometas llevándolos fuera de la re- gión planetaria, generando una distribución aleatoria de sus órbitas lo cual le confirió una configuración esférica. La Nube de Oort sería entonces una aglomeración más o menos esférica de cometas que se encuentra a aproximadamente 50.000 UA del Sol. Su límite ex- terior define los confines de la influencia gravitatoria del Sistema Solar. La Nube de Oort se podría pensar como constituida por dos estructuras: una región interna y una externa, esta 4 última de forma esférica. Los miembros de la Nube de Oort serían objetos que se formaron en el interior del Sistema Solar y que, debido a interacciones gravitatorias con los planetas gigantes, fueron dispersados hacia el exterior. Se cree que los cometas de largo período, y los de tipo Halley, provienen de la Nube de Oort. Hasta la fecha no se ha confirmado la observación de objetos pertenecientes a dicha región, aunque Sedna y otros dos objetos transneptunianos son fuertes candidatos a ser miembros de la Nube de Oort interior. Se estima que en la Nube de Oort hay alrededor de 1012 cometas, con una masa total de algunas masas terrestres (∼ 10 M⊕). Entre el 50-80 % de su población se encuentra en una región más densa, a unas 10.000 UA del Sol. 1.3. Clasificación de los objetos subestelares Los objetos subestelares son cuerpos celestes cuya masa es inferior a la necesaria para que la fusión de hidrógeno se produzca o sea sostenida durante escalas de tiempo signifi- cativamente largas. La masa mínima necesaria estimada para que la fusión de hidrógeno sea posible estaría en el rango 0,07 - 0,092 M⊙ , dependiendo de la composición química de la estrella (0,07 M⊙ para objetos de metalicidad solar y 0,092 M⊙ para metalicidad cero). Los objetos subestelares pueden, a su vez, subclasificarse en enanas marrones, planetas, cuerpos de masa “planetaria”, satélites y todo tipo de cuerpos pequeños (como asteroides y cometas). En los últimos 20 años, con el progreso en las capacidades tecnológicas y en las técnicas observacionales, se han encontrado numerosos objetos de masa planetaria alrededor de otras estrellas (los llamados planetas extrasolares). Pero además se descubrieron nuevos cuerpos orbitando alrededor del Sol, fundamentalmente en la región del cinturón de Kuiper, algunos de los cuales tienen dimensiones comparables a la de Plutón. Todos estos descubrimientos pusieron ba jo la lupa el concepto mismo de planeta. Por este motivo, la Unión Astronómica Internacional (IAU por sus siglas en inglés) designó un grupo de traba jo que tuvo como objetivo redefinir este término. Finalmente, luego de acalorados debates, la resolución B5 de la IAU, tomada en la Asamblea General celebrada en Praga el 24 de Agosto de 2006, define a un planeta como un cuerpo celeste que: (a) está en órbita alrededor del Sol, (b) tiene la masa suficiente como para que su autogravedad contrarreste las fuerzas de cuerpo rígido, de modo que su forma sea aquella correspondiente a la de equilibrio hidrostático (cuasi esférica), y (c) que haya “limpiado” la vecindad de su órbita. El punto (c) hace referencia a que el objeto en cuestión sea gravitatoriamente dominante en la región que comprende su órbita y adyacencias, con lo cual no debe haber en sus cercanías ningún otro objeto de tamaño comparable, exceptuando sus lunas si las tuviera. Como se desprende inmediatamente de la definición, la palabra planeta, estrictamente hablando, puede ser usada solo para hacer referencia a objetos del Sistema Solar. De 5 hecho, los únicos planetas son, por ahora, Mercurio, Venus, La Tierra, Marte, Júpiter, Saturno, Urano y Neptuno. Una definición más abarcativa, que tenga en cuenta a objetos extrasolares, sea que estos estén en órbita entorno a una estrella o no, está actualmente siendo estudiada. De todos modos, y por las reminiscencias históricas del término, a los cuerpos de masa planetaria que orbitan una estrella o una enana marrón se los sigue llamando planetas. Los planetas extrasolares, o exoplanetas, son planetas que no pertenecen al Sistema Solar y que orbitan alrededor de una estrella. La mayoría de los detectados hasta el mo- mento son planetas gigantes de características jovianas, aunque seguramente la detección de planetas de tipo terrestres será corriente en los próximos años, tal como lo predice la teoría (Ida & Lin 2008a, 2008b). En última instancia, lo que se persigue con esta intensa exploración de sistemas extrasolares es la búsqueda de vida. Claro que, probablemente, la detección de vida tal como la conocemos en nuestro planeta todavía esté muy lejos de ocu- rrir. Sin embargo, como veremos en el capítulo 2, en las últimas dos décadas la detección de exoplanetas ha aumentado de una manera extraordinaria, al punto que las teorías de formación y evolución planetaria quedaron varios pasos por detrás de las observaciones. Pero no solo se han detectado planetas entorno a estrellas y enanas marrones, sino también “planetas sueltos” en el espacio (ver, por ejemplo, Zapatero Osorio et al. 2002). A estos objetos de masa planetaria se los denomina planemos, rogue planets o free floa- ting planets (“planetas flotando a la deriva”). El descubrimiento de estos objetos que no están asociados a ninguna estrella (en general pertenecen a cúmulos abiertos muy jóvenes (Caballero et al. 2006)) fue otro hito en las Ciencias Planetarias, ya que despertó nuevos interrogantes, tanto sobre la formación estelar como sobre la formación de sistemas pla- netarios. Si estos “planetas” en sus inicios orbitaban una estrella y por algún encuentro fueron separados de ella, o si pudieron formarse en soledad en alguna región de formación estelar, es todavía una incógnita. Si bien la definición de “planeta extrasolar” no ha sido oficialmente establecida todavía, el grupo de traba jo de la IAU traba jando en este tema propone que: Objetos con masas deba jo de la fusión termonuclear del deuterio (alrededor de 13 masas de Júpiter para la metalicidad solar), que orbitan estrellas o remanentes estela- res son considerados “planetas”, sin importar cómo se formaron. El límite inferior de masa requerido para un planeta extrasolar es el mismo que el utilizado en el Sistema Solar. Objetos subestelares con masas superiores por encima del límite correspondiente a la fusión del hidrógeno son “enanas marrones”, sin importar su mecanismo de formación. Free-floating objects en cúmulos estelares con masas por deba jo de la necesaria para la fusión del deuterio no son “planetas” sino “sub-enanas marrones”. Su existencia es de particular importancia para determinar los mecanismos de formación planeta- ria. Podrían ser planetas que fueron arrancados de sus órbitas por perturbaciones 6 gravitatorias. Sin embargo, en la región de Orión no hay suficientes estrellas que pudieran albergar sistemas planetarios. Luego, estos planetas podrían provenir del colapso gravitatorio por fragmentación directa de la nube. 1.3.1. Enanas marrones Con el término de enana marrón se denomina a los objetos subestelares que no tienen la masa necesaria para mantener la fusión de hidrógeno en su núcleo3 pero que, sin embargo, pueden quemar deuterio. Las enanas marrones son completamente convectivas, tanto en su interior como en su superficie. El posible rango de valores para su masa está entre aquellas correspondientes a la de los planetas gigantes y a la de las estrellas más pequeñas de la Secuencia Principal, esto es, entre 13 y 80 masas de Júpiter (MJ ). Es en este intervalo de masas donde se alcanzaría la temperatura de 0, 5 × 106 K necesaria para la fusión de deuterio. Las enanas marrones de mayor masa (0,06 - 0,08 M⊙ ) pueden también quemar litio transitoriamente durante algunas etapas de su evolución ya que la temperatura en su interior podría llegar a los 2, 5 × 106 K. La presencia de litio en un espectro es lo que permite la identificación de las enanas marrones y su diferenciación de las estrellas de muy ba ja masa. Las estrellas queman su contenido de litio muy rápidamente, no dejando rastros de su presencia, mientras que en los objetos subestelares el litio no llega a ser completamente consumido. Entonces, la diferencia más importante entre las estrellas y las enanas marrones es que estas últimas no alcanzan la secuencia principal correspondiente a la quema de hidrógeno. Así, el proceso evolutivo de una enana marrón es simplemente su enfriamiento. En este aspecto son más cercanas a los planetas gaseosos, siendo la mayor diferencia entre ambos su temperatura para una determinada edad, la cual depende de la masa del objeto. La primera enana marrón confirmada no estaba sola en el espacio sino en órbita alre- dedor de una estrella. A partir de ese momento, las enanas marrones se han encontrado en forma aislada o como miembros de sistemas múltiples, pudiéndose distinguir dos sub- tipos de sistemas: aquellos formados solo por enanas marrones y aquellos en los que la enana marrón tiene un compañero de masa estelar (Bate 2006). Este último caso (enana marrón-estrella) es de particular interés a la hora de comparar a las enanas marrones con los planetas. En el caso de sistemas binarios de enanas marrones se observan dos grandes diferencias respecto de los sistemas binarios estelares: por un lado la separación entre las enanas marrones es menor que la que ocurre en los pares estelares, pero además la relación de masa entre ambos miembros es siempre cercana a la unidad, mientras que la frecuen- cia de estrellas binarias tiene un pico cuando la relación de masas es del orden de 0,25. Por otro lado, tenemos a las enanas marrones como acompañantes de una estrella. Como mencionaremos más adelante, la mayor cantidad de exoplanetas fueron descubiertos por el método de velocidades radiales. Es esperable entonces que esta técnica de como resultado el descubrimiento de muchas enanas marrones. Sin embargo, eso no ha sido así. Muy pocas 3La temperatura necesaria para la quema de hidrógeno es aproximadamente 3 × 106 K. 7 enanas marrones fueron descubiertas en órbitas cercanas a estrellas, dando lugar a lo que se conoce como “desierto de enanas marrones”. Esta ausencia hace pensar en un mecanismo de formación diferente entre los planetas y las enanas marrones. Existen varios escenarios propuestos para la formación de las enanas marrones, dependiendo de si forman parte de un sistema (y de qué clase de sistema) o si son objetos aislados. Aquellas enanas marrones (y posiblemente también los planemos) que se encuentran aisladas han seguido, muy pro- bablemente, un proceso de formación análogo al de las estrellas de masa solar, atravesando en sus orígenes una fase T-Tauri. Como evidencia de esto se tiene generalmente la presencia de un disco circumestelar, el cual se detecta en la mayoría de las enanas marrones jóvenes, además de líneas de emisión anchas y asimétricas (las cuales indican procesos de acreción) y líneas prohibidas (Jayawardhana & Ivanov 2006). Estas analogías entre las estrellas de ba ja masa y las enanas marrones sugieren un escenario de formación común. Sin embargo, debido a su ba ja masa, las reacciones termonucleares de fusión de hidrógeno que ocurren en el núcleo no producen la luminosidad suficiente para mantener su estructura. El colapso gravitatorio de la protoestrella no calienta en forma efectiva al núcleo y antes que se alcan- ce la temperatura necesaria para el desarrollo eficiente de la fusión nuclear, la densidad se vuelve muy alta, llegando al estado de degeneración electrónica. Ba jo estas condiciones, la contracción gravitatoria se detiene y el resultado es una “estrella fallida” o enana marrón. A partir de ahí, el objeto simplemente se enfría, radiando su energía interna. Se ha sugerido que otro mecanismo posible para la formación de enanas marrones podría tener lugar durante el proceso de formación estelar en regiones donde hay varios objetos en formación simultánea. Allí se producirían eyecciones por interacciones dinámicas y las enanas marrones serían objetos que fueron eyectados antes de llegar a alcanzar la masa necesaria para convertirse en una estrella. Este escenario fue estudiado mediante simulaciones hidrodinámicas que mostraron su factibilidad (ver Bate 2006 y las referencias que allí se mencionan). Otra posibilidad es que las enanas marrones se formen debido a una finalización abrupta de la acreción de gas debido a que la radiación de estrellas masivas cercanas produciría la evaporación del resto de gas necesario. Si bien este escenario es factible, no podría ser el único mecanismo de formación de enanas marrones ya que no podría explicar por sí solo la abundancia de estos objetos. Una característica sobresaliente de las enanas marrones es que todas tienen más o menos el radio de Júpiter. En todo el rango de masas posibles, el tamaño del radio solo varía en un 10−15 %. Distinguir una enana marrón de un planeta gigante no es una tarea sencilla, sobre todo cuando, como veremos más adelante, de las observaciones de planetas extrasolares generalmente solo se puede estimar su masa mínima. El primer objeto candidato a enana marrón fue descubierto en 1988, GD 165B, compañera de una enana blanca (Becklin & Zuckerman 1988). Por años existió la controversia de si era realmente una enana marrón. Hoy es el prototipo de las estrellas menos masivas, las de tipo espectral L, que están en la frontera de los objetos estelares. Fue recién en 1995 que se encontraron tres candidatos que no presentaban controversia. El más significativo fue el descubrimiento de Gliese 229B. Este objeto es lo suficientemente frío como para no ser una estrella pero demasiado luminoso y 8 distante de su compañera como para ser un planeta. Su espectro no solo muestra que su temperatura y su luminosidad están por deba jo de la estelar (Oppenheimer et al. 1995), sino que también presenta líneas de absorción de metano en 2 micrones (estas líneas no son esperables en los espectros estelares), una característica asociada, hasta ese momento, solo a nuestros planetas gigantes. Otra característica que distingue a los objetos subestelares son las líneas de litio. Si bien el litio también está presente en los espectros de las estrellas jóvenes, en esos casos se las puede distinguir de las enanas marrones por su tamaño. Las enanas marrones emiten la mayor parte de su flujo en las bandas del infrarro jo cercano. La temperatura superficial se encuentra entre los 2.200 y los 750 K. 1.3.2. Planetas gigantes Planetas gigantes o jovianos es la denominación con la que se conoce colectivamente a Júpiter, Saturno, Urano y Neptuno, los planetas del Sistema Solar caracterizados por sus masivas envolturas gaseosas.4 Júpiter y Saturno son conocidos desde tiempos remotos y estuvieron asociados a la mi- tología y religión de numerosos pueblos. Fueron los romanos quienes les dieron los nombres con los que los conocemos en la actualidad. Júpiter (Zeus para los griegos) era el padre de todos los dioses y, entre otras cosas, era el protector de las actividades militares de los romanos fuera de los límites del Imperio. Saturno, por su parte, era considerado el dios de la cosecha y para los romanos era equivalente al dios Kronos de los griegos. Júpiter es el planeta de mayor masa del Sistema Solar. Su magnitud aparente es -2.8 en oposición y -1.6 en conjunción, lo cual lo hace el tercer objeto más brillante del cielo nocturno, después de La Luna y Venus (aunque en ciertos momentos del año Marte puede ser aún más brillante). En tanto, el nombre de Urano proviene de la antigua Grecia. Urano era una deidad del cielo, padre de Kronos (Saturno) y abuelo de Zeus (Júpiter). Si bien Urano es visible a o jo desnudo, no fue reconocido como un planeta sino hasta fines del siglo XVIII ya que se lo confundía con una estrella. En 1781 William Herschel anunció el descubrimiento de lo que él pensaba era un cometa. Sin embargo, luego de sucesivas observaciones, no logró identificar su cola y, por otra parte, la estimación de sus elementos orbitales indicaban que la órbita era bastante circular. Estos datos sugerían que se trataba de un nuevo planeta, pero Herschel no se atrevía a darle esa clasificación. Sin embargo, rápidamente, el resto de la comunidad científica lo reconoció como tal y luego de varios debates se le dio finalmente el nombre de Urano. Por su parte, Neptuno fue el primer objeto del Sistema Solar predicho matemáticamen- te. Apartamientos de su trayectoria real respecto de la órbita calculada de Urano indujeron a los astrónomos de la época a proponer la existencia de un octavo planeta. Así fue como gracias a los cálculos del francés Le Verrier, se descubrió Neptuno en septiembre de 1846, en la región del cielo en la cual se lo buscaba. El nombre de Neptuno (dios del mar en 4Las características de estos planetas serán desarrolladas en detalle en los capítulos que siguen. Esta sección fue incluida aquí por completitud. 9 la mitología romana) lo propuso el astrónomo F. Struve a fines de ese mismo año y fue ampliamente aceptado por la comunidad científica internacional. 1.3.3. Planetas terrestres Los planetas terrestres (rocosos, telúricos, o interiores) son cuerpos similares a La Tie- rra, compuestos fundamentalmente por silicatos. Según su radio orbital creciente, ellos son: Mercurio, Venus, La Tierra y Marte. Los planetas terrestres son los más cercanos al Sol. Todos los planetas terrestres tienen estructuras similares: un núcleo metálico compuesto fundamentalmente de hierro, rodeado por un manto de silicatos. Las atmósferas de estos planetas son secundarias, ya que fueron generadas por vulcanismos o impactos cometarios, y no como la de los planetas gigantes donde el gas proviene de la nebulosa protoplanetaria. Si bien la mayoría de los planetas extrasolares descubiertos hasta el momento tienen características asimilables a las de nuestros planetas gigantes, el primer sistema de exo- planetas detectado (Wolszczan & Frail 1992) resultó estar compuesto por planetas de tipo terrestres, siendo sus masas 0,02, 4,3 y 3,9 masas terrestres (M⊕ ). La gran diferencia es que estos planetas no orbitan una estrella como el Sol sino un púlsar (PSR B1257+12). Su des- cubrimiento fue completamente accidental: durante la observación del púlsar se detectaron interrupciones en las emisiones de radio, que resultaron ser causadas por los tránsitos de estos planetas. Si no hubieran estado orbitando un púlsar, no habrían sido descubiertos. Cuando en 1995 se descubrió el primer exoplaneta entorno a una estrella de tipo solar (Mayor & Queloz, 1995), 51 Peg b, muchos astrónomos pensaron que se trataba de un gigante terrestre puesto que no se pensaba que fuera posible que un planeta gaseoso pudiera estar tan cerca de su estrella (a = 0,052 UA). Recién en junio de 2005 se detectó un planeta de tipo terrestre alrededor de una enana ro ja, Gliese 876 (Rivera et al. 2005). Este objeto tendría una masa de entre 5 y 7 M⊕ y un período orbital de tan solo 2 días. En agosto del mismo año se observó, con la técnica de microlensing, un planeta de 5,5 M⊕ en una ubicación respecto de su estrella parecida a la del cinturón de asteroides local (OGLE-05-390LB, Beaulien et al. 2006). En abril de 2007, con el telescopio de La Silla (ESO), se descubrió que alrededor de la estrella Gliese 581 (enana ro ja), en su región de habitabilidad5 , hay un planeta cuya masa sería 5 veces la de la Tierra (Udry et al. 2007, Selsis et al. 2007). Sin embargo, todavía no se tiene la certeza de que este planeta sea de tipo rocoso. 1.3.4. Planetas enanos y cuerpos pequeños En la década del ’90 comenzaron a confirmarse observacionalmente los primeros objetos del Cinturón de Kuiper (hasta ese momento solo se conocía el sistema Plutón - Caronte). A 5La región de habitabilidad de una estrella se define como el rango de distancias orbitales en las cuales podría haber potenciales reservorios de agua en estado líquido, ingrediente fundamental para el desarrollo de organismos biológicos complejos (Kasting et al.1993). 10 comienzos de este siglo se descubrieron los primeros objetos transneptunianos comparables con Plutón tanto en tamaño como en sus características orbitales: Quoar, Sedna y Eris. Se sabía entonces que en un corto tiempo aparecerían muchísimos objetos similares y que, aplicando el criterio aceptado hasta ese momento, todos deberían ser llamados planetas (situación parecida a la que ocurrió en el Siglo XIX luego del descubrimiento de Ceres). Esto fue lo que motivó que la IAU redefiniera la palabra planeta, hecho que tuvo como consecuencia que surgieran otras dos clasificaciones nuevas: planetas enanos y cuerpos pequeños del Sistema Solar. Un planeta enano es un cuerpo celeste que: (a) está en órbita alrededor del Sol, (b) tiene la masa suficiente como para que su autogravedad contrarreste las fuerzas de cuerpo rígido, de modo que su forma sea aquella correspondiente a la de equilibrio hidrostático (cuasi esférica), (c) no limpió la vecindad de su órbita, y (d) no es un satélite. Cualquier otro objeto, exceptuando los satélites, que no sea un planeta o un planeta enano será clasificado como un cuerpo pequeño del Sistema Solar. Entre ellos encon- tramos fundamentalmente a los asteroides, la mayoría de los objetos transneptunianos y los cometas. Hasta el momento, la IAU reconoce cinco planetas enanos: Ceres (en el Cinturón de Asteroides), Plutón, Haumea, Makemake y Eris (pertenecientes al Scattered Disk). Se sospecha que otros 40 objetos ya conocidos también clasificarían como planetas enanos, y se estima que se encontrarán muchos más cuando se explore mejor el Cinturón de Kuiper. De hecho, los candidatos corresponden a un survey del hemisferio norte. Por simetría se esperaría encontrar otros tantos en el hemisferio sur. El 11 de Junio de 2008, la IAU anunció una categoría dentro de los planetas enanos, los plutoides. Se define como plutoide a todo cuerpo celeste en órbita alrededor del Sol a una distancia superior a la de Neptuno, con masa suficiente como para que su autogra- vedad sea capaz de contrarrestar la fuerza de cuerpo rígido de modo que su forma sea la correspondiente a la de equilibrio hidrostático (cuasi esférica) y que no haya limpiado la vecindad de su órbita. No hay que confundir el término plutoide con plutino. Un plutino es un objeto transneptuniano en resonancia 2:3 con Neptuno. La denominación plutino remite a Plutón, el cual está atrapado en dicha resonancia. Esta clasificación, que incluye al propio Plutón y a sus lunas, solo se refiere a las propiedades dinámicas de los objetos y no a sus características físicas. 1.4. ¿Por qué estudiar la formación de planetas gi- gantes? El ser humano ha sido siempre curioso, y ha buscado incansablemente respuestas a todas sus preguntas. Exceptuando Urano y Neptuno, el resto de los planetas del Sistema Solar han sido familiares a los o jos de millones de espectadores del cielo desde el principio de la humanidad. En los finales del siglo XX se sumaron, a los planetas ya conocidos, 11 centenares de exoplanetas orbitando otros soles, la mayoría de ellos en configuraciones completamente distintas a la de nuestro Sistema Solar. La búsqueda de otros mundos tiene como objetivo principal encontrar análogos a nuestra Tierra donde se pueda haber desarrollado la vida. El desafío que esto plantea para la ciencia está recién en sus comienzos. En su exploración del Universo, los astrónomos han encontrado sistemas planetarios en configuraciones inesperadas, las cuales seguramente nunca hubieran sido propuestas en una reunión científica de no mediar evidencias irrefutables: Tierras orbitando púlsares, Júpiters a distancias de su estrella central más pequeñas que la de Mercurio al Sol, o mucho mayores que la de Plutón; Uranos y Neptunos calientes... ¿Cómo se formaron estos planetas? La teoría todavía no ha podido encontrar una respuesta. Pero, quizá, lo más sorprendente, es lo poco que todavía sabemos de nuestros “propios planetas”. En la actualidad, no podemos precisar la estructura interna de Júpiter, ni si los planetas gigantes se formaron donde se encuentran hoy. Tampoco sabemos exactamente cuánta de la masa de Urano y Neptuno está en estado gaseoso, ni por qué Urano es el único de los planetas gigantes que emite menos radiación que la esperada, en relación a la que recibe del Sol. De hecho, no contamos con una teoría unívocamente aceptada para explicar la formación de los planetas gaseosos. Los datos observacionales nos muestran, por un lado, la gran variedad de sistemas pla- netarios que pueden existir en el Universo. Pero por otra parte, imponen serias restricciones a los modelos que intentan explicar su origen. Los planetas gigantes están presentes en la mayoría de los sistemas planetarios detectados hasta el momento. En lo que respecta al Sistema Solar, se cree que jugaron un papel fundamental para que en la Tierra pudiera desarrollarse la vida. Sin embargo, la teoría de formación de los planetas gigantes dista mu- cho de estar cerrada. Más allá de las bases del modelo, que podrían considerarse “simples”, existen numerosos actores que intervienen directamente, y pueden modificar de manera sustancial, el escenario de formación de estos objetos. Para un estudio consistente de la formación de los planetas gigantes se necesita contar tanto con resultados precisos para la ecuación de estado de un gas compuesto por hidrógeno y helio, como de la dinámica a gran escala del Sistema Solar. A diferencia de otros problemas astrofísicos, es difícil en este caso aislar y, al mismo tiempo, estudiar en forma realista un determinado aspecto de la cuestión. Por otro lado, es imprescindible hacer simplificaciones, muchas de ellas necesariamente gruesas, si se quiere comenzar a comprender cómo ocurre este proceso. Debido a la dificultad que plantea el problema de la formación de los planetas gigantes, existen todavía muchas incógnitas, y se hace necesario que los modelos adopten hipótesis tan realistas como sea posible si se quiere llegar a una descripción confiable de este proceso. En función de todas las preguntas sin respuesta que rodean al origen de los sistemas planetarios creemos que vale la pena estudiar, en particular, la formación de los planetas gigantes. Dada la diversidad de situaciones que rodean a este problema, hemos decidido circunscribir este estudio a los planetas del Sistema Solar, esperando que en el futuro nuestro modelo pueda ser extendido para atacar el problema de la formación de planetas extrasolares. 12 Capítulo 2 Datos observacionales La naturaleza de los planetas gigantes es muy distinta a la de sus pares terrestres. En este capítulo presentaremos la información que aportan los datos observacionales, tanto en lo que respecta a los planetas del Sistema Solar como a los planetas extrasolares. 2.1. Características de los planetas gigantes del Siste- ma Solar La característica fundamental que distingue a los planetas gigantes de los planetas terrestres es su masa y su composición. Júpiter y Saturno son dos órdenes de magnitud más masivos que la Tierra, y Urano y Neptuno, si bien relativamente mucho más pequeños , superan a la masa de la Tierra en un factor 10. Pero además, en el caso de Júpiter y Saturno, más del 90 % de su masa está compuesta por hidrógeno y helio. A continuación describiremos los aspectos más relevantes que caracterizan al interior de los planetas gigantes. Es importante destacar que esta información se infiere de la construcción de modelos teóricos, en conjunto con los datos observacionales disponibles. Los aspectos generales de los párrafos que siguen corresponden a un curso dictado por T. Guillot (Guillot, 2001). La masa de los planetas gigantes se puede calcular de la observación del movimiento de sus satélites naturales (sus valores se listan en la tabla 1.1). Del análisis de las trayectorias descriptas por los vuelos de las misiones espaciales, sobre todo de aquellas que estuvieron en órbita polar, se pueden obtener estimaciones más precisas de su campo gravitatorio. Debido a la rápida rotación que tienen los planetas, su campo gravitatorio se aparta del correspondiente a un cuerpo esférico. Si expandimos el potencial gravitatorio en polinomios de Legendre, Pi (cos θ), obtenemos, ba jo la hipótesis de simetría axial: i r 1 − ∞ GMp Xi=1 (cid:18) Rec r (cid:19) 13 JiPi (cos θ)! φ(r, θ) = − (2.1) siendo r la coordenada radial, θ el ángulo polar, Mp la masa del planeta, Rec su radio ecuatorial y Ji los momentos gravitatorios. Dado que los planetas gigantes se encuentran casi en equilibrio hidrostático, solo los coeficientes correspondientes a los términos pares resultan no despreciables (los Ji impares son nulos debido a la simetría que existe respecto del ecuador). La estructura interna de estos objetos se puede estimar a partir de su masa Mp , su radio ecuatorial Rec y de los momentos gravitatorios. Los momentos gravitatorios son medidos por las sondas y misiones espaciales que sobrevuelan ocasionalmente a los planetas. Una primera estimación de la densidad se puede obtener de dividir la masa total del planeta, Mp , por el volumen de la esfera de radio Rec . De la tabla 1.1 se puede ver que las densidades de los planetas gigantes rondan la unidad, siendo muy ba jas comparadas con las de los planetas terrestres. Saturno es el menos denso de todos, con una densidad promedio inferior a 1. Cálculos más detallados se obtienen utilizando el período de rotación y las mediciones de los momentos J2 , J4 y J6 . Los momentos gravitatorios están directa- mente relacionados con los momentos de inercia, con lo cual dependen sensiblemente de la velocidad de rotación, de la distribución de la masa, etc. De este modo, el conocimiento de los momentos gravitatorios junto con una ecuación de estado para el gas permiten estimar la densidad ρg = ρg (r, θ), la cual es equivalente (vía la ecuación de estado) al perfil de densidad ρg (P , T ), donde P es la presión y T la temperatura. El procedimiento para obtener información del interior de los planetas es integrar la ecuación de equilibrio hidrostático (incluyendo el término de rotación) junto con una ecua- ción de estado y opacidades apropiadas. Las condiciones de borde para esta integración surgen de los datos observacionales (masa, temperatura superficial, gravedad superficial, etc.). De este modo, el conocimiento de la estructura interna es indirecto y depende fuer- temente del modelo físico empleado. En este proceso, la ecuación de estado (EOS, por sus siglas del término en inglés Equation of State) juega un papel fundamental en la descripción de la estructura interna de los planetas gigantes. La EOS que se utiliza en la mayoría de los cálculos de formación, evolución y estructura de los planetas gigantes es la calculada por Saumon, Chabrier & van Horn (1995, en adelante SCVH; ver capítulo 4 para comentarios al respecto de la EOS). Estos autores calculan la ecuación de estado para el hidrógeno y el helio por separado y sugieren una determinada interpolación para el caso de una mez- cla de ambos gases. Hay que notar que al hacer esto se desprecian las interacciones entre ambas especies. Por otra parte, la presencia de otros elementos que no sean el hidrógeno y el helio, se incorpora utilizando otras ecuaciones de estado, las cuales resultan aún menos confiables. Es importante subrayar que la ecuación de estado SCVH se calcula sobre la base de numerosas simplificaciones pero es, hasta el momento, la más detallada con la que se cuenta. Las atmósferas de los planetas gigantes están compuestas, fundamentalmente, por hi- drógeno molecular (H2 ) y helio (He). Estas moléculas son difíciles de detectar puesto que su momento dipolar es nulo. En cuanto a las líneas producidas por las transiciones electró- nicas, las mismas corresponden a altitudes atmosféricas altas y, por lo tanto, no aportan información sobre la estructura interior en regiones más profundas. La abundancia de ele- mentos pesados, por otra parte, resulta fundamental para entender el proceso de formación 14 de estos objetos. Sin embargo, exceptuando el caso de Júpiter, de donde se tiene informa- ción gracias a la sonda Galileo, las abundancias en el resto de los planetas gigantes ha sido muy pobremente estimada y solo la abundancia de metano (CH4 ) pudo ser medida con confianza. Se cree que tanto en Júpiter como en Saturno los elementos que componen sus envolturas están bien mezclados. En el caso de Júpiter, las mediciones in situ de la sonda Galileo dan como resultado que la abundancia de helio es Y /(X + Y ) = 0, 238 ± 0, 007, siendo Y la fracción en masa de todas las especies de helio, y X la correspondiente al hidrógeno. En cuanto a Saturno, las incertezas todavía son bastante grandes. Los datos con los que se cuenta hasta el momento son los correspondientes a la sonda Voyager 2, que dan como resultado Y /(X + Y ) = 0, 06 ± 0, 05. Según se cree, estos últimos valores son poco confiables y se piensa que en realidad serían mayores. Se espera que, en breve, los resultados de las mediciones de la misión Cassini-Huygens puedan aportar datos más precisos. De todos modos, la abundancia superficial de helio en estos planetas resulta siempre inferior a la solar1 . Sin embargo, si aceptamos que la nebulosa primordial tenía abundancias solares, debe haber operado algún mecanismo para que las abundancias superficiales observadas en los planetas exteriores sean menores a las solares. Una posibilidad es que exista una separación de fases entre el hidrógeno y el helio, en la cual “gotitas” de helio puedan crecer lo suficientemente rápido como para precipitar hacia el interior por la acción de la gravedad y no ser afectadas por la convección. Una evidencia que apoya esta hipótesis es la deficiencia en la abundancia de neón, dado que el neón tiende a disolverse en el helio, con lo cual habría descendido al interior dentro de estas gotas. En la atmósfera de Júpiter se detecta la presencia de metano, vapor de agua, amoníaco, carbono y oxígeno, entre otros “elementos pesados”. En lo que se refiere a elementos pesados en general, el enriquecimiento sería de un factor entre 2 y 4 en relación a la abundancia solar. De todas formas, todavía hay elementos cuyas abundancias no han podido ser bien determinadas. Basados en datos espectroscópicos, la composición de Saturno podría ser muy similar a la de Júpiter, o incluso aún más rico en elementos pesados. Utilizando la EOS de SCVH, Saumon & Guillot (2004) estimaron, dentro del rango de incertezas de la EOS, el estado del gas en los interiores de Júpiter y Saturno. Mientras que el interior de Saturno se encuentra dentro de una región bastante bien estudiada de la ecuación de estado de la mezcla hidrógeno-helio (donde el hidrógeno es molecular), en el caso de Júpiter la situación es diferente. Puesto que gran parte de su estructura se encuentra a presiones intermedias, donde la ecuación de estado es bastante menos conocida, las incertezas entorno a estas estimaciones son grandes (ver capítulo 4 y figura 4.4). El modelo teórico que se propone para el cálculo del interior de estos planetas distingue tres estructuras: un núcleo central, denso, presumiblemente formado por rocas (silicatos, hierro) e hielos (agua, amoníaco, metano); un interior rico en helio y una envoltura externa deficiente en helio (ésta última hipótesis basada en las observaciones). Una representación 1Los modelos evolutivos estelares estiman que la relación del helio relativa al hidrógeno en el Sol es Y /(X + Y ) = 0, 270 ± 0, 005. 15 esquemática del interior de los planetas gigantes se muestras en la figura 2.1. La existencia del núcleo sólido resulta generalmente necesaria para reproducir los campos gravitatorios medidos. Es importante destacar, sin embargo, que para los modelos de formación, las cantidades más relevantes que se quieren estimar son: la masa del núcleo (Mc ) y la masa de elementos pesados contenida en el interior del planeta (MZ ). La determinación de estos parámetros surge de aquellos modelos que mejor a justen los valores del Rec y los momentos gravitatorios J2 y J4 . En el caso de Júpiter, de acuerdo con este modelo, la masa del núcleo podría acotarse entre 0 y 11 M⊕ , mientras que la masa de elementos pesados estaría entre 1 y 39 M⊕ . Comparado con la abundancia de elementos pesados en el Sol, estos valores corresponden a un enriquecimiento de un factor entre 1,5 y 6. El caso de Saturno es menos problemático, ya que al ser menos masivo que Júpiter resulta menos sensible a las incertezas de la EOS en los regímenes correspondientes al interior de este planeta. Los valores estimados para el núcleo son 9 M⊕ < Mc < 22 M⊕ , y para el resto de los elementos pesados en la envoltura 1 M⊕ < MZ < 8 M⊕ . Esto representa un enriquecimiento de un factor entre 6 y 14 por encima de la abundancia solar. Recientemente, Militzer y colaboradores (2008) calcularon el interior de Júpiter utili- zando una EOS derivada a partir de primeros principios. El modelo empleado no presupone la estructura del planeta dividida en tres regiones sino simplemente compuesta por una parte “sólida” (el núcleo) y una envoltura gaseosa. Según este modelo, la masa del núcleo sería de 16 ± 2 M⊕ , siendo la metalicidad de la envoltura mucho menor que la estima- da por Saumon & Guillot (2004). A diferencia de los modelos calculados con la EOS de SCVH, donde la masa del núcleo era bastante pequeña y las abundancias en el manto eran grandes, en el modelo de Militzer et al. la situación se revierte. Según las mediciones de la sonda Galileo, si se extrapolan los valores superficiales (con un modelo adecuado) a toda la envoltura, la masa total de hielos fuera del núcleo no superaría las 6 M⊕ . En el modelo de Militzer et al., la masa de hielos sería de 4 M⊕ , en buen acuerdo con los datos observacionales. Por otra parte, los valores de los momentos multipolares también reflejan la rotación del planeta; cuanto más altos sean más afectados se verán por rotaciones no uniformes. Para a justar el valor de J4 , Militzer et al. no necesitan proponer la diferenciación de la envoltura como lo hacen Saumon & Guillot. Por las características de su modelo, ellos sugieren que el valor de J4 está afectado por los vientos superficiales y, de hecho, ba jo esta propuesta ellos logran a justar este valor sin necesidad de modificar otros parámetros. Los resultados de Militzer et al. (2008) son muy recientes y probablemente no hayan sido todavía discutidos en profundidad por la comunidad científica. Además, por el momento, la EOS que emplearon para calcular sus resultados no fue publicada en forma de tabla, por cuanto no es para nada trivial intentar hacer cálculos bajo las hipótesis de su modelo. Por su parte, Nettelman y colaboradores (2008) también estimaron, a partir de una EOS obtenida por primeros principios, el interior de Júpiter. Sin embargo, sus resultados no concuerdan con los de Militzer et al. De este modo, la masa del núcleo y la abundancia total de sólidos presente en el interior de de los planetas gigantes sigue siendo uno de los ejes principales de discusión en relación a la estructura interna de los planetas. 16 Figura 2.1. Representación esquemática del interior de Júpiter, Saturno, Urano y Neptuno (figura tomada del artículo de Guillot, 1999). 17 En lo que respecta a las atmósferas de Urano y Neptuno se observa un enriqueci- miento en elementos pesados de hasta un factor 30 por encima de la composición solar, fundamentalmente de CH4 . Los perfiles de densidad que se pueden inferir de los momen- tos gravitatorios hacen pensar que estos planetas están compuestos fundamentalmente por hielos (mezcla de H2O, CH4 y NH3 ), y que solo en su atmósfera predominan el hidrógeno y el helio. Los modelos del interior de Urano y Neptuno proponen también una estructu- ra de tres capas: un núcleo central rocoso (compuesto fundamentalmente por silicatos e hierro), una capa de hielos y una atmósfera gaseosa. Cada una de estas capas no estaría homogéneamente mezclada. La existencia de estas regiones inhomogéneas confirmaría la presunción de que el hidrógeno se encuentra confinado solo a la atmósfera. Los modelos predicen que la relación entre hielos y rocas es del orden de 10, mucho mayor que el valor protosolar, que sería 2,5. Por otra parte, si se impone que la relación entre hielo y rocas sea acorde a la composición solar se obtiene que, en rasgos generales, ambos planetas están formados por un 25 % de rocas, un 60 − 70 % de hielos y el restante 5 − 15 % de hidrógeno y helio. Esto resulta en que la estimación de la masa de gas para Urano sea de 3 M⊕ y para Neptuno de 5 M⊕ . La estructura característica de los planetas gigantes puede observarse en la figura 2.1. Las opacidades son otro punto fundamental a la hora de explicar el interior de los planetas. A ba jas temperaturas (∼ 1500 − 2000 K) la opacidad media de Rosseland está dominada por los granos de polvo. Las tablas de opacidad que se usan en astrofísica se calculan generalmente imponiendo la existencia de un equilibrio químico global donde las especies condensadas son retenidas y además, suponiendo que la distribución de tamaños de las partículas es la correspondiente al medio interestelar. Sin embargo, no está claro que esta última hipótesis sea válida para el caso de los planetas gigantes y las enanas ma- rrones. Dado que el interior de los planetas sería mayormente convectivo, esto modificaría la distribución de tamaños de los granos, hecho que tendría profunda repercusión en los modelos de formación de los planetas gigantes (Podolak 2003). En lo que respecta a Urano y Neptuno, la opacidad de Rosseland es todavía incierta. De las estimaciones hechas al mo- mento, ambos planetas tendrían una región radiativa a temperaturas de alrededor de 1500 K. Sin embargo, estas regiones podrían desaparecer de existir ciertas fuentes de opacidad como el sodio y el potasio, presentes en las enanas marrones. Sin embargo, vale la pena notar que si bien la existencia de una zona radiativa es bastante significativa a la hora de calcular la evolución de estos objetos, no es determinante para el cálculo de la estructura interior. Júpiter, Saturno y Neptuno emiten mucha más energía de la que reciben del Sol. El caso de Urano es menos claro: si bien su flujo intrínseco es mucho menor que el del resto de los planetas gigantes, está en duda si emite o no más energía de la que recibe del Sol. En el caso de Júpiter se mostró que esto puede ser explicado por la continua contracción y enfriamiento del planeta. Una consecuencia importante de la existencia del flujo de calor intrínseco en un planeta es que requiere de temperaturas del orden de los 104 K o más, lo cual implica que los planetas gigantes son fluidos, y que su interior resultaría ser esencialmente convectivo. Estimar las temperaturas superficiales es por demás complicado pero, asumiendo como 18 superficie la correspondiente a 1 bar de presión se obtiene: para Júpiter, T = 165 ± 5 K, para Saturno T = 135 ± 5 K, para Urano T = 76 ± 2 K y para Neptuno T = 72 ± 2 K. Las atmósferas de los planetas gigantes son de naturaleza compleja y turbulenta. Por ejemplo, los vientos son muy variables dependiendo de la latitud. En el caso de Júpiter y Saturno, los vientos en el ecuador tienen una velocidad mucho mayor que en los polos, mientras que para Urano y Neptuno se da a la inversa. Sin embargo, dado que los cuatro planetas tienen altas velocidades de rotación entorno a su propio eje, las velocidades de los vientos son más ba jas que la velocidad de rotación (la velocidad típica de los vientos en Júpiter es de 360 km h−1 ). Los vientos en Saturno tienen velocidades mucho más altas que en Júpiter, pudiendo alcanzar los 1.800 km h−1 . Sin embargo, es Neptuno el planeta con los vientos más intensos del Sistema Solar, llegando a velocidades de hasta 2.100 km h−1 . Además de los vientos, estos planetas presentan tormentas, que pueden tener la extensión de todo el planeta, y durar por semanas o por siglos. La gran mancha ro ja de Júpiter es un centro anticiclónico de 12.000 km de diámetro, situado 22◦ al sur del ecuador con, al menos, 300 años de edad. Los modelos indican que la tormenta es estable y que podría ser una característica permanente del planeta. Júpiter está cubierto por nubes compuestas de cristales de amoníaco y posiblemente bisulfuro de amonio. Las nubes están en la tropopausa y cubren su superficie en forma de bandas. Los colores marrones y anaranjados de las nubes provienen de compuestos que suben desde el interior a la superficie y que cambian de color cuando son expuestos a la luz ultravioleta solar. Si bien la composición exacta se desconoce, las posibles sustancias que la componen podrían ser fósforo, sulfuros o hidrocarburos. En lo que a Saturno respecta, se observaron tormentas desarrollándose en todo el planeta. Por último, debemos hacer mención a los satélites y anillos de los planetas gigantes. De acuerdo a sus características orbitales, los satélites se pueden clasificar en regulares e irregulares. Los primeros generalmente se encuentran en órbitas casi circulares en el plano del ecuador, rotando en sentido directo. Los irregulares, por su parte, no tienen una forma definida y describen órbitas excéntricas, alargadas y retrógradas. Al momento, sabemos que Júpiter cuenta con 63 satélites, de los cuales ocho son regulares y los restantes son irregulares. Los satélites regulares se dividen en dos grupos: los cuatro interiores (Grupo Amaltea) y los cuatros satélites galileanos, descubiertos por Galileo hace 400 años. Estos últimos, en orden decreciente según su tamaño, son: Ganímides, Calisto, Io y Europa. Saturno, por su parte, tiene 60 satélites naturales, siendo el más grande Titán que, después de Ganímides, es el segundo en tamaño en el Sistema Solar. Urano y Neptuno cuentan con 27 y 13, respectivamente. Se cree que los satélites regulares se formaron en el disco de acreción del planeta, mientras el planeta se encontraba en sus últimas etapas del proceso de formación. Los irregulares, en cambio, serían planetesimales capturados por el campo gravitatorio del planeta. Los cuatro planetas gigantes también tienen anillos y se supone que los satélites son los que, constantemente, les proveen el material debido a las colisiones que se producen entre ellos. De entre todos, el sistema de anillos de Saturno es el más espectacular y el único que puede observarse con binoculares. De hecho, el área que ocupan los anillos es tan grande que al reflejar la luz que proviene del planeta resultan ser tan o más brillantes que el propio Saturno. 19 2.2. Planetas extrasolares Si bien esta Tesis no trata el problema específico de los planetas extrasolares, es inne- gable que el interés por los planetas gigantes del Sistema Solar fue reavivado a partir del descubrimiento de estos objetos alrededor de otras estrellas. Creemos entonces adecuado hacer mención de las características más importantes de los planetas extrasolares, como así también de los mecanismo que permiten su detección, y de las preguntas y teorías que surgen en base a estos descubrimientos. 2.2.1. Métodos de detección2 Los métodos de detección de planetas extrasolares pueden clasificarse en dos grandes grupos: los indirectos y los directos. Se dice que un método es indirecto si a través de él se puede inferir la presencia de un planeta, lo cual significa que lo que se detecta son los efectos que tiene el planeta sobre la estrella y no una señal propia proveniente del planeta. Variaciones regulares en la posición de una estrella en el cielo, corrimientos por efecto Doppler de las líneas espectrales y cambios periódicos en la luminosidad son algunos de los efectos medibles que pueden indicar la presencia de un planeta orbitando entorno a una estrella. Veremos a continuación que hay numerosas técnicas de detección indirecta de planetas extrasolares. Por otra parte, cuando podemos observar la luz proveniente del planeta, cualquiera sea la longitud de onda involucrada, decimos que el método de detección es directo. Métodos indirectos Existen numerosas técnicas que permiten la detección indirecta de exoplanetas. En general, cada una de ellas resulta más efectiva que otra dependiendo de la configuración del sistema planetario, por cuanto cada técnica tiene asociado un sesgo intrínseco respecto de las características del objeto detectado. A continuación describiremos brevemente cuatro de las técnicas más utilizadas las cuales, a su vez, presentan marcadas diferencias conceptuales entre sí. Velocidades radiales El método de las velocidades radiales o de corrimiento Doppler es, hasta el momento, el método de detección de planetas extrasolares más exitoso, puesto que la mayoría de los descubrimientos han estado relacionados a esta técnica. De hecho, el primer exoplaneta orbitando alrededor de una estrella de tipo solar, 51 Peg b, fue descubierto en 1995 de 2La bibliografía consultada para el armado conceptual de esta sección corresponde a los cursos dictados durante la XVI Canary Island Winter School por T. Brown, L. Doyle y S. Udry (ver bibliografía: Brown 2008, Doyle 2008, Udry 2008). 20 esta manera (Mayor & Queloz 1995). Consideremos un sistema estrella-planeta. Tanto la estrella como el planeta describen una órbita en torno al centro de masa del sistema. Si nos focalizamos en la estrella, un observador en la Tierra podrá detectar el movimiento de la misma en la dirección de la visual a través, por ejemplo, de un estudio de su espectro. Las variaciones en la velocidad radial pueden inferirse a través mediciones altamente pre- cisas de corrimientos en las líneas del espectro de la estrella. Así, mediante un detallado análisis espectroscópico, se podrá determinar el corrimiento periódico al ro jo o al azul de las líneas espectrales, cuya magnitud será ∆λ/λ = v/c, siendo v la velocidad de la estrella alrededor del baricentro del sistema. Las variaciones de las líneas espectrales solo miden la componente del movimiento relativo que se aleja o se acerca del observador. Del análisis de los espectros obtenidos a lo largo de uno o varios períodos orbitales se puede inferir la curva de velocidad radial de la estrella, cuya amplitud máxima, que puede ser deducida de las ecuaciones de Newton y de la segunda ley de Kepler, será: K = 2πG Porb Mp sin i (Mp + M∗ )2/3 1 (1 − e2 )1/2 (2.2) donde Porb es el período orbital del planeta, e es la excentricidad de la órbita, e i es su inclinación entre el plano del cielo y el plano orbital. K , Porb y e se pueden determinar si se toman varios espectros durante una órbita completa del planeta. Notemos que como Mp ≪ M∗ , Mp + M∗ ∼ M∗ , la ecuación anterior no depende de la masa del planeta sino de la masa mínima, Mp sin i. De este modo, estimando M∗ de la clasificación espectral, se puede despejar Mp sin i. Si bien la masa del planeta no podrá ser determinada ya que la geometría del problema es a priori desconocida, podemos estimar qué probabilidades hay de que la masa del planeta sea muy distinta a la masa mínima. Claramente, si i ≃ 90 ◦ , Mp sin i ≃ Mp , pero si sin i ≃ 0 el valor estimado será muy distinto del valor real de la masa del objeto. Sin embargo, estos casos son poco probables: si suponemos que las orientaciones orbitales están distribuidas al azar, la probabilidad que sin i < 0,1 será tan solo del 5 %. La detección de exoplanetas con esta técnica depende de la precisión con la que se puede medir el corrimiento Doppler de las líneas. Por ejemplo, para medir variaciones del orden de 10 m s−1 en la velocidad radial de la estrella, la sensibilidad de la medición tiene que ser tal que permita detectar una variación de 10−3 en el ancho de una línea (un valor típico es 10−4 Å). El efecto de Júpiter sobre el Sol es de 12 m s−1 , el de Saturno 4 m s−1 y el de la Tierra de tan solo 0,1 m s−1 . Por otra parte, existe un límite para este tipo de detección impuesto por las fluctuaciones intrínsecas de la superficie estelar. Este límite es de 1 m s−1 . Sin embargo, la detección de planetas como Júpiter resulta muy útil para comparar sistemas planetarios extrasolares con nuestro propio Sistema Solar, como así también para estimar la efectividad de planetas de tipo jovianos como escudos protectores de potenciales planetas terrestres, en órbitas interiores, que estén en condiciones de desarrollar alguna forma de vida. 21 Tránsitos Supongamos que el observador, el plano de la órbita del planeta y la estrella se encuen- tran los tres sobre la línea de la visual (i ≃ 90◦). Cuando el planeta pasa por delante de la estrella, oculta parte del disco estelar, de modo que el observador detecta una disminución del brillo de la estrella mientras este evento ocurre. Si se acepta que la orientación de las órbitas sigue una distribución al azar, la probabilidad de observar un tránsito se puede estimar con tres parámetros: R∗ , Rp y a, y resulta ser: . (2.3) Probabilidad = R∗ + Rp a a , lo cual indica que es más probable observar Como R∗ ≫ Rp , la probabilidad se reduce a R∗ tránsitos cuanto menor sea la separación orbital entre el planeta y la estrella. En el caso del Sistema Solar, la probabilidad de que un observador externo detecte un tránsito de Mercurio es del orden del 1 %, mientras que para el caso de Júpiter cae a apenas 0,1 %. En el caso de los hot Jupiters3 la probabilidad puede ascender a un 20 % debido a los pequeños semiejes de estos objetos. Dado que no hay una forma clara de mejorar la probabilidad de detección de un tránsito planetario, la única estrategia disponible es la observación de un gran número de estrellas, ya sea con programas automatizados de búsqueda de variaciones fotométricas en estrellas muy brillantes, o bien a través del estudio detallado de las estrellas más débiles confinadas a una determinada región del cielo. Existen dos técnicas de detección de tránsitos: la fotométrica y la espectroscópica. En los tránsitos fotométricos lo que se detecta es la sombra que produce el planeta cuando pasa por delante del disco estelar. El área de cobertura del disco estelar dependerá del radio del planeta ya que cuanto mayor sea el tamaño del planeta, más significativa será la disminución del brillo estelar. De hecho, el cambio en la luminosidad ∆L será: 2 ∆L L ∝ (cid:18) Rp R∗ (cid:19) donde L es la luminosidad total del sistema cuando no se encuentra en situación de tránsito. Dado que esta técnica depende del radio y no de la masa del planeta, permite la detección de planetas no necesariamente muy masivos. Además, para un dado límite de detección, la observación de estrellas pequeñas permitiría la detección de planetas pequeños. Si el sistema tiene observaciones espectroscópicas que permitan calcular el corrimiento Doppler de las líneas, se puede obtener la masa real de objeto, ya que ésta será aproximadamente igual a la masa mínima. De este modo, como conocemos el radio y la masa del objeto podemos estimar ciertos parámetros físicos muy relevantes, como ser su densidad media. Del análisis de la curva de luz se puede obtener información sobre la estructura del planeta, de posibles satélites en órbita como también la potencial presencia de otros compañeros planetarios. (2.4) , 3Los llamados hot Jupiters son planetas extrasolares con masa similar a la de Júpiter pero que se encuentran muy cerca de su estrella central, como en el caso de 51 Peg. 22 El análisis espectroscópico de un tránsito puede aportar muy valiosa información, pero es mucho más difícil de detectar que su contraparte fotométrica ya que el número de fotones que involucra la obtención de un espectro es menor y la precisión requerida es, por otra parte, mayor. Supongamos al planeta y a la estrella como cuerpos esféricos, el observador medirá radios diferentes dependiendo de la opacidad de la atmósfera del planeta. Así, el radio del objeto que provoca el ocultamiento depende de la longitud de onda en que se observe, la cual, según la región de la atmósfera que atraviese, será completa o parcialmente absorbida. Esto lleva directamente a la observación de la atmósfera de un planeta extrasolar. En el caso de la Tierra, por ejemplo, su atmósfera es opaca por deba jo de los 3.000 Å con lo cual, si se la observara durante un tránsito en longitudes de onda cortas, se la vería 60 km más grande que en longitudes de onda largas. Microlensing Consideremos que tenemos en el campo de la visual dos estrellas, una de las cuales se encuentra mucho más lejos de nosotros que la otra. Supongamos que la más cercana, a la que llamaremos lente L, en su desplazamiento relativo, se alinea con la otra (S), ocultándola por un cierto intervalo de tiempo. Debido a efectos relativísticos relacionados con la curvatura del espacio-tiempo, la luz que proviene de la estrella más lejana S se curvará al encontrarse con la estrella L. Este encuentro causa un incremento temporario del brillo combinado de ambas estrellas debido a la amplificación que sufre la estrella S. Este efecto fue observado por primera vez en galaxias y se conoce como lentes gravitacionales. Si la estrella lente L tiene, a su vez, un planeta en órbita, el patrón de amplificación del brillo se desviará del patrón estándar debido a que el planeta modifica el campo gravitatorio de su estrella. Si bien la probabilidad de alineación de dos estrellas es de una en un millón, si esto ocurre, la probabilidad de que un planeta provoque una amplificación que exceda el 5 % de la propia amplificación producida por la estrella L es de uno en cinco. La duración de estos eventos depende de la velocidad relativa con que se desplaza la lente, de la distancia, y de la masa del planeta. Por ejemplo, si la estrella lente se encuentra a 5 kiloparsecs, para un planeta de la masa de Júpiter, el evento durará 3 días y para un planeta como la Tierra tan solo 4 horas. En el primer caso, la amplificación del brillo será de 3 magnitudes y en el segundo de 1 magnitud. Como se ve, estos eventos son detectables con los instrumentos actuales. Sin embargo, no son previsibles ni repetibles, por cuanto se requiere la observación constante de un gran número de estrellas lejanas para tener la posibilidad de registrarlos. Este método permite obtener información sobre los datos orbitales del planeta como así también la razón entre la masa de la estrella y del planeta. Al momento, es el método que mejor funciona a grandes distancias y el que, seguramente, aportará información planetaria teniendo en cuenta poblaciones estelares diferentes, como pueden ser la del disco y del bulge galáctico. 23 Astrometría α = , La búsqueda astrométrica de exoplanetas es la más antigua de todas estas técnicas. De hecho, desde mediados del siglo XX se hacen anuncios de planetas extrasolares debido mediciones astrométricas pero, después de un análisis más profundo, estos objetos resultan ser demasiado masivos para ser considerados planetas. La astrometría mide variaciones periódicas de la posición de una estrella en el plano del cielo (eliminando previamente el movimiento aparente de la estrella debido al movimiento propio y la parala je anual). Si la estrella tiene un planeta girando en torno a ella describirá un movimiento elíptico alrededor del baricentro del sistema estrella-planeta cuyo semieje mayor, medido en segundos de arco, será Mp ap M∗ d si d (distancia entre el observador y la estrella) se mide en parsecs y ap en UA. Para que esta técnica sea efectiva, se deben tomar numerosas imágenes de la estrella durante al menos una fracción importante del período orbital del planeta. Este método es complementario al de velocidades radiales ya que es más sensible a largos períodos (lo cual es equivalente a grandes semiejes), mientras que el método de velocidades radiales favorece la detección de exoplanetas en órbitas cercanas a la estrella. De la ecuación 2.5 se puede ver que la detección de exoplanetas con esta técnica depende de la distancia al sistema, por cuanto, por el momento, queda limitada a estrellas cercanas. Para tener una idea de las magnitudes relacionadas con esta técnica, si consideramos una estrella a 5 parsecs de distancia, el efecto que veríamos si un “Júpiter” estuviera en órbita alrededor de ella sería un desplazamiento de un milisegundo de arco en el plano del cielo, mientras que para una “Tierra” sería de tan solo 0,6 microsegundos de arco. En la actualidad, las mediciones astrométricas están en el límite de la precisión necesaria para poder detectar estas variaciones. Recién en junio de 2009 se anunció el primer exoplaneta descubierto mediante esta técnica4 (Pravdo & Shakland 2009) y se espera que la astrometría aporte numerosos hallazgos en el futuro cercano. De hecho, la misión de la NASA Kepler, destinada a la búsqueda de tránsitos permitirá también la búsqueda astrométrica usando los mismos datos fotométricos. (2.5) Métodos directos La obtención directa de la imagen de un planeta en la región del visible depende de la luz que éste refleja de la estrella central, lo cual, a su vez, depende de la separación orbital con la estrella, del tamaño del planeta y de la naturaleza de su atmósfera. Registrar esta luz requiere de telescopios muy poderosos. La razón entre el brillo del planeta y de la estrella es el factor determinante para la detección ya que de eso depende que el planeta quede enmascarado por el brillo de los anillos de difracción que provoca la estrella en el telescopio. 4Previamente, las mediciones astrométricas permitieron la confirmación de Gliese 876b como exopla- neta (Benedict et al. 2002), el cual fue descubierto espectroscópicamente (Delfosse et al. 1998, Marcy et al. 1998). 24 Por ejemplo, para una estrella a 5 parsecs y un brillo de 1 magnitud, el brillo de un planeta como Júpiter será de 23 magnitudes y el de una Tierra, de 25. La búsqueda de exoplanetas con esta técnica requiere de herramientas específicas como ópticas adaptativas para corregir los efectos atmosféricos que distorsionan la imagen y de coronógrafos para bloquear la luz que proviene de la estrella. El uso de un coronógrafo es la idea del proyecto TPF–C. Por su parte, TPF–I hará su búsqueda utilizando interferometría destructiva. La idea es que la luz proveniente de la estrella interfiera destructivamente, cancelando así el flujo proveniente de la estrella. El planeta, por su parte, como se encuentra en otra ubicación angular, no se verá afectado. El gran contraste que existe ente la estrella y el planeta se puede mejorar si en vez de observar en el visible se observa en el infrarro jo. En estas longitudes de onda el planeta no solo refleja parte de la luz que recibe de la estrella sino que también emite su propia energía térmica. Por otra parte, los planetas de tipo joviano que tengan asociado un gran campo magnético (por ejemplo, esto puede ocurrir si los planetas tienen un núcleo metálico y si, a su vez, rotan relativamente rápido) pueden emitir un flujo de radiación bastante importante en las longitudes de onda de radio. Si consideramos un planeta como Júpiter alrededor de una estrella de tipo solar, en la frecuencia de 10 MHz, el cociente de flujo planeta-estrella podría llegar a 4 (el planeta más brillante que la estrella), en una etapa de ba ja actividad estelar. Sin embargo, los electrones presentes en el medio interestelar añaden mucho ruido a la detección en este rango de longitudes de onda, por cuanto, por el momento, no es posible la detección en este rango de frecuencias. 2.2.2. Programas espaciales Actualmente en funcionamiento Spitzer Space Telescope El telescopio espacial Spitzer fue puesto en órbita alrededor del Sol en agosto de 2004 y se encuentra orbitando detrás de la Tierra. Traba ja en el infrarro jo. Entre las detecciones más importantes se encuentra un posible cinturón de asteroides alre- dedor de una estrella cercana y la primera detección de emisión térmica de varios planetas extrasolares. Spitzer Homepage: http://www.spitzer.caltech.edu/ CoRoT (Convection, Rotation and planetary Transits) La misión espacial CoRoT fue idea- da con el objeto de obtener información sobre la estructura estelar, mediante técnicas astrosismológicas, y, por otra parte, detectar planetas extrasolares utilizando el método de tránsitos. Además, es la primera misión diseñada para la búsqueda de planetas extrasolares de tipo terrestres. El satélite fue lanzado el 27 de diciembre de 2006 y se encuentra en una órbita polar circular a 896 km de altitud. Está previsto que la misión se desarrolle por el plazo de dos años y medio. Cuenta con un telescopio afocal de 27 cm y 4 cámaras CCD. CoRoT es una misión liderada por la French National Space Agency, aunque el consorcio está integrado también por ESA (Agencia Espacial Europea), Alemania, España, Bélgica, Austria y Brasil. Hasta el momento, los descubrimientos más significativos fueron: la de- tección de dos planetas extrasolares, la detección de un objeto que clasificaría entre enana 25 marrón y planeta gigante (una masa de alrededor de 20 masas de Júpiter y su radio de aproximadamente 0,8 radios de Júpiter) y otro exoplaneta que presenta un período de translación igual al período de rotación de su estrella. Corot Homepage: http://smsc.cnes.fr/COROT/ EPOCh (Extrasolar Planet Observations and Characterization) Es un proyecto que uti- liza las cámaras de la estación espacial Deep Impact (puesta en órbita en 2005) para buscar tránsitos de exoplanetas, observar el movimiento en el plano del cielo de estrellas con planetas y analizar la luz reflejada por los planetas. Se encuentra funcionando desde comienzos de 2008. Homepage: http://discovery.nasa.gov/epoxi.html Proyectos TPF (Terrestrial Planet Finder) TPF es un proyecto de la NASA para la búsqueda de planetas de tipo terrestres en regiones potencialmente habitables. TPF consta de dos instrumentos a bordo: un coronógrafo (TPF-C) y un interferómetro (TPF-I). Actualmente el proyecto está muy poco avanzado y si bien las estimaciones de lanzamiento son: TPF-C en 2014 y TPF-I en 2020, el proyecto siempre es postergado. TPF Homepage: http://planetquest.jpl.nasa.gov/TPF Kepler se encargará de monitorear fotométricamente, con un telescopio de un metro de diámetro, cientos de miles de estrellas en búsqueda de tránsitos de planetas tipo terrestres. Está previsto el lanzamiento de Kepler para el año 2009. De esta manera se podrá obtener una estimación estadística de cuál es la frecuencia de planetas como la Tierra en sistemas extrasolares. Kepler Homepage: http://kepler.nasa.gov/ SIM (Space Interferometry Mission) se encuentra en desarrollo y su objetivo será deter- minar las posiciones y distancias a cientos de estrellas en forma más precisa que cualquier programa previo. SIM está siendo desarrollada por el Jet Propulsion Laboratory ba jo con- trato de la NASA. Tendrá como objetivo el primer censo de planetas tipo terrestres que orbiten estrellas cercanas. Este último hecho tiene que ver con que se hará detección direc- ta de estos objetos de modo de poder obtener información sobre sus propiedades físicas, sus masas y los elementos orbitales. Esta misión está siendo específicamente diseñada para optimizar las mediciones astrométricas y será capaz de detectar el desplazamiento en el cielo de una estrella, con una precisión de un microsegundo de arco. Kepler y SIM son com- plementarias a TPF: Kepler brindará estadísticas de la frecuencia de planetas terrestres utilizando estrellas distantes y SIM hará censos en estrellas cercanas, con lo cual ambos proveerán una selección de objetos para ser luego observados por TPF. SIM Homepage: http://planetquest.jpl.nasa.gov/SIM 26 Darwin Darwin está previsto que sea una flotilla de cuatro o cinco telescopios espacia- les cuyo objetivo será la búsqueda de planetas terrestres extrasolares y, consecuentemente el análisis atmosférico en busca de señales químicas de vida. Dado que en el óptico sería casi imposible detectar este tipo de objetos debido a la luminosidad de la estrella central, Darwin hará su búsqueda en el infrarro jo. Por otra parte, es en el infrarro jo donde la vida tal como la conocemos en la Tierra deja señales detectables. Uno de los fundamentos para que el telescopio se encuentre en el espacio es la necesidad de que esté a muy ba jas tem- peraturas, de modo que él mismo no emita radiación en la longitud de onda a observar, para poder así detectar señales muy tenues de planetas lejanos. Tres de las naves espaciales transportarán el telescopio que tendrá un diámetro de aproximadamente 3 metros. Darwin operará desde el punto lagrangiano L2, que se encuentra a 1,5 millones de km de la Tierra en dirección opuesta al Sol. Si bien todavía no hay fecha estimada para el lanzamiento, dada la similitud de objetivos y la complejidad de los proyectos Darwin y TPF, ESA y NASA están pensando en unificarlos y hacerlos operar en conjunto. Darwin Homepage: http://www.esa.int/science/darwin GAIA Gaia es un proyecto de la ESA. Entre otros, su objetivo es medir, con una precisión de 20 microsegundos de arco, las posiciones de mil millones de estrellas (que se encuentran hasta 200 parsecs del Sol), tanto de nuestra Galaxia como de otros miembros del Grupo Local (observaciones astrométricas). También hará observaciones espectroscópicas y foto- métricas de todos los objetos. Su lanzamiento está previsto para fines del año 2011. GAIA Homepage: gaia.esa.int/ 2.2.3. Características de los planetas extrasolares detectados Al día de hoy, y durante los 13 años desde la confirmación del primer planeta extra- solar descubierto alrededor de una estrella de tipo solar (hay registros previos de planetas orbitando alrededor de púlsares), se han detectado 340 candidatos a planetas extrasolares (ver La Enciclopedia de Exoplanetas http://exoplanet.eu). El hecho que se los mencione como candidatos (aunque, generalmente, se utilizará el término planeta o exoplaneta) está relacionado con que no teniéndose una definición de la palabra planeta afuera del Sistema Solar, se considerará ad-hoc que son planetas los objetos subestelares de masa inferior a 20 MJ (no entran dentro de esta clasificación los llamados “free-floating planets” u objetos con masa planetaria que se encuentran solos en el espacio). Teniendo en cuenta que más del 90 % de los mismos fue detectado por el método de velocidades radiales (menos del 25 % presentan, a su vez, tránsitos), este límite para la masa es todavía más arbitrario si se tiene en cuenta que, en general, solo se puede estimar la masa mínima del objeto. Otro motivo para no hacer mucho hincapié en el número de candidatos detectados es que aumentan día a día debido a la alta tasa de descubrimientos, consecuencia del mejoramiento en la precisión instrumental. De hecho, el 36 % de los exoplanetas fueron detectados entre los años 2007–2009. 27 Del análisis de los datos disponibles se puede decir que, de los 340 exoplanetas, habría 16 (∼ 5 %) con masa menor a 10 M⊕ . Claramente, esto no significa que los planetas más pequeños sean menos abundantes sino que las capacidades tecnológicas no han permitido todavía su detección en un gran número. Como veremos a continuación, las limitaciones tecnológicas traen aparejado un sesgo en la detección, viéndose naturalmente favorecido el descubrimiento de planetas alrededor de estrellas cercanas (d < 400 parsecs, con un máximo entorno a los 30 parsecs). El método de velocidades radiales es el que ha tenido más éxito en la búsqueda de exoplanetas y, por sus características, favorece el descubrimiento de planetas de gran masa y con semiejes pequeños. De hecho, más del 50 % de los exoplanetas tiene un período orbital inferior a un año y una masa superior a 0,1 MJ . Sin embargo, si bien las técnicas utilizadas favorecen este tipo de descubrimientos, fue una gran sorpresa que sistemas planetarios tan distintos al Sistema Solar existieran. Otros datos que despertaron el asombro son el amplio abanico de excentricidades (alrededor del 80 % de los exoplanetas tiene una excentricidad superior a 0,1), la presencia de planetas en estrellas binarias y la alta metalicidad de muchas de las estrellas con planetas. Al momento, alrededor del 60 % de las estrellas que albergan planetas muestran una metalicidad mayor a la solar. La detección de planetas entorno a estrellas de metalicidad alta provocó que en los subsiguientes programas de búsqueda de exoplanetas se favorecie- ran ciertos blancos (Santos et al. 2005, Wuchterl 2008). Por otra parte, por las dificultades intrínsecas de los métodos de detección se suelen descartar cierto tipo de estrellas, como aquellas que presentan altas velocidades de rotación, alta actividad cromosférica, estrellas sin una clase de luminosidad bien definida, etc. Existe además un sesgo en la búsqueda espectroscópica que favorece la detección de planetas entorno de estrellas más metálicas, ya que estas presentan líneas de absorción metálicas muy fuertes. La variación en la velocidad aumenta desde algunos m seg−1 para estrellas de metalicidad solar hasta 16 m seg−1 para estrellas poco metálicas (1/4 o menos). Este hecho podría llevar a que la frecuencia de exoplanetas en relación a la metalicidad de las estrellas que los albergan esté sobrestimada en hasta un factor 2 (Durisen et al. 2007). Todo esto sugiere que la correlación entre la metalicidad de las estrellas y la frecuencia de planetas deba, por ahora, ser analizada con cautela. Es notable, además, que en los últimos dos años el número de descubrimientos de exoplanetas entorno a estrellas más y menos metálicas que el Sol se ha ido balanceando. Pa- recería, entonces, necesario esperar a tener muestras estadísticamente más representativas de los sistemas extrasolares antes de precipitar conclusiones respecto de las características que presentan los sistemas planetarios en general. Luego del descubrimiento de los primeros exoplanetas, y debido a que estos resultaron ser muy diferentes de los planetas del Sistema Solar, se pensó que el Sistema Solar era un caso atípico y no representativo de un sistema planetario “normal”. Ahora bien, estas conclusiones suenan al menos un poco prematuras si tenemos en cuenta lo mencionado en los párrafos anteriores respecto de los efectos de selección que son intrínsecos a la disponibilidad tecnológica actual. Además, no hay que perder de vista que el Sistema Solar es el sistema planetario más estudiado y mejor conocido por los científicos. Y que el acceso que se tiene a diversos datos, como ser su estructura, su composición, la distribución de 28 masa y de momento angular, la información detallada del Sol, las propiedades superficiales de los planetas, los datos que aportan los cometas, asteroides y satélites; en fin, toneladas de información que la humanidad lleva siglos recopilando y analizando, y que demorará mucho tiempo, si es que alguna vez es posible, obtenerla con este nivel de detalle para sistemas extrasolares. Es así que la sugerencia de la violación del principio copernicano aplicado a la cosmogonía, que ha surgido en debates frente a la aparición de tan inesperados sistemas extrasolares, parece, como mínimo, muy apresurada. Es innegable la existencia de sistemas planetarios muy diferentes al nuestro, que presentan condiciones extremas que no eran imaginables hace 20 años atrás. Pero entonces, una teoría general de formación y evolución de los sistemas planetarios tiene que poder explicar en forma autoconsistente las diversas configuraciones observadas, incluida la local. Y éste es, sin duda, el desafío más grande de las Ciencias Planetarias. 2.3. Teorías de migración La presencia de exoplanetas en órbitas muy cercanas a la estrella central provocó cons- ternación en la comunidad astronómica ¿Pudieron estos objetos formarse in situ? ¿Pueden haber sobrevivido en esa ubicación el tiempo estimado de vida de su estrella? ¿Qué otras alternativas hay para explicar la existencia de estos sistemas? Poco tiempo después del descubrimiento de 51 Peg b (Mayor & Queloz 1995), Guillot et al. (1996) demostraron que el planeta pudo, efectivamente, permanecer tan cerca de su estrella durante todo su tiempo estimado de vida. Por su parte, Wuchterl (1996) probó que este planeta se pudo haber formado in situ si en su zona de alimentación había suficiente material disponible para ser acretado. Pero, a su vez, otras alternativas fueron propuestas. Basadas en la idea generalizada de que los planetas gigantes deben formarse más allá de la línea de hielo (ver 4.1), puesto que en la región interior no habría la cantidad de sólidos necesarios para construir planetas gigantes, estas teorías sugieren que debe haber algún mecanismo capaz de llevar a los protoplanetas desde órbitas exteriores a órbitas muy cercanas a la estrella como las observadas (Wuchterl 2008). Los mecanismos propuestos para el cambio en los elementos orbitales fueron dos: jumping Jupiters, para lo cual se necesita de un sistema planetario que en sus orígenes haya contado con varios planetas gigantes, migración orbital, donde se propone un decrecimiento gradual del radio orbital debido a la interacción del protoplaneta con el disco de gas y de planetesimales. La primera alternativa, propuesta por Weidenschilling & Marzari (1996), necesita de la formación de varios planetas gigantes, en un corto lapso de tiempo que permita que se generen perturbaciones destructivas que den por resultado la supervivencia de solo algunos de ellos en órbitas excéntricas y cercanas a la estrella central. En estos casos, los planetas de 29 menos masa son expulsados fuera del sistema. Más allá de que se requiere de configuraciones muy particulares para que esto ocurra, las distancias orbitales resultantes serían todavía mayores a las observadas, por cuanto se requeriría de algún otro mecanismo para continuar el acercamiento entre la estrella y los planetas. La migración orbital, por su parte, ha sido favorecida por muchos especialistas. La idea básica que engloba este concepto es el cambio sistemático del semieje mayor orbital de un planeta (Papaloizou et al. 2007). Históricamente, la teoría de migración estuvo asociada al desplazamiento hacia afuera de Urano y Neptuno desde su supuesto lugar de formación hasta su ubicación actual, lo cual dio una alternativa para acortar los tiempos de crecimiento de estos planetas. La migración se habría producido por un intercambio de momento angular entre los planetas y el disco de planetesimales remanentes. Después del descubrimiento de los primeros exoplanetas surgieron otras teorías de naturaleza muy distinta que se conocen como migración de tipo I, de tipo II y de tipo III, las dos primeras sugeridas por Ward en 1997 y la tercera así bautizada por analogía. En estos casos, la migración sería consecuencia del intercambio de momento angular entre el planeta y el disco de gas, siendo importante tanto para los planetas terrestres como para los gigantes. Esta propuesta ofrece una alternativa para explicar la existencia de los “hot Jupiters”. Para un planeta inmerso en un disco de gas, en órbita circular entorno a una estrella, el intercambio de momento angular entre el planeta y el gas ocurre en las ubicaciones de las resonancias de corrotación y de Lindblad (Armitage 2007). De su interacción con las resonancias de Lindblad interiores el planeta gana momento angular, lo cual hace que se desplace hacia afuera. Por su parte, el gas pierde momento angular y se mueve hacia adentro. De la interacción con las resonancias de Lindblad exteriores, el planeta pierde momento angular (lo que hace que se desplace hacia adentro) y el gas, que gana momento angular, se mueve hacia afuera. De este modo, el planeta repele al gas entorno a su órbita. El flujo de momento angular intercambiado en cada resonancia es proporcional a la masa del planeta al cuadrado y a la densidad superficial de gas en ese punto. Cuando la masa del planeta es ba ja, digamos de alrededor de 1 M⊕ aunque esto depende de las propiedades del disco, el flujo de momento angular es despreciable comparado con el transporte de momento angular por la viscosidad del disco, con lo cual el perfil de densidad de gas no se altera, y el torque neto que actúa sobre el planeta es la suma de los torques de las resonancias. En este caso la escala de tiempo de migración va con el inverso de la masa del planeta con lo cual, mientras se mantenga válida la hipótesis de que el perfil de gas no cambia, los planetas más grandes son los más afectados. El sentido en el cual se produce la migración no resulta evidente de la suma de los torques. Sin embargo, se encuentra invariablemente que las resonancias de Lindblad exteriores son más importantes que las interiores provocando que el planeta migre hacia el interior. El torque en la resonancia de corrotación no invierte el sentido de la migración. En una primera aproximación, la eficiencia del transporte de momento angular tiene poco impacto en la estimación de la tasa de migración de tipo I. Sin embargo, cuando se hacen simulaciones más realistas, considerando discos turbulentos, se producen fluctuacio- nes en la densidad del disco protoplanetario, las cuales ejercen torques aleatorios sobre el 30 planeta. Esto provoca que la variación del semieje no sea en un único sentido, producién- dose una migración estocástica o random walk. Simulaciones recientes muestran que para planetas de masa menor a 10 M⊕ domina la componente de random walk sobre la migración de tipo I “clásica”. Si consideramos planetas de mucha masa, el flujo de momento angular domina localmente al debido a la viscosidad. En este caso se produce la “repulsión” entre el planeta y el gas que lo rodea generándose una brecha de ba ja densidad entre ambos. Para que se forme la brecha se deben satisfacer dos condiciones: por un lado, el radio de Hill del planeta debe ser comparable a la altura de escala del disco; por el otro, los torques deben ser capaces de remover el gas más rápido de lo que los efectos viscosos tardan en llenarlo. Un planeta de la masa de Saturno estaría en el límite para la apertura de la brecha. Cuan- do un planeta tiene la masa necesaria para abrir una brecha, su evolución orbital ocurre en una escala de tiempo comparable a la del disco protoplanetario. Esta es la migración de tipo II. Para el caso de un Júpiter, el tiempo característico de migración sería de 0,75 millones de años aproximadamente. La migración de tipo III se refiere a una variación violenta del semieje del protoplaneta por inestabilidades en la interacción entre el planeta y el disco. Este tipo de migración surge de la transferencia de material a través de la resonancia de corrotación. Este modo de migración se manifestaría en aquellos planetas que hayan vaciado su región de corrotación y que se encuentren inmersos en discos masivos. La migración puede ser hacia adentro o hacia afuera dependiendo de las condiciones iniciales. En algunas simulaciones se vio que el semieje del planeta puede variar considerablemente en menos de 100 órbitas. Hasta el momento no está claro cuánto tiempo puede operar este mecanismo ni de que manera un planeta que sufre esta migración puede ser frenado. El estudio de la interacción planeta disco lejos está de ser sencillo. El estudio analítico de Ward (1997) toma como modelo las ecuaciones de la mecánica de fluidos linealizadas, una ley de potencias para la densidad nebular y un cierto potencial gravitatorio que refleje el problema. Esta aproximación, sin embargo, no tiene en cuenta la perturbación en la densidad del disco como consecuencia de la presencia del protoplaneta (Wuchterl 2008). Las partes más densas del protoplaneta, en la región más interior de su esfera de Hill, difieren enormemente del resto del disco. De este modo, la presión en el interior de la esfera de Hill estaría siendo altamente subestimada. El protoplaneta entra entonces en los cálculos como una fuente gravitatoria pero la presión que ejerce el gas en el sentido inverso está siendo omitida. Otra cuestión es la consideración, necesaria para un análisis lineal, de un disco kepleriano sin perturbar. Sin embargo, la mera presencia de un cuerpo masivo hace que la idea de un disco kepleriano sea, en estos casos, artificial. Por su parte, los cálculos hidrodinámicos bidimensionales (2D) y tridimensionales (3D) son terriblemente costosos y requieren, a su vez, de numerosas simplificaciones para funcionar. Sin embargo, son este tipo de simulaciones las que pueden aportar una aproximación más realista de cómo es la interacción planeta-disco. Masset, D’Angelo & Kley (2006) estudiaron la migración de tipo I con simulaciones tridimensionales para un disco isotermo y mostraron que los efectos no lineales se hacen importantes cuando el objeto tiene una masa de, aproximadamente, 5 M⊕ , reduciendo el torque neto, lo que lleva a tasas de migración más lentas. Debemos notar 31 que, en modelos unidimensionales de protoplanetas en discos isotermos (Pečnik & Wuchterl 2007) la masa de cruce5 es de apenas 0,1 M⊕ , un factor 100 por deba jo del valor estimado para situaciones más realistas. Entonces, los regímenes típicos que se estudian en modelos tridimensionales son supercríticos en un factor 100. Por ejemplo, la interacción dinámica planeta-disco, la acreción y la migración calculada en un disco isotermo para un objeto de 5 M⊕ de algún modo se corresponden con un objeto de 500 M⊕ (aproximadamente 1 1 2 MJ ). En realidad, escalear correctamente ambos casos no es trivial y requiere relacionar ambos cálculos con mucho cuidado. Por otra parte, la dificultad intrínseca de los modelos 3D limita todavía la realización de cálculos con diversas condiciones iniciales que permitan una comparación directa con los modelos unidimensionales (1D). Sin embargo, seguramente estas simulaciones serán posibles en un futuro cercano, arro jando luz sobre valores más realistas de la tasa de migración, como también sobre otras cuestiones de la formación de los planetas gigantes que hasta ahora solo se han podido estudiar en forma lineal. 5Masa de cruce: corresponde a la masa del núcleo (o de la envoltura de gas) de un planeta cuando, en el proceso de formación, la masa del núcleo es igual a la de la envoltura. (Ver capítulo 6, Ec. (6.8)). 32 Capítulo 3 Escenarios de formación 3.1. Formación de un sistema planetario 3.1.1. Nubes moleculares gigantes: la nursery Las nubes moleculares son las progenitoras de las estrellas, y por consiguiente, de los sistemas planetarios. Estas nubes están formadas fundamentalmente por moléculas de hi- drógeno, aunque también se observan CO, 13CO y NH3 (por ejemplo, Armitage 2007). Algunas de ellas se ven en el cielo como regiones oscuras; ésto se debe a que el polvo que contienen oscurece la luz de las estrellas de la Galaxia que se encuentran detrás de ellas. Generalmente, dentro de las nubes moleculares se detectan estrellas muy jóvenes (de pocos millones de años de edad), las cuales corresponden a objetos de pre-secuencia principal, lo cual significa que están aún en contracción y que todavía no alcanzaron las condiciones necesarias como para comenzar la fusión de hidrógeno. La formación estelar ocurre en los brazos espirales de la Galaxia (por ejemplo, Wuch- terl 2008). Las nubes moleculares se fragmentan en subestructuras que a su vez colapsan para dar lugar a las estrellas. El resultado de la fragmentación de la nube y del colapso de sus componentes es un cúmulo estelar. El proceso de formación estelar involucra pa- sar por múltiples escalas tanto de tamaño como de masa. La masa típica de una nube molecular es de alrededor de 106 M⊙ y su extensión de 100 pc, siendo su velocidad de rotación muy ba ja. Por su parte, los fragmentos que dan lugar a las estrellas tienen un tamaño de aproximadamente 0,1 pc y masas típicamente estelares de, aproximadamen- te, 1 M⊙ . La fragmentación de la nube tiene su origen en las mareas provocadas por la Galaxia, los campos magnéticos y la turbulencia, los cuales juegan un rol fundamental para que se formen “grumos” de alta densidad que luego colapsan para dar origen a las estrellas. Sin embargo, la física que domina a los fragmentos no es la misma que provoca los fragmentos, y la formación estelar estará gobernada por la competencia entre la fuerza de gravedad y la presión. En las primeras etapas de colapso la velocidad de rotación es dinámicamente despreciable. Sin embargo, el momento angular de cada grumo es grande. 33 De este modo, durante el colapso de los fragmentos, la rotación se vuelve cada vez más importante. Mientras se produce la contracción, la conservación del momento angular lleva a que estas estructuras entren en una rotación rápida. Sin embargo, en el caso del Siste- ma Solar el momento angular actual es mucho menor que el momento angular estimado para el estado inicial del colapso. Esto se conoce como “el problema del momento angular durante la formación estelar” (ver Armitage 2007). La solución a este problema sería que probablemente exista también una redistribución del momento angular entre las distintas componentes que surgen del fragmento (sistemas binarios o múltiples, formación de discos circumestelares, pérdida de momento angular en forma de jets), lo cual a su vez impediría que cada uno de los fragmentos alcance las velocidades de ruptura. Durante el colapso de una nube se pueden distinguir dos etapas: la fragmentación, y el posterior colapso y acreción de los fragmentos. La fragmentación de la nube daría origen a núcleos de masas superiores a 0,01-0,007 M⊙ . Por deba jo de este límite la fragmentación es poco probable ya que, en ese caso, los fragmentos serían tan densos que se volverían lo suficientemente opacos como para impedir el transporte de energía, llevando a que la temperatura aumente sin límites. Los fragmentos pueden caracterizarse básicamente de la siguiente manera: son cuasi-homogéneos, fríos (T ∼ 10 K), opacos para las longitudes de onda correspondientes al visible pero transparentes para aquellas encargadas del transporte de energía y, consecuentemente, isotermos. 3.1.2. Formación del disco protoplanetario El colapso gravitacional de una nube protoestelar se detiene cuando su núcleo se vuelve opaco y es entonces capaz de calentarse lo suficiente como para que la presión contrarreste a la gravedad (Wuchterl 2008). Dado que tanto la presión como la gravedad actúan en forma isotrópica, las estrellas resultantes son esféricas. Sin embargo, dado que el momento angular no es nulo, hay un segundo agente que puede llevar a detener el colapso: la fuerza centrífuga. Durante el colapso de la nube, la conservación del momento angular lleva a que la velocidad de rotación aumente en varios órdenes de magnitud. A diferencia de la presión, la fuerza centrífuga no es isotrópica y actúa en forma perpendicular al eje de rotación. De este modo el colapso, que se produce paralelo al eje de rotación, queda prácticamente inalterado. A partir del momento en que la nube alcanza un determinado radio, la fuerza centrífuga se vuelve lo suficientemente importante como para que el gas que caiga hacia el centro lo haga en forma paralela al eje de rotación. El material llega entonces al plano ecuatorial de la estrella, donde la componente vertical de la fuerza gravitatoria es cero y la componente radial se equilibra con la fuerza centrífuga. Se forma así un disco circumestelar en el plano ecuatorial. El ancho del disco queda determinado por el equilibrio entre la presión del gas y la componente vertical de la fuerza gravitatoria que ejerce la estrella. El balance de fuerzas en la dirección radial se da entre la fuerza gravitatoria y la fuerza centrífuga más el gradiente de presión del gas. Luego, a diferencia de un problema de dos cuerpos donde solo actúa la gravedad, en este caso el balance de fuerzas involucra también a la presión del gas. 34 Entonces, que exista una fuerza debida a la presión del gas es lo que hace que la velocidad de rotación del disco sea un poco más ba ja que en el modelo de un disco kepleriano. El resultado global del colapso de una condensación de gas es una protoestrella con un disco de acreción, el cual eventualmente puede convertirse en un disco protoplanetario. Es importante notar que tanto a nivel observacional como teórico es poco lo que se sabe respecto de los discos protoplanetarios y su origen. En el primer caso el problema radica en que la sensibilidad y la resolución necesaria para este tipo de estudios debe ser muy grande. En cuanto a los estudios teóricos, existen numerosas dificultades a nivel computacional para calcular un disco protoplanetario como resultado del colapso de una nube, y además el transporte de momento angular es algo que todavía no está satisfactoriamente resuelto. A grandes rasgos, la información que se tiene de las observaciones muestra que la mayo- ría de las estrellas jóvenes presentan un disco circumestelar cuya masa máxima no superaría el 10 % de la masa de la estrella central. Los discos observados tienen una extensión que en general no supera las 1.000 UA (Millan-Gabet et al. 2007). Debido a los límites actuales de resolución, la información sobre propiedades físicas como la temperatura y la densidad, solo pudo obtenerse a partir de las 50 UA en el caso de estrellas cercanas (Wuchterl 2008). Para radios menores a las 50 UA, el disco no puede ser resuelto espacialmente, y su estructura puede inferirse al combinar la distribución espectral de energía con los modelos teóricos. Además, las observaciones muestran una disminución en la emisión infrarro ja de discos con diferencias de edad de algunos millones de años, lo cual indicaría que se encuentran en distintos estadios evolutivos. Es probable que en aquellos discos con ba ja emisión esté ocurriendo la formación planetaria o que los propios discos se encuentren en proceso de disipación. 3.1.3. Del polvo a los planetesimales El escenario estándar de la formación de un sistema planetario se puede pensar en cuatro etapas (Nagasawa et al. 2007). Primero, el polvo sedimenta hacia el plano de la nebulosa protoplanetaria formando un disco. Luego, la formación de un sistema planetario comprende las tres etapas restantes, que están caracterizadas por la aparición en el disco de: 1) los planetesimales, 2) los protoplanetas, y finalmente 3) los planetas. El disco circu- mestelar está formado originalmente por gas y polvo. La formación de los planetas requiere entonces del crecimiento de las partículas de polvo en, al menos, 12 órdenes de magnitud (Armitage 2007). Se considera polvo a las partículas con tamaños por deba jo del micrón y hasta aproximadamente el centímetro. El movimiento de estas partículas está acoplado al del gas y su crecimiento ocurre por colisiones inelásticas. A medida que estas partículas van aumentando su tamaño se van desacoplando dinámicamente del gas, estando su movimien- to bien descripto como una combinación entre la correspondiente a una órbita kepleriana y la fricción aerodinámica producto de que estos objetos se desplazan más rápido que el gas que los rodea. El crecimiento de las partículas en este régimen debería ser rápido, aunque todavía no está bien entendido. Los planetesimales son cuerpos del orden del kilómetro, su- 35 ficientemente masivos como para que se los pueda considerar dinámicamente desacoplados del gas. Los planetesimales son los cuerpos más pequeños que se tienen en cuenta cuando se estudia la formación de los planetas dado que sus interacciones pueden ser considera- das puramente gravitatorias. Cuando la acreción de planetesimales da lugar a cuerpos de tamaño terrestre, estos vuelven a acoplarse al disco de gas, pero en este caso predomina la interacción gravitatoria: aparecen los fenómenos de migración (ver Capítulo 2, § 2.3). Su subsecuente crecimiento dará lugar a los núcleos de los planetas gaseosos: cuando un embrión alcanza aproximadamente 10 M⊕ es capaz de acretar grandes cantidades del gas circundante dando origen a las extensas atmósferas de los planetas gigantes (Mizuno 1980, Stevenson 1982). En síntesis, los primeros cuerpos masivos en aparecer en el disco son los planetesimales, que se forman por la coagulación de granos de polvo. Luego, los protoplanetas se forman por acreción de planetesimales. Finalmente, la tercera etapa es la que distingue la forma- ción de los planetas terrestres de los planetas gigantes: mientras que los primeros surgen de las colisiones de protoplanetas, los segundos son protoplanetas capaces de acretar grandes cantidades de gas de la nebulosa. Este escenario, conocido como escenario estándar, es el más aceptado actualmente, pero hay que remarcar que su plausibilidad está ligada a la for- mación de los planetesimales, hecho que todavía no pudo ser explicado satisfactoriamente. En el proceso de formación de un sistema planetario hay dos cuestiones que no están bien entendidas: la formación de los planetesimales y la formación de los planetas gigantes. Esta Tesis tiene como objetivo estudiar lo segundo, dando por sentado la existencia de los planetesimales. De todos modos, en esta sección daremos un breve resumen de lo que se sabe hasta el momento del proceso de formación de los planetesimales. Lo que sigue está basado en el curso dictado por A. Youdin durante la escuela de invierno Les Houches Winter School 2008. La estimación del tiempo de formación de los planetesimales puede hacerse de dos ma- neras completamente independientes. Por un lado, una cota superior proviene de la escala de tiempo de vida de los discos observados alrededor de las estrellas T Tauri. Por otra par- te, se pueden hacer estimaciones a partir del estudio de los meteoritos. Afortunadamente, ambos procedimientos arro jan resultados concordantes. Dado que los planetesimales se forman en regiones donde hay abundante polvo, es importante estimar la vida media de los discos óptimamente gruesos. Las observaciones de objetos estelares jóvenes en el infrarro jo cercano revelan la existencia de un disco interior que se encuentra a altas temperaturas. Si bien la parte gaseosa del disco es más difícil de estimar que su componente de polvo, existe gran evidencia de que las zonas de altas temperaturas en el disco correlacionan con abundantes cantidades de gas. De este hecho surge que en el cálculo de la formación de los planetesimales se debe tener en cuenta la presencia del gas, que implica que no se puede despreciar el frenado gaseoso. Uno de los motivos por los cuales los planetesimales deben formarse mientras todavía hay gas en el disco, tiene que ver con que los planetas gigantes deben a su vez formarse mientras el gas esté presente, y como dijimos, los planetesimales son los bloques fundamentales para 36 la construcción de un planeta. Dadas las estimaciones actuales de la vida media de los discos circumestelares, este proceso no puede durar más que algunos millones de años. Se concluye entonces que los planetesimales deben formarse durante los primeros millones de años de evolución de la estrella en la pre-secuencia principal, cuando los discos tienen todavía abundante cantidad de polvo. Sin embargo, dado que las partículas centimétricas no emiten radiación en forma eficiente, no se puede observar a los planetesimales durante su proceso de formación. Por otra parte, de los meteoritos del Sistema Solar se puede obtener información en ciertos aspectos más detallada. Los meteoritos primitivos se conocen como condritas (no están diferenciados en un núcleo rico en hierro y un manto rocoso como los planetas y los asteroides más grandes) y se mantienen relativamente inalterados desde su origen. Las condritas están formadas mayoritariamente por condrulos. Sin embargo, algunas presentan inclusiones de calcio y aluminio, las cuales son altamente refractarias. Se conoce a estas inclusiones como CAIs (Calcium Aluminun Inclusions) y representan el material más anti- guo conocido del Sistema Solar, generalmente utilizado para datar su edad. Si las condritas forman parte de la primera generación de planetesimales, la diferencia de edad entre las condritas y las CAIs da un límite inferior a la escala de tiempo de formación de los me- teoritos primitivos. Del decaimiento radiactivo del 26Al se puede estimar que las CAIs son más viejas que los condrulos por algunos millones de años. Esto significa que la formación de los planetesimales en el Sistema Solar fue un proceso que se extendió por varios millones de años, en acuerdo con las estimaciones de vida de los discos protoplanetarios. Sin embargo, las partículas están inmersas en un disco de gas que no rota con velocidad kepleriana, lo cual a efectos prácticos resulta en un “viento” que las frena y las hace espiralar hacia la estrella central. La escala de tiempo involucrada en este proceso para objetos del tamaño del metro es tan solo de ¡100 años!. Luego, los planetesimales en crecimiento deberían superar este tamaño crítico con mucha rapidez o, alternativamente, debería existir algún mecanismo que retenga al material y le impida caer hacia la estrella central. Actualmente existen dos mecanismos propuestos para la formación de los planetesima- les: (1) colisiones sucesivas entre las partículas de polvo que terminen en la aglomeración de las mismas y (2) el colapso gravitatorio de pequeñas partículas sólidas en cuerpos más grandes. Estos dos mecanismos no son mutuamente excluyentes ya que podría ocurrir que las partículas de polvo crezcan primero como producto de colisiones inelásticas hasta que los cuerpos que formen se desacoplen del gas y luego colapsen como consecuencia de su autogravedad dando como resultado la aparición de los planetesimales. Las colisiones entre partículas podrían llevar a la formación rápida de los planetesimales si el “pegoteo” entre las partículas fuera altamente efectivo. La eficiencia de la acreción por colisiones inelásticas no se conoce bien todavía, aunque actualmente se están llevando a cabo numerosos ensayos de laboratorio para ver ba jo qué condiciones estas colisiones provocan el crecimiento de las partículas. Las velocidades relativas entre las partículas en el momento de una colisión determinan si la colisión dará lugar a acumulación o no. La velocidad de colisión crece monótonamente con el tamaño de las partículas hasta que el 37 desacoplamiento entre ambas cantidades ocurre para cuerpos del orden del metro, cuando las velocidades relativas alcanzan los 100 m seg−1 . Sin embargo, los experimentos muestran que partículas del tamaño del milímetro tendrían ya dificultad en aglomerarse en forma efectiva, independientemente de las velocidades involucradas. Por otra parte, la idea del crecimiento de un planetesimal por colisiones con partículas pequeñas parece también difícil debido a las altas velocidades a las cuales estos eventos deberían producirse. La otra posibilidad para la formación de los planetesimales es la inestabilidad gravitato- ria. Este mecanismo es responsable de la formación de muchos de los objetos del Universo, en las más diversas escalas. En el caso de los planetesimales, la idea es que estos se for- marían en el disco protoplanetario, donde los sólidos forman un disco frío (debido a que pierden rápidamente su energía cinética por el frenado gaseoso y las colisiones inelásticas), el cual se fragmentaría debido a su autogravedad y provocaría de esta forma el colapso de pequeñas partículas sólidas en planetesimales. El problema de este modelo es la presencia de turbulencia. En ausencia de turbulencia, las partículas se acumularían en el plano cen- tral del disco hasta llegar a densidades suficientemente altas como para que se produzca el colapso gravitatorio. Pero aún el propio decantamiento de los sólidos provoca turbulen- cia en el disco, con lo cual ésta no puede ser despreciada. Recientemente, Johansen et al. (2007), hicieron simulaciones del colapso gravitatorio de los sólidos con un código híbrido MHD/partícula-malla donde se incluyeron diversas fuentes de turbulencia en el disco. En estas simulaciones las partículas se consideraron con un tamaño inicial de entre 15 y 60 cm. Si bien es incierto todavía cómo alcanzarían este tamaño, las simulaciones muestran la viabilidad de la formación de planetesimales por inestabilidad gravitatoria, ya que las partículas colapsan rápidamente hasta formar objetos del tamaño del kilómetro en tan solo un período orbital. La formación de planetesimales presenta muchos interrogantes y quedan todavía serias cuestiones por resolver. Los recientes resultados de simulaciones numéricas muestran que el colapso gravitatorio es una solución posible, aunque el crecimiento de los granos de polvo hasta el tamaño del centímetro ocurra mediante mecanismos por ahora desconocidos. Las capacidades tecnológicas actuales no permiten tener registros observacionales del proceso de formación de los planetesimales, quedando limitada la información disponible a los registros obtenidos de estrellas T Tauri y a aquella que aportan lo meteoritos encontrados en La Tierra. 3.1.4. Planetas rocosos y planetas gaseosos Los protoplanetas se forman por acreción entre los planetesimales. Clásicamente, se distinguen dos regímenes de crecimiento entre los planetesimales: el ordenado y el runaway. En el crecimiento ordenado todos los cuerpos que conforman la población de planetesimales tienen aproximadamente la misma tasa de crecimiento, y la razón entre sus masas resulta ser del orden de la unidad. En cambio, en el crecimiento runaway los objetos más grandes crecen más rápidamente que los pequeños, con lo cual la razón entre sus masas crece 38 monótonamente. El escenario clásico de la formación planetaria, planteado en forma teórica por Safro- nov en la década del ’60, asumía que los embriones planetarios surgían por el crecimiento ordenado de los planetesimales. Sin embargo, si esto fuera así, los núcleos de los planetas gigantes no se podrían formar lo suficientemente rápido como para capturar el gas necesario para su envoltura antes que la nebulosa se disipe (Safronov 1969). A partir de la década del ’80, el crecimiento de los planetesimales comenzó a estudiarse en más detalle a través de simulaciones numéricas. Estas simulaciones plantearon un escenario diferente, ya que mos- traron que si un objeto del disco de planetesimales se distingue de los demás por ser más masivo que la media seguirá un crecimiento runaway y no un crecimiento ordenado como se pensaba antes (Kokubo & Ida 1998, 2000, 2002). En el crecimiento runaway, el planete- simal que, por cuestiones probabilísticas, surge más masivo que el resto continúa acretando planetesimales de su vecindad, aumentando así su masa, diferenciándose cada vez más del resto de los planetesimales del disco e impidiendo que éstos crezcan en forma significativa. Es destacable, sin embargo, que la mayor parte de la masa del sistema continúa contenida en los planetesimales más pequeños. El crecimiento runaway no implica que el tiempo de crecimiento de un planetesimal decrezca con la masa sino que, la razón entre la masa del cuerpo en crecimiento runaway y la del resto de los planetesimales crece con el tiempo. Una vez que el embrión alcanza la masa necesaria como para excitar gravitatoriamente a los planetesimales que lo rodean, el crecimiento runaway se autolimita. Los embriones resultantes tienen, aproximadamente, la masa de la Luna y se encuentran separados por una distancia de ∼ 10 RH (con RH nos referimos al radio de Hill de un objeto1). Estos embriones continúan creciendo, pero no ya dentro de un régimen runaway sino siguiendo lo que se conoce como crecimiento oligárquico, ya que solo estos objetos siguen acretando planetesimales. Entonces, el crecimiento oligárquico es el régimen que corresponde a los embriones que surgieron por el crecimiento runaway de los planetesimales y que siguen su crecimiento alimentándose de los planetesimales que pueblan el disco pero que se vie- ron impedidos de crecer. La tasa de crecimiento correspondiente al régimen oligárquico es menor que durante el crecimiento runaway, pero mayor que en un crecimiento ordenado. El resultado del crecimiento oligárquico es un sistema protoplanetario: el crecimiento oligárquico es el que, de alguna manera, establece cuáles son las configuraciones orbitales finales, como también las masas resultantes y las escalas de tiempo involucradas (Kokubo 2001). En la tabla 3.1 se muestran los resultados de simulaciones numéricas realizadas para distintos radios orbitales en el Sistema Solar. Llegado a este estado, la fase final de la formación planetaria ocurre dependiendo de la región del disco donde se encuentran los protoplanetas. En la región de los planetas terrestres, la masa final y la separación orbital estimadas de los embriones resultantes es mucho más pequeña que la de los planetas ac- tuales. Luego, para alcanzar las masas planetarias el proceso de acreción debe continuar con colisiones entre los protoplanetas. El sistema protoplanetario resultante del crecimien- to oligárquico se puede volver dinámicamente inestable en escalas de tiempo largas. Las 1Ver definición del radio de Hill en la sección “Símbolos y Unidades” o en el capítulo 4, ecuación (4.13). 39 excentricidades aumentarían entonces por perturbaciones gravitatorias de otros protopla- netas o de los planetas jovianos. De esta manera se daría el cruce orbital necesario para que se produzcan las colisiones entre los protoplanetas. Sin embargo, las simulaciones de N -cuerpos muestran que las excentricidades finales de los planetas resultarían mucho más elevadas que las actuales. Una posibilidad para ba jar las excentricidades provendría de la fricción dinámica de los planetesimales residuales en el disco; otra posibilidad serían las interacciones con el disco de gas. Por otra parte, para los embriones que se encuentran en la región de los planetas exteriores la abundancia de material gaseoso permite que la gravedad de los cuerpos sólidos atraiga al gas circundante del disco, formando de esta manera la característica estructura de los planetas gaseosos. 3.2. Teorías de formación de planetas gigantes Se han propuesto varios modelos para explicar la formación de los planetas gigantes. Entre otros argumentos, el hecho observacional de la muy ba ja presencia de objetos con masas entre 5 y 50 MJ en órbita alrededor de estrellas de tipo solar (conocido como “desierto de enanas marrones”), lo cual representa una discontinuidad en la función inicial de masa, sugiere que el mecanismo de formación de los planetas es diferente al de las estrellas. En la sección 1.3.1 mencionamos algunos escenarios de formación de enanas marrones, fundamentalmente aquellos que podrían explicar la existencia de estos objetos en forma aislada. Para el caso de las enanas marrones con compañeros de masa estelar, la escasa población de estos objetos con períodos menores a cinco años podría estar relacionada con las dificultades del modelo de inestabilidad del disco como escenario de formación de objetos subestelares, como veremos a continuación. Por otra parte, en el Sistema Solar, el hecho que los planetas terrestres estén formados fundamentalmente por sólidos y que, a medida que la masa de los planetas crece, aumenta su componente gaseosa pero evidenciando siempre Tabla 3.1. Masa final, mf , y tiempo requerido, tf , para el crecimiento oligárquico de embriones planetarios ubicados en las actuales posiciones de La Tierra, Júpiter, Saturno, Urano y Neptuno (Kokubo 2001). Σ0 es la densidad superficial de sólidos y ∆a la separación orbital de los embriones (∆a = 10 RH ). a [UA] Σ0 [g cm−2 ] ∆a [UA] mf [M⊕ ] 0,16 0,07 10 1 5 1 4 5 10 1,4 3 9 14 6 0,5 20 30 0,3 10 20 tf [106 años] 0,7 40 300 2000 7000 40 un enriquecimiento respecto de la abundancia solar, lleva a pensar también en un escenario de formación común para todos los planetas. Además, las propiedades observadas en otros sistemas planetarios, como por ejemplo el caso de HD 149026b que tendría un núcleo sólido de 67 M⊕ (Sato et al. 2005), sugiere que los exoplanetas se habrían formado en una manera análoga a la de los planetas del Sistema Solar. Actualmente, existen dos modelos que intentan explicar la formación de los planetas gigantes: el modelo de inestabilidad nucleada y el modelo de inestabilidad gravitatoria. Ambos modelos son conceptualmente diferentes. Si bien el modelo de inestabilidad nuclea- da es el que cuenta con mayor aceptación por parte de la comunidad científica en general (y es el modelo en el cual está basado el desarrollo de esta Tesis), la falta de argumen- tos irrefutables contra el modelo de inestabilidad gravitatoria hace que éste se encuentre todavía en consideración. 3.2.1. Modelo de inestabilidad nucleada Siguiendo a Lissauer & Stevenson (2007), el modelo de inestabilidad nucleada puede resumirse de la siguiente manera: Las partículas de polvo presentes en la nebulosa solar se acumulan para formar plane- tesimales, los cuales a su vez son acretados mutuamente dando origen al núcleo sólido del planeta (este proceso es completamente análogo al de la formación de los planetas terrestres). Mientras esto ocurre, el núcleo comienza a rodearse de una envoltura de gas que en principio es poco masiva. Por su parte, la tasa de acreción de sólidos es mucho más alta que la de gas. A medida que la zona de alimentación2 del planeta se va vaciando, la acreción de sólidos disminuye mientras que la de gas aumenta. En el transcurso del proceso de acreción, se llega a que, en algún momento, la masa del núcleo y la masa de la envoltura gaseosa se hacen iguales. Cuando esto ocurre, la tasa de acreción de gas cambia, haciéndose cada vez más alta. Esta etapa se conoce como crecimiento runaway de la envoltura, ya que el protoplaneta incorpora grandes cantidades de gas en un corto período de tiempo. Estas dos primeras etapas se conocen como estadío nebular ya que el borde externo del planeta permanece en contacto con la nebulosa protoplanetaria, con lo cual la densidad y la temperatura en la interfase son las correspondientes a la de la nebulosa. Estas dos etapas son las que se estudiarán en detalle en el presente traba jo. Siguiendo en el proceso de formación, se tiene que: El crecimiento runaway de la envoltura queda luego limitado a la tasa a la cual la nebulosa es capaz de transportar gas a la vecindad del planeta. Una vez alcanzado este estado, la región de la envoltura que se encuentra en equilibrio se contrae y continúa la acreción de gas en forma hidrodinámica. 2Región adyacente a la órbita del planeta de donde éste obtiene el material para su formación. 41 Finalmente, la acreción de gas termina, ya sea por la apertura de una brecha en el disco (gap) como consecuencia de los efectos de marea que provoca el planeta, o por la propia disipación de la nebulosa. Una vez finalizada la acreción, el planeta evoluciona aislado, contrayéndose y enfriándose a masa constante. El modelo de inestabilidad nucleada fue investigado por primera vez por Perri & Ca- meron (1974) y Cameron (1978) quienes, proponiendo una envoltura completamente adia- bática, estudiaron las soluciones de equilibrio para la envoltura. Más tarde, Mizuno et al. (1978), perfeccionaron el modelo aproximando la estructura de la envoltura por capas iso- termas y adiabáticas. Finalmente, en 1980, Mizuno publicó el traba jo que es considerado referente para los cálculos que se realizaron a posteriori. Mizuno (1980) estudió la formación de los planetas gigantes en base a modelos com- pletamente hidrostáticos y en equilibrio térmico. El procedimiento del que se valió fue la construcción de una secuencia de modelos donde se incrementaba la masa del núcleo con- forme transcurría el tiempo y, para cada modelo resolvía el estado de la masa de gas ligado. El límite externo de la envoltura correspondía al radio tidal (o radio de Hill) del objeto, en donde las condiciones de borde eran las que caracterizan a la nebulosa protoplanetaria para el dado radio orbital. Mizuno encontró que existía un valor máximo para la masa del núcleo a partir del cual no existía una solución estática para la envoltura. De acuerdo a sus cálculos, la “masa crítica” del núcleo a partir de la cual se perdía el estado de equilibrio hidrostático era de aproximadamente 12 M⊕ . Esto concordaba con las estimaciones para los núcleos de los planetas gigantes del Sistema Solar. Y dado que este valor resultaba ser insensible a la posición del planeta en la nebulosa, esto es, las condiciones de borde no afectaban notoriamente la masa final del núcleo, se daba así una explicación lógica para el hecho de que las masas de los núcleos de los planetas gigantes del Sistema Solar fueran similares. Seis años más tarde, Bodenheimer & Pollack (1986) construyeron el primer código capaz de resolver numéricamente y en forma autoconsistente las cuatro ecuaciones diferenciales características de la evolución estelar aplicadas a la envoltura del planeta (ver capítulo 4, § 4.4.1), proponiendo así un modelo más realista que el de Mizuno para estudiar la forma- ción de los planetas gigantes. Adoptando una tasa de acreción de sólidos constante para la formación del núcleo, la resolución de estas ecuaciones en conjunto con las condiciones de borde, dan lugar a una secuencia de modelos donde el crecimiento del núcleo y de la envoltura se calculan en forma acoplada y consistente. Una vez que el planeta llegaba a la masa final, Bodenheimer & Pollack estudiaron además la evolución del mismo durante su proceso de enfriamiento. En la primera fase de la formación del planeta, el núcleo crece a una tasa mucho más alta que a la que lo hace la envoltura. En esta etapa, la energía de la envoltura proviene principalmente de la energía gravitatoria de los planetesimales ingresantes. Cuando la masa de la envoltura se hace igual a la masa del núcleo, la energía proveniente de la acreción de sólidos se vuelve insuficiente para contrarrestar “el peso” de las capas de gas, y la luminosidad de la envoltura comienza a estar dominada por la contracción de las capas 42 de gas. Esta transición, si bien bastante rápida, se produce de manera suave y continua. El valor de la masa del núcleo para el cual ella ocurre puede identificarse con la “masa crítica” definida por Mizuno. Sin embargo, es importante notar que, a diferencia de lo que se pensaba con anterioridad, la envoltura nunca deja de estar en equilibrio hidrostático. El concepto de “masa crítica” puede reformularse como la masa del núcleo a partir de la cual el crecimiento del planeta se debe fundamentalmente a la acreción de gas. En la práctica, este punto se alcanza cuando el núcleo y la envoltura tienen masas aproximadamente iguales. Bodenheimer & Pollack encontraron que la masa del núcleo para la cual esto ocurría era de entre 10 y 30 M⊕ . El traba jo de Bodenheimer & Pollack (1986) fue pionero en lo que respecta al cálculo autoconsistente de la formación de planetas gigantes en el marco del modelo de inesta- bilidad nucleada. Si bien presenta numerosas simplificaciones (como considerar la tasa de acreción de sólidos constante), sus resultados continúan siendo conceptualmente válidos. El estudio de Bodenheimer & Pollack incorporó por primera vez, ba jo la hipótesis de un modelo unidimensional, toda la física relevante del problema. Probablemente uno de los resultados más importantes de este traba jo, más allá de la confirmación de la plausibilidad del modelo de inestabilidad nucleada, fue el hecho de que todo el proceso de acreción de gas ocurre con la envoltura en equilibrio hidrostático, aún después de alcanzar la “masa crítica” para el núcleo. Las simulaciones más detalladas que continuaron con el estudio de este problema confirmaron también este hecho. Cabe, sin embargo, mencionar que Wuch- terl (1990, 1991, 1995), con un tratamiento completamente hidrodinámico, encuentra que cuando el núcleo alcanza la masa crítica se desencadena una inestabilidad en la envoltura que provoca la eyección de grandes cantidades de gas. Mientras que estos eventos están presentes cuando la densidad nebular es ba ja (como podría atribuirse a las regiones de Urano y Neptuno), desaparecen para densidades más altas. Wuchterl propone que este mecanismo fue el que le impidió a los gigantes helados continuar la acreción de gas y por este motivo presentan envolturas menos masivas que Júpiter y Saturno. Las inestabilidades encontradas por Wuchterl no pudieron ser reproducidas por otros autores. Sin embargo, el tratamiento que Wuchterl hace del transporte de la energía (considerando que la convec- ción depende del tiempo), no fue tampoco incluido en ningún otro modelo de formación de planetas gigantes. 3.2.2. Modelo de inestabilidad del disco El modelo de inestabilidad del disco propone un escenario completamente distinto al de la inestabilidad nucleada para la formación de los planetas gigantes. Según este modelo, los planetas gigantes se forman por la contracción de “grumos” de gas que surgen debido a inestabilidades gravitatorias en el disco protoplanetario. Las simulaciones numéricas (Boss 1997, 1998) muestran que condensaciones de ∼ 1MJ se podrían formar en un disco de gas gravitatoriamente inestable. Sin embargo, debido al costo computacional que estas simulaciones requieren la evolución de estas condensaciones solo se siguen por algunos períodos orbitales y resulta incierto si ellas son capaces de sobrevivir lo suficiente como para 43 que luego contraerse hasta alcanzar el tamaño de un planeta. De hecho, las inestabilidades gravitatorias que provocan las condensaciones generan también ondas de densidad, las cuales transportan momento angular lo cual haría que el disco se ensanche y, por lo tanto, disminuya su densidad en la región. Esto provocaría que los grumos finalmente se desarmen y el disco vuelva a ser estable. Raficov (2005) mostró que los discos que podrían dar lugar a planetas gaseosos por inestabilidad gravitatoria, a una distancia de algunas decenas de UA tendrían que ser muy calientes (T > 103 K), luminosos y de gran masa, propiedades que, de ser frecuentes, serían fácilmente detectables en discos circumestelares. Para discos con propiedades más o menos estándar, la formación por inestabilidad gravitatoria podría producirse más allá de las 100 UA. En estos casos no solo habría que invocar mecanismos de migración para posicionarlos en las ubicaciones actuales sino que, además, las masas de estos objetos no podrían ser inferiores a las 10 MJ . Este hecho estaría en concordancia con el “desierto de enanas marrones”, dado que si el mecanismo de inestabilidad del disco fuera eficiente para semiejes pequeños deberían detectarse muchos objetos en el rango de masas de las enanas marrones. Sin embargo, la frecuencia de enanas marrones alrededor de estrellas de tipo solar aumenta con la separación orbital. Dado que, en cambio, sí se observan cuerpos de masa planetaria para pequeños semiejes, se considera a este hecho como un argumento a favor de que la formación de planetas gigante y enanas marrones correrían por caminos diferentes. Por otra parte, otro argumento en contra del modelo de inestabilidad del disco es que, para impedir que las condensaciones se desarmen ba jo circunstancias plausibles para el disco, se necesitaría que el disco se enfríe demasiado rápido y/o que haya acreción de masa, de modo que se mantenga inestable. Aún si esto fuera posible, la composición del planeta sería la misma de la nebulosa, con lo cual se necesitaría un proceso independiente para explicar el enriquecimiento respecto de la composición solar que tienen Júpiter y Saturno. Además, en este escenario la formación de planetas como Urano y Neptuno (que, de algún modo, se puede pensar que están a mitad de camino entre los planetas terrestres y los gigantes gaseosos) resulta bastante difícil de explicar. El estudio de la evolución de los grumos requiere que las simulaciones sean integra- das por largos períodos de tiempo, con lo cual habría que considerar a su vez los procesos dinámicos que ocurren en el disco. Por otra parte, la capacidad computacional actual es in- suficiente para hacer un seguimiento detallado de la evolución de estos grumos por un lapso de tiempo mayor a algunos miles de años. Además, si se quiere estudiar en profundidad la capacidad de supervivencia de los grumos es vital incluir en el modelo una descripción ter- modinámica más detallada de la que se ha estado empleando hasta ahora. En una reciente revisión sobre la formación de los planetas gigantes por inestabilidades gravitatorias en el disco, Durisen et al. (2007) plantean que en simulaciones donde la ecuación de estado con- siderada es isoterma, los grumos crecen hasta aproximadamente 10 MJ en algunos miles de años. Sin embargo, al usar la EOS del gas ideal (que no es isoterma) estos grumos apenas llegan a 1 MJ . Este tipo de resultados pone en evidencia la importancia de un tratamiento termodinámico adecuado para analizar tanto la supervivencia de los grumos como también su crecimiento. 44 3.2.3. ¿Un modelo híbrido? En los últimos años, varios autores han sugerido que la formación de planetas gigantes por inestabilidad nucleada podría verse favorecida si se considerara la presencia de inesta- bilidades gravitatorias en el disco de gas. La complejidad de este tipo de simulaciones hace que los resultados obtenidos hasta el momento estén lejos de ser concluyentes; sin embargo son sugestivos y proponen nuevos caminos para la investigación de este problema. En base a simulaciones 3D hidrodinámicas, Durisen et al. (2007) especulan con que en los discos protoplanetarios las inestabilidades gravitatorias lleven a que se produzcan, en una escala de tiempo de algunos miles de años, arcos o anillos donde la densidad del disco se ve incrementada en algunos factores respecto del valor original. Si bien no necesariamente estos anillos llevarían a la formación de planetas por inestabilidades gravitatorias en el gas, podrían acelerar el proceso de la inestabilidad nucleada. La formación de los anillos no solo implica un aumento en la densidad sino también un pico de presión en esa región. En la vecindad de regiones de alta presión la velocidad del gas puede ser super o sub- kepleriana dependiendo del gradiente de presión. Los planetesimales, que sufren la fricción producida por el gas, se desplazan hacia afuera o hacia adentro respectivamente, siendo el efecto neto una acumulación de sólidos en las regiones de alta presión. Esta acumulación podría llevar a la formación, más rápida y en gran cantidad, de planetesimales en esta zona y, consecuentemente, se aceleraría la formación del núcleo de los planetas gigantes, haciendo que todo el proceso sea más eficiente. Además, la formación de cuerpos sólidos por aglomeración de partículas de polvo disminuye la opacidad del gas, lo cual favorece la acreción y la contracción de las capas de gas. Si bien no se han realizado simulaciones que demuestren que realmente estas sean las consecuencias de la formación de anillos en el disco de gas, ponen de manifiesto que es necesario estudiar estos escenarios en más profundidad. Otro mecanismo “híbrido” de formación de planetas gigantes es el propuesto por Klahr & Bodenheimer (2003). La idea de estos autores es que el núcleo sólido se forme en el centro de un vórtice anticiclónico, dado que los vórtices colectan material del disco. Los vórtices se forman por perturbaciones que produce la materia cuando ésta cae hacia el disco o bien como resultado de inestabilidades hidrodinámicas. De simulaciones numéricas se obtiene que, una vez capturadas, las partículas de hasta el tamaño del metro, quedan dentro del vórtice por varios cientos de períodos orbitales, aún si el disco es turbulento. En el caso de cuerpos del orden del kilómetro o más grandes ya la situación no es tan clara. Es posible que los planetesimales sean también eyectados del vórtice, con lo cual sería poco probable que llegue a formarse un cuerpo lo suficientemente masivo como para ser considerado un embrión. Sin embargo, este mecanismo podría dar origen a objetos de hasta un kilómetro de tamaño que, como se mencionó anteriormente, es algo que todavía no está bien resuelto. Por otra parte, los planetesimales que se formen quedan radialmente confinados, lo cual facilita la acreción mutua para formar el núcleo, y un enriquecimiento en la zona de alimentación para que el proceso sea rápido. Con lo cual, la presencia de vórtices en el disco aceleraría la formación de un planeta, aún en discos de ba ja masa. Además, pone de manifiesto un mecanismo alternativo para la formación de planetesimales. El mayor inconveniente que 45 hasta el momento tiene esta hipótesis es que no existen observaciones que confirmen la formación de vórtices en los discos protoplanetarios. Está entre los objetivos de ALMA3 la búsqueda de estas estructuras. Es importante entonces que se continúe el estudio de los procesos de inestabilidad gravitatoria en los discos protoplanetarios ya que, si bien puede que éstos no den lugar a la formación directa de planetas, pueden favorecer la formación de los planetesimales y de los núcleos planetarios. De hecho, aún cuando no sea posible que se formen cuerpos de tamaños superiores al metro, las inestabilidades gravitatorias impedirían la migración abrupta de las partículas sólidas hacia la estrella central. La estructura compleja y variable en el tiempo de las inestabilidades gravitatorias aumentaría el tiempo de permanencia de las partículas en el disco dándoles el tiempo suficiente como para crecer hasta el tamaño necesario para desacoplarse del gas. 3Atacama Large Millimiter/submillimiter Array. Proyecto astronómico internacional que consiste en un conjunto de radio telescopios. Estudiará la formación estelar en el universo temprano e intentará detectar en forma directa planetas extrasolares. 46 Capítulo 4 Modelo teórico de formación de planetas gigantes En los capítulos anteriores se resumió el estado actual del conocimiento de los proce- sos involucrados en la formación de los planetas gigantes. También, se hizo referencia a los últimos datos observacionales disponibles, tanto en lo que respecta a los planetas del Sistema Solar como a los planetas extrasolares. En lo que sigue, se presentarán en detalle las componentes del modelo adoptado para la elaboración de esta Tesis sobre la formación de los planetas gigantes. Como ya mencionamos, el objetivo de esta Tesis es calcular la formación de planetas gigantes en el marco de la hipótesis de inestabilidad nucleada. Este estudio se hará utili- zando un código numérico, cuya versión original fue desarrollada por Benvenuto & Brunini (2005), y posteriormente actualizada por Fortier, Benvenuto & Brunini (2007). Para este tipo de cálculo deben elegirse modelos de crecimiento, tanto para el núcleo sólido como para la envoltura gaseosa. El estudio detallado del crecimiento de un embrión sólido requiere de simulaciones de N -cuerpos, lo cual está fuera del alcance de los objetivos de este traba jo. Sin embargo, de este tipo de simulaciones, y en conjunto con modelos semi-analíticos (Ida & Makino 1993, Kokubo & Ida 1996) pueden obtenerse expresiones analíticas aproximadas, las cuales nos permiten incorporar al código un modelo del régimen de acreción de sólidos. En cuanto a la formación de la envoltura gaseosa, ésta se calcula en forma autoconsistente resolviendo las ecuaciones estándar de evolución estelar junto con las condiciones de borde adecuadas. Es importante notar que la tasa de acreción de sólidos está acoplada al cálculo de la formación de la envoltura. De este modo, el régimen de acreción de los planetesimales y la interacción de éstos con la envoltura tendrán una incidencia importante en la forma en que ocurra la acreción de gas. Por este motivo, describiremos los dos regímenes involu- crados en la formación del núcleo de los planetas gigantes, el crecimiento de tipo runaway y, en más profundidad, el crecimiento oligárquico (post-runaway); siendo éste último el que adoptaremos para nuestro modelo. Por otra parte, dado que durante su formación el planeta está inmerso en un disco de gas y sólidos, debemos considerar un modelo que lo describa. Utilizaremos uno sencillo, el modelo de la nebulosa solar estándar (Hayashi 1981), 47 el cual determinará parte de las condiciones iniciales y de borde necesarias a la hora de resolver las ecuaciones diferenciales. 4.1. El disco protoplanetario El disco protoplanetario será caracterizado según el modelo propuesto por Hayashi (1981). Consideraremos que la nebulosa solar se encuentra en equilibrio, en lo cual está implícito que ha transcurrido el tiempo suficiente como para que los granos de polvo hayan sedimentado hacia el plano ecuatorial del disco, de modo que la estructura gaseosa estará compuesta fundamentalmente por moléculas de hidrógeno y átomos de helio. Se supone que cualquier efecto magnético es despreciable y, por lo tanto, no serán tenidos en cuenta en el modelo. La masa mínima de la nebulosa solar corresponde a la mínima cantidad de materia que debe contener el disco protoplanetario para dar cuenta de todos los cuerpos presentes actualmente en el Sistema Solar. Para estimar su valor se calcula la masa total de material sólido que se encuentra hoy en día presente en los planetas, asteroides, cometas, satélites, etc., y se calcula la cantidad de gas necesario para obtener la abundancia solar. El perfil de densidad surge de dividir la masa total de sólidos que contiene cada planeta por la región del disco que ocupa. La masa total se estima en 10−2 M⊙ , en un rango de distancias entre 0,35 y 36 UA. A este modelo del disco protoplanetario se lo conoce como Nebulosa Solar de Masa Mínima (NSMM). Para el desarrollo de esta Tesis supondremos que el disco protoplanetario está bien descripto por los siguientes tres parámetros: el perfil de temperatura, la densidad superficial de sólidos y la densidad volumétrica de gas. Aceptando la hipótesis de que los granos de polvo presentes en el disco se comportan como un cuerpo negro, con lo cual toda la energía que absorben es reemitida, y planteando la ley de Steffan-Boltzmann se obtiene para el perfil de temperatura una ley de potencias dependiente de la distancia: T (a) = αTeff (1 UA) T ⋆ eff a−q , (4.1) donde a es la distancia a la estrella central, T ⋆ eff es la temperatura efectiva relativa a la del Sol (es una cantidad adimensional y para los casos considerados en este traba jo eff = 1), Teff (1 UA) es la temperatura efectiva del disco a 1 UA, fijada en 280 K, que T ⋆ es la temperatura actual en el Sistema Solar en esa ubicación, y α es una constante que se introduce para tener en cuenta que en sus inicios el Sol era más luminoso que en la actualidad (α fue fijado arbitrariamente en 1,08; vale destacar que esta consideración no tiene ninguna incidencia en los resultados). La línea de hielo, asnow , es la ubicación en el disco donde la temperatura es tal que el agua se encuentra en estado sólido. Este cambio de fase en el estado del agua provoca un aumento en la densidad de sólidos lo cual, según 48 se cree, sería fundamental para la formación de planetas de gran masa. La línea de hielo se sitúa donde la temperatura es T = 170 K, que en este modelo corresponde a a = 3,16 UA. La definición de la NSMM lleva a que el a juste del perfil de densidad de sólidos siga una ley de potencias: Σ (a) = ηice σz (1 UA) a−p , (4.2) donde consideramos que σz (1 UA) = 10 g cm−2 , p = 3/2 y ηice es el factor de enriqueci- miento más allá de la línea de hielo dado por ηice = ( 1 4 si a < asnow si a > asnow . Si bien en esta ecuación está implícita la hipótesis de que los sólidos se encuentran distribuidos uniformemente, para el cálculo de la formación de planetas gigantes es también necesario adoptar un valor para la masa de los planetesimales del disco, como así también un valor para su radio. En la mayoría de las simulaciones, consideraremos que la densidad en masa de los planetesimales es para todos la misma e igual a 1,5 g cm−3 . En cuanto al tamaño, se considerarán poblaciones de planetesimales con radio único y poblaciones donde se adopta una función para representar la distribución de tamaños. Como veremos en los próximos capítulos, los resultados dependen fuertemente de estas hipótesis. En cuanto a la densidad volumétrica de gas, el perfil está dado por: a−b 2H donde σg (1 UA) es la densidad superficial de gas 1 UA, siendo ésta igual a 2 × 103 g cm−2 , b = 3/2 y H es la altura de escala del disco de gas, ρ (a) = σg (1 UA) , (4.3) H (a) = 0,05 a5/4 × (1,5 × 1013cm). Notemos que a está expresada en UA pero H debe estar en cm para que ρ (a) esté en g cm−3 , con lo cual aparece un factor de conversión 1,5 × 1013 cm. Los valores de las constantes σg (1 UA) y σz (1 UA) no son los propuestos en el artículo original de Hayashi (1981) sino que han sido actualizados y son un poco más elevados, correspondiendo en este caso al modelo propuesto por Kokubo (2001). Las características más importantes de la NSMM en la posición de Júpiter (a = 5,2 UA) se listan en la Tabla 4.1. (4.4) 4.2. Sobre el crecimiento de los planetesimales A la hora de hablar del crecimiento de los planetesimales es necesario mencionar bre- vemente cómo operan los efectos de enfocamiento gravitatorio y fricción dinámica. Sea 49 Tabla 4.1. Características de la NSMM en la ubicación actual de Júpiter (a = 5,2 UA). Temperatura Densidad superficial de sólidos Densidad volumétrica de gas 133 K 3,37 g cm−2 1, 5 × 10−11g cm−3 m la masa de un planetesimal, rm su radio, vrel la velocidad relativa promedio entre este planetesimal y los que lo rodean, y sea ρPl la densidad de planetesimales (masa por unidad de volumen del disco). Si el planetesimal acreta todos los planetesimales con los que tiene encuentros en un tiempo d t, su tasa de crecimiento será: dm d t = πr2 mρPlvrel . (4.5) Sin embargo, debido al efecto de atracción gravitatoria mutua entre los planetesimales, su trayectoria relativa no es lineal sino hiperbólica. Entonces, los planetesimales acreta- dos por m no serán solo aquellos cuyo parámetro de impacto sea menor o igual a rm (o sea, todos aquellos que directamente chocan con él), sino también aquellos que, por acción gravitatoria, sean deflectados de su trayectoria original y ésta termine en una colisión (ver figura 4.1). El valor máximo del parámetro de impacto para el cual se satisface esta con- dición puede obtenerse imponiendo que durante un encuentro debe conservarse la energía y el momento angular. De este modo, la acción de la gravedad produce un aumento en la sección eficaz de colisión que se denomina enfocamiento gravitatorio. El radio efectivo del planetesimal satisface la ecuación: donde θ se conoce como número de Safronov y está definido como: ef = r2 r2 m(1 + 2θ), (4.6) 1 2 v 2 esc v 2 rel (4.7) , θ ≡ donde vesc es la velocidad de escape de la superficie del planetesimal. Luego, cuanto más pequeña es la velocidad relativa o cuanto más masivo es el objeto, mayor es la sección eficaz de colisión. Entonces, los planetesimales más masivos o que se encuentren en ba jos regímenes de velocidades relativas tendrán mayores probabilidades de crecimiento. Otro efecto que favorece el crecimiento de los planetesimales de mayor masa es la fricción dinámica. Consideremos un cuerpo de masa M que se desplaza a través de un mar infinito y homogéneo de partículas de masa m. El efecto neto, luego de sucesivos encuentros entre M y los pequeños cuerpos que lo rodean, es similar al que experimenta un objeto moviéndose en un fluido, solo que en el primer caso la naturaleza de la “fricción” que sufre M es de carácter gravitatorio. El cuerpo M experimenta una desaceleración, 50 Figura 4.1. Representación esquemática del enfocamiento gravitatorio que produce un cuerpo masivo sobre los planetesimales más pequeños que se encuentran en su zona de influencia y presentan una velocidad relativa no nula respecto de éste. Los planetesi- males, que de seguir su trayectoria original no colisionarían con el embrión, sufren una deflección debido al efecto gravitatorio del mismo, resultando finalmente acretados por el embrión. cuya componente actúa en dirección paralela a su velocidad, que provoca una reducción de la misma. Se dice que esta continua desaceleración es debida a la fricción dinámica. La fricción dinámica, al oponerse a la velocidad orbital, provoca el decaimiento orbital del objeto, pero también afecta la excentricidad del mismo, tendiendo a circularizar las órbitas. El frenado es proporcional a la masa del objeto: cuanto mayor es su masa, mayor es la desaceleración. El efecto neto resulta ser una transferencia de energía cinética desde los objetos de mayor tamaño hacia los más pequeños. Ésto es de crucial importancia, pues gracias al enfocamiento gravitatorio los cuerpos más masivos conservan su status en la medida en que las velocidades evolucionan, mientras que los objetos más pequeños pierden ese privilegio debido a que la fricción dinámica no es operativa sobre ellos. En lo que respecta al crecimiento de los planetesimales para dar origen a los proto- planetas se puede decir que la formación de los planetas terrestres atraviesa tres etapas, mientras que la formación de los núcleos de los planetas gigantes se completa en tan so- lo las dos primeras (Thommes & Duncan 2006). En la primera etapa, los planetesimales crecen siguiendo un régimen runaway, lo cual significa que los más grandes pobladores del disco serán favorecidos en su crecimiento y rápidamente ganarán la masa suficiente para distinguirse del resto de los planetesimales. Cuando aparecen los primeros embriones pro- toplanetarios (cuerpos del orden de 1024 g, aunque este valor depende de la ubicación en el disco), el crecimiento se vuelve autorregulado puesto que, en estos casos, los objetos más grandes perturban gravitatoriamente a los planetesimales de su adyacencia lo cual restrin- ge la probabilidad de acreción dado que aumenta el régimen de velocidades relativas. Esta etapa se conoce como crecimiento oligárquico ya que solo los cuerpos de mayor masa alcan- zan su “masa de aislación”, lo cual significa que llegan a consumir todos los planetesimales que se encuentran ba jo su influencia gravitatoria. En la región de los planetas terrestres, 51 la masa de aislación típica resulta ser aproximadamente igual a la masa de Marte. La ter- cera etapa de crecimiento se produce por la colisión de estos embriones, en un crecimiento típicamente ordenado. Los modelos de formación de planetas terrestres estiman un perío- do de acreción del orden de 108 años cuando se consideran escenarios en ausencia de gas (Chambers 2001), o de solo algunas decenas de millones de años si se tienen en cuenta los efectos dinámicos de las resonancias con los planetas gigantes (Nagasawa, Lin & Thommes 2005). Este último hecho requiere de la pre-existencia de los planetas gigantes, con lo cual éstos deben formarse antes que los planetas terrestres. Para que ésto ocurra, y teniendo en cuenta el tiempo que involucra el crecimiento por colisión entre embriones, que corresponde a un régimen de crecimiento ordenado, la tercera etapa no debería ser significativa en la formación de los núcleos de los planetas gigantes, los cuales serían entonces el producto final del crecimiento oligárquico. O sea que, el crecimiento runaway y el oligárquico deben ser capaces de dar origen a embriones sólidos de 10 M⊕ en la región exterior del Siste- ma Solar para poder formar los núcleos de los planetas gigantes sin necesidad de invocar además colisiones entre embriones para llegar a masas de ese orden. A continuación describiremos en detalle los regímenes de crecimiento runaway y oligár- quico, y estableceremos las características del modelo utilizado en nuestras simulaciones. 4.2.1. Régimen de crecimiento runaway Aceptando que, por algún mecanismo, planetesimales de tamaño del orden del kilóme- tro se forman en el disco protoplanetario, existe un acuerdo general en que las primeras semillas de los embriones planetarios surgen debido a la acreción mutua entre los planete- simales. En el escenario clásico de la formación planetaria (Safronov 1969), el crecimiento de los planetesimales se considera ordenado. Sin embargo, ba jo este régimen, el tiempo de crecimiento de los núcleos de Júpiter y Saturno excedería el tiempo de vida estimado para la nebulosa protoplanetaria, con lo cual estos objetos no tendrían la posibilidad de capturar el gas necesario para formar la envoltura. Además, la escala de tiempo de formación de Urano y Neptuno superaría, en ese caso, la edad del Sistema Solar. A partir de la década del ’80, el crecimiento de los planetesimales comenzó a ser profundamente investigado ba jo la motivación de resolver este problema. Mediante simulaciones numéricas se encontró que el crecimiento de los planetesimales en los primeros estadíos de la acreción era el corres- pondiente al de un régimen de crecimiento en fuga (Greenberg et al. 1978, Kokubo & Ida 1996). De acuerdo con Ida & Makino (1993), este tipo de crecimiento se puede resumir de la siguiente manera: al principio, dado que las excentricidades e inclinaciones de los pla- netesimales se mantienen ba jas (las velocidades aleatorias de los planetesimales resultan proporcionales a su excentricidad e inclinación), las velocidades aleatorias de los planete- simales más chicos son pequeñas debido a que están reguladas por la dispersión mutua entre los planetesimales. Esto favorece al crecimiento del protoplaneta, el cual se acelera a medida que aumenta su masa debido al enfocamiento gravitatorio. Se desencadena enton- ces un proceso retroalimentado: los planetesimales más grandes crecen más rápido porque son gravitatoriamente dominantes entre sus vecinos, comenzando así el crecimiento en fuga 52 o crecimiento runaway. Los cuerpos en crecimiento runaway rápidamente se vuelven más masivos que el resto, siendo este el mecanismo que da origen a los primeros embriones planetarios en el disco. Durante la etapa de crecimiento runaway, tanto la tasa de creci- miento del protoplaneta como la proporción de masa entre el protoplaneta y el resto de los planetesimales se incrementa con el tiempo. 4.2.2. Régimen de crecimiento oligárquico Ida & Makino (1993) investigaron la evolución post-runaway de las excentricidades e inclinaciones de los planetesimales debido a la dispersión producida por la acción del cam- po gravitatorio del protoplaneta. Cuando los embriones en crecimiento runaway adquieren la masa suficiente como para afectar gravitatoriamente a los planetesimales que pueblan su zona de influencia, las velocidades aleatorias de los planetesimales aumentan y el creci- miento del embrión se vuelve más lento. Ida & Makino (1993) bautizaron a esta etapa como “etapa dominada por el protoplaneta”. Para estudiarla, los autores realizaron simulaciones tridimensionales de N -cuerpos, con las cuales analizaron la evolución de las excentricidades e inclinaciones de un sistema de planetesimales, todos de ellos de igual masa, frente a la pre- sencia de un protoplaneta. Tuvieron en cuenta las interacciones protoplaneta-planetesimal y planetesimal-planetesimal, pero no consideraron que hubiera acreción entre ellos. Ida & Makino (1993) encontraron que los planetesimales son fuertemente dispersados por el protoplaneta y que parte de la energía que adquieren en la dispersión se distribuye luego entre el resto de los planetesimales. Los planetesimales excitados de esta manera quedan confinados en una región alrededor del protoplaneta que es aproximadamente el doble del ancho de su zona de alimentación. En esta región, las excentricidades (em) y las inclina- ciones (im ) de los planetesimales son fuertemente perturbadas, y las velocidades aleatorias pueden aproximarse mediante la ecuación: vrel ≃ qhe2 m i + hi2 m i vk . Luego, es claro que un incremento en em e im se traduce en un aumento de las velocidades aleatorias. Aquí vk es la velocidad kepleriana correspondiente a la posición del protoplaneta (vk = a Ωk donde Ωk es la velocidad angular kepleriana de un objeto que gira en torno al m i1/2 (hi2 m i1/2 ) es la excentricidad (inclinación) Sol en una órbita circular de radio a), y he2 cuadrática media de los planetesimales en el disco. En base a los resultados de sus simulaciones, Ida & Makino (1993) proponen un creci- miento de tipo escalonado para los protoplanetas. En una primera etapa, los protoplanetas siguen un crecimiento runaway en el sentido usual: la perturbación que sufren los pla- netesimales está dominada por los propios planetesimales, siendo las velocidades relativas generalmente ba jas, lo cual favorece el rápido crecimiento de los embriones. Un crecimiento runaway no significa necesariamente que el tiempo de crecimiento del embrión disminuya a medida que aumenta su masa, sino que la razón de la masa entre dos cuerpos (embrión - planetesimal) aumenta con el tiempo. Además, durante el crecimiento runaway, la mayor (4.8) 53 parte de la masa del sistema se encuentra en los cuerpos pequeños, cuya distribución de masa puede a justarse con una ley de potencias, mientras que la masa del embrión se separa de esta distribución. Además, debido a la fricción dinámica, la excentricidad e inclinación del embrión se mantienen ba jas, lo cual facilita también el proceso de acreción. Cuando un protoplaneta es suficientemente masivo como para perturbar a los planetesimales de su vecindad, el sistema entra en una etapa en la cual su dinámica está dominada por el protoplaneta. Las simulaciones numéricas de N -cuerpos muestran que esta etapa ocurre en el régimen dominado por la dispersión, donde la relación entre los valores medios de las excentricidades y las inclinaciones es: m i1/2/hi2 m i1/2 ≃ 2. he2 (4.9) En esta segunda etapa de crecimiento post-runaway, la tasa de crecimiento del pro- toplaneta decrece como consecuencia del aumento en las velocidades relativas entre los planetesimales y el protoplaneta. Sin embargo, el cociente de masas entre el protoplaneta y los planetesimales sigue aumentando con el tiempo. Ida & Makino (1993) derivaron, en base a un modelo semi-anlítico la condición para el dominio de la dispersión protoplaneta-planetesimal sobre la dispersión planetesimal- planetesimal: 2 ΣM M > Σ m, (4.10) donde M es la masa del embrión sólido, m es el valor promedio de la masa de los planete- simales, Σ es la densidad superficial de masa debida a los planetesimales del disco y ΣM se define como: ΣM = M 2πa∆a , (4.11) donde 2πa∆a es el área de la región que ocupan los planetesimales excitados. Para una nebulosa solar estándar, esta condición puede ser trasladada a una relación entre M y m, la cual depende del perfil de densidad superficial del disco, del semieje mayor a y de la masa de los planetesimales. Para un modelo de nebulosa de ba ja masa, la transición entre los dos regímenes de crecimiento, ocurre cuando M/m ≃ 50 − 100. Esta condición, derivada en forma semi-analítica, fue corroborada por los autores con simulaciones de N -cuerpos. La segunda etapa en el régimen de crecimiento del protoplaneta comienza cuando el protoplaneta es todavía mucho menos masivo que su correspondiente masa de aislación. De este modo, cuando la masa del protoplaneta excede en, aproximadamente, 100 veces la masa típica de un planetesimal, el protoplaneta excita las velocidades de los planetesimales que lo rodean y las velocidades aleatorias comienzan a depender de su masa. Ida & Makino (1993) estimaron el tiempo característico de crecimiento del protoplaneta 54 Tgrow (“the mass–doubling time”), ba jo la suposición de una densidad superficial de sólidos constante, y obtuvieron que: Tgrow ≡ 1 M ∝ M −1/3 he2 m i. dt !−1 dM (4.12) Para estimar he2 m i, consideraron un estado de equilibrio donde la dispersión producida por el embrión se encuentra compensada por el amortiguamiento debido a la viscosidad con el gas nebular. Como resultado, en la primera etapa (correspondiente al crecimiento runaway, donde las perturbaciones están dominadas por los planetesimales), he2 m i ∝ m8/15 (notar que no depende de la masa del protoplaneta), mientras que en la etapa post-runaway, m i ∝ m1/9M 2/3 . Así, en la primera etapa Tgrow ∝ M −1/3 , de modo que los protoplanetas he2 crecen lo suficientemente rápido como para que la relación M/m ≃ 50 − 100 se alcance en una escala de tiempo despreciable. En la segunda etapa, es el protoplaneta el perturbador dominante y Tgrow ∝ M 1/3 , con lo cual el crecimiento del protoplaneta se hace más lento a medida que su masa aumenta. Este tipo de dependencia temporal es característica del crecimiento de tipo ordenado (que es el régimen correspondiente a la última etapa de formación de los planetas terrestres). Sin embargo, la etapa dominada por el protoplaneta es mucho más corta que la correspondiente al crecimiento ordenado porque los protoplanetas crecen por la acreción de planetesimales y no por la colisión entre cuerpos de tamaños similares, donde la fricción dinámica está ausente. Continuando la investigación de Ida & Makino (1993), Kokubo & Ida (1998) hicieron simulaciones tridimensionales de N -cuerpos para estudiar el escenario de acreción post- runaway. Como condición inicial consideraron la presencia de dos protoplanetas, ahora sí, creciendo por la acreción de planetesimales. En la primera etapa de acreción, confirmaron que los protoplanetas crecen según el régimen runaway. El crecimiento runaway se inte- rrumpe cuando M ∼ 50 m, tal como lo predecía el modelo semi-analítico de Ida & Makino (1993). Luego, los protoplanetas continúan su crecimiento manteniendo una separación orbital típica de 10 RH , donde RH es el Radio de Hill del protoplaneta, . 1/3 RH = a (cid:18) M 3M⋆ (cid:19) Durante su crecimiento se manifiesta una repulsión orbital entre los protoplanetas, la cual es consecuencia del acoplamiento entre la dispersión provocada por los cuerpos de mayor masa y la fricción dinámica en la población de planetesimales. Los protoplanetas crecen manteniendo el cociente entre sus masas cercano a uno, pero la relación de las masas entre los protoplanetas y los planetesimales se incrementa en el tiempo. De este modo, se dice que los protoplanetas siguen un crecimiento de tipo “oligárquico” ya que no se produce acreción entre los planetesimales y solo crecen los cuerpos de mayor masa. La distribución de masa en el disco se vuelve bimodal: hay un pequeño grupo de embriones, embebidos en un disco poblado por un gran número de planetesimales. Kokubo & Ida (1998) definieron (4.13) 55 formalmente a este tipo de crecimiento como crecimiento oligárquico en el sentido de que no solo uno sino varios son los protoplanetas que dominan al sistema de planetesimales. Kokubo & Ida (2000) investigaron, también con simulaciones tridimensionales de N - cuerpos la evolución de un enjambre de planetesimales hasta que algunos alcanzan la masa necesaria para ser considerados embriones, pero ahora incluyendo el efecto de la fricción nebular. Confirmaron la existencia de un crecimiento runaway en la primera etapa de la for- mación protoplanetaria y una segunda etapa de crecimiento oligárquico de los embriones, en el mismo sentido que lo encontrado por Kokubo & Ida (1998). Así mismo, al analizar la evolución de la excentricidad y la inclinación cuadrática media de los planetesimales duran- te el crecimiento oligárquico de los embriones, encontraron que sus resultados concordaban con aquellos predichos por la teoría semi-analítica de Ida & Makino (1993). En función de estos resultados, tanto numéricos como analíticos, es ampliamente acep- tado que el crecimiento de los embriones sólidos atraviesa dos etapas: la primera de tipo runaway, cuya duración es corta comparada con las escalas de tiempo involucradas en la formación de los planetas gigantes; y la segunda, de tipo oligárquico, que es la que domina durante el proceso de formación. 4.3. La aparición de los embriones planetarios Cuando la dinámica de la población de planetesimales está dominada por la dispersión, la tasa de crecimiento de un embrión sólido está bien descripta por la aproximación de partículas en una ca ja (Safronov 1969), Σ 2h (4.14) π R2 eff vrel , dMc dt ≃ F donde Mc es la masa de la componente sólida del protoplaneta (que en nuestro caso corres- ponde al núcleo de un planeta gigante), h es la altura de escala del disco de planetesimales, Reff es el radio efectivo de captura del embrión y F es un factor que se introduce para compensar la tasa de crecimiento producto de utilizar la aproximación de dos cuerpos en el cálculo de la velocidad de dispersión de los planetesimales, la cual está modelada por un único valor para la excentricidad y la inclinación, correspondiente al valor cuadrático medio. F está estimado en ∼ 3 (Greenzweig & Lissauer 1992). Debido al enfocamiento gravitatorio, el radio efectivo de captura del protoplaneta, Reff , es mayor que el radio geométrico del embrión sólido, Rc , c 1 + (cid:18) vesc vrel (cid:19)2! , eff = R2 R2 con vesc la velocidad de escape de su superficie y vrel la velocidad relativa entre el proto- planeta y los planetesimales, (4.15) 56 √e2 + i2 a Ωk , (4.16) vrel ≃ m i1/2 (i ≡ hi2 mi1/2 ); e (i) es el valor cuadrático donde, a partir de ahora, definimos e ≡ he2 medio de la excentricidad (inclinación) con respecto al plano del disco protoplanetario (e, i << 1) y Ωk es la velocidad angular kepleriana. En lo que sigue utilizaremos la aproximación i ≃ e/2 y h ≃ ai (h es la altura de escala de los planetesimales). Siguiendo a Thommes, Duncan & Levison (2003), adoptamos para e la expresión de equilibrio que se deduce para el caso donde las perturbaciones gravitatorias debidas al protoplaneta están balanceadas por el arrastre gaseoso. Sea TVS la escala de tiempo en la cual un cuerpo de masa Mp (en nuestro caso, Mp es la masa total del protoplaneta, ésto es, la masa del núcleo más la masa de la envoltura) excita a los planetesimales que lo rodean (Ida 1990), GM !2 1 40 Ω2 k a3 y sea la escala de tiempo asociada al amortiguamiento del gas (Adachi, Hayashi & Naka- zawa 1976): e4Mp ΣM a2Ωk TVS ≃ , (4.17) Tgas ≃ 1 e m (CD/2)πr2 mρaΩk , (4.18) donde CD es el coeficiente de amortiguamiento de la componente gaseosa del disco (adi- mensional y de orden uno para estos casos), y rm el radio de un planetesimal de masa m. Igualando las ecuaciones (4.17) y (4.18) y despejando e, se obtiene para la excentricidad de equilibrio de los planetesimales: e = 1,7m1/15M 1/3 p ρ2/15 m β 1/5C 1/5 D ρ1/5M 1/3 ⋆ a1/5 , (4.19) donde ρm es la densidad en masa de un planetesimal, ρ es la densidad volumétrica de gas en el disco protoplanetario, y β es al ancho, en unidades del radio de Hill, de la región que ocupan los planetesimales excitados por el protoplaneta (considerando que, potencialmente, hay otros embriones creciendo en la vecindad, β es del orden de 10). Según Chambers (2006), utilizar los valores de equilibrio para e e i es una aproximación aceptable cuando se consideran embriones m > 10−2 M⊕ , consistente con las masas de los núcleos iniciales de todas nuestras simulaciones (ver capítulos 6 y 7). Sin embargo, cuando un embrión sólido alcanza la masa necesaria como para ligar gravitatoriamente gas de la nebulosa en la cual está inmerso, la presencia de esta envoltura debe ser considerada en el cálculo del radio efectivo de captura, Reff . La envoltura gaseosa modifica la trayectoria de los planetesimales ingresantes, ya que estos se ven afectados por la fricción del gas, lo cual aumenta el radio de captura del protoplaneta. De este modo, 57 b’ b Figura 4.2. Esquema del enfocamiento gravitatorio y viscoso de un protoplaneta. De- bido a la gravedad del planeta y a la viscosidad del gas la sección eficaz de captura de un protoplaneta es mayor que su sección eficaz geométrica. La gravedad y el frenado viscoso deflectan la trayectoria de un planetesimal. Cuando el parámetro de impacto se va reduciendo, de b’ a b, llegará un momento en que el planetesimal seguirá una trayectoria que lo hace colisionar con el núcleo. El valor de b para el cual esto ocurre se define como el radio efectivo del planeta. 58 además del enfocamiento gravitatorio, tendremos un “enfocamiento viscoso” debido a la fricción producida por la atmósfera. El radio efectivo Reff , en la forma dada en la Ec. (4.15) domina cuando la masa de gas ligado es despreciable. Pero cuando un embrión adquiere la cantidad de gas suficiente para formar apenas una delgada atmósfera, su radio efectivo aumenta, separándose consi- derablemente del calculado anteriormente (ver figura 4.2). Para calcular el radio efectivo del protoplaneta en presencia de una atmósfera hay que considerar que sobre los planete- simales ingresantes actúan dos fuerzas: la gravedad y el arrastre viscoso. Consideremos un planetesimal que ingresa a la atmósfera del protoplaneta con una velocidad vrel (Ec. (4.16)). Su ecuación de movimiento se obtiene de considerar, entonces, la acción de la gravedad, ~fG = − junto con la acción de la fricción gaseosa. En la Ec. (4.20), r es la coordenada radial con origen en el centro del protoplaneta y Mr es la masa contenida hasta un radio r . La fuerza de arrastre viscoso actuando sobre un cuerpo esférico de radio rm que via ja a una velocidad v es: (4.20) r , GMrm r2 ~fD = − donde aquí ρg es la densidad de la envoltura y CD = 1 (Adachi, Hayashi & Nakazawa 1976). Luego, la ecuación de movimiento surge de combinar ambas ecuaciones, CD π r2 m ρg v 2 v , (4.21) 1 2 GMrm r2 1 2 d~v dt m (4.22) CD π r2 m ρg v 2 v . = ~fG + ~fD = − r − Para comprender la importancia de la viscosidad en el cálculo del radio efectivo, la figura 4.3 muestra, para una de nuestras simulaciones, los valores que toman, a lo largo de todo el proceso de formación, el radio efectivo considerando solo el enfocamiento gravitatorio (R∗ eff ) y el que surge de considerar también el efecto viscoso de la envoltura (Reff ). Los detalles de la simulación que caracterizan este ejemplo se discutirán en el capítulo 6. Como se puede apreciar, aún cuando el gas ligado al embrión es despreciable (corresponde a cuando el radio del planeta es inferior al radio efectivo), la viscosidad del gas igualmente contribuye a aumentar el radio efectivo (en este caso, el gas no pertenece a la envoltura del planeta pero sí pertenece a su esfera de Hill). El apartamiento entre ambos radios se acentúa cuando la masa de la envoltura domina la masa total del planeta. De hecho, la eff es de más de un orden de magnitud. Cuando el planeta es muy diferencia entre Reff y R∗ masivo, notamos que R∗ eff decrece. Ésto se debe a que la dispersión de los planetesimales es muy grande (la velocidad relativa, vrel , es alta), lo cual atenúa el enfocamiento gravitatorio (Ec. (4.15)). Sin embargo, como la viscosidad domina el cálculo de Reff , este hecho no tiene un efecto apreciable. 59 4 3.5 3 2.5 2 1.5 1 0.5 0 ] m c 9 0 1 [ ) R ( g o l Rp Reff R * eff −0.5 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 log(Rc) [10 9 cm] Figura 4.3. En esta figura se muestra la evolución del radio efectivo del planeta, Reff , y del radio efectivo por enfocamiento gravitatorio sin considerar la viscosidad del gas, R∗ eff . La abscisa corresponde al logaritmo del radio del núcleo. Se incluyó también la evolución del radio del planeta a modo de referencia (el radio del planeta se calcula teniendo en cuenta la masa del núcleo y la masa de gas). Debido, justamente, a la naturaleza de la escala logarítmica, los estadíos iniciales de la formación del planeta, que corresponden a un embrión sólido rodeado de una pequeña atmósfera de gas, se pueden observar en detalle (log Rc ∼< 0, 2). Notemos que, muy al principio de la formación, el radio efectivo es mayor que el radio del planeta (log Rc ∼< −0, 3). Esto se debe a que el valor del radio efectivo está dominado por el enfocamiento gravitatorio. Cuando la masa de la envoltura crece, la viscosidad se vuelve cada vez más importante y el Reff se aparta de R∗ eff . El radio del planeta es generalmente mayor que el radio efectivo debido a que las capas de gas que conforman la parte exterior del planeta tienen una densidad muy baja y forman una envoltura muy extendida. 60 4.4. La envoltura gaseosa 4.4.1. Ecuaciones constitutivas Las ecuaciones que gobiernan la estructura del planeta son las ecuaciones diferenciales estándar de la evolución estelar (ver, por ejemplo, Kippenhahn & Weigert 1990). Si des- preciamos la rotación del planeta y la presencia de campos magnéticos, y adoptamos la formulación lagrangiana considerando simetría esférica, donde las variables independientes son la masa Mr y el tiempo t, tenemos que la estructura gaseosa del planeta está bien descripta por cuatro ecuaciones diferenciales acopladas: ∂ r ∂Mr ∂P ∂Mr ∂L ∂Mr ∂T ∂Mr = 1 4πr2ρg GMr = − 4πr4 ∂S = ǫac − T ∂ t GMr T = − 4πr4P ∇ Ecuación de conservación de la masa (4.23) Ecuación de equilibrio hidrostático (4.24) Ecuación de balance energético Ecuación de transporte (4.25) (4.26) donde r es la coordenada radial, ρg es la densidad, P es la presión, G es la constante de gravitación universal, L es la luminosidad, ǫac es luminosidad debido a la acreción de planetesimales, T es la temperatura, S es la entropía por unidad de masa y ∇ es el gradiente de temperatura adimensional, d ln T ∇ ≡ d ln P que depende del tipo de transporte de energía. Si el transporte es radiativo, ∇ está dado por: (4.27) , (4.28) 3 κLP ∇rad = 16πar cG Mr T 4 con ar la constante de densidad de radiación, c la velocidad de la luz en el vacío y κ la opacidad media de Rosseland. Si el transporte es convectivo se considerará que el gradiente es el correspondiente al caso adiabático, ∇ = ∇ad . El tipo de transporte (radiativo o convectivo) estará determinado por el Criterio de Estabilidad de Schwarzschild, si ∇rad < ∇ad ∇ =  ∇rad si ∇rad ≥ ∇ad . ∇ad  En lo que respecta a la energía, notemos que la ecuación (4.25) está caracterizada por dos términos: 61 ǫac representa el calor, por unidad de tiempo y unidad de masa de la envoltura, que es suministrado por los planetesimales cuando estos son capturados por el planeta. Supondremos que los planetesimales son suficientemente grandes como para no de- sintegrarse antes de llegar al núcleo. De hecho, supondremos que en su trayectoria no sufren ningún proceso de ablación (lo cual es una hipótesis simplificadora impor- tante), con lo cual llegan al núcleo con la misma masa con la que ingresaron. El intercambio de energía que tienen con la envoltura será solo como consecuencia de la fricción con el gas. El procedimiento para el cálculo de ǫac será detallado en § 4.4.5. ∂S ∂ t T = que, según la primera ley de la termo- el segundo término de la ecuación es −T dinámica, resulta ∂S ∂V ∂U dQ ∂ t ∂ t dt ∂ t donde Q es el calor, U es la energía interna y V es el volumen, todos por unidad de masa. De esta expresión se puede ver que hay otros dos mecanismos que controlan ∂V ) y la variación la energía: la contracción gravitatoria (reflejada en el término P ∂ t ∂U de la energía interna, . En la práctica, el término T d S será evaluado según la ∂ t relación termodinámica: (4.29) + P = T d S = CP d T − δ ρg d P , (4.30) donde CP es el calor específico a presión constante (por unidad de masa) y δ se define como: . δ ≡ (4.31) ∂ ln ρg ∂ ln T (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)P Dependiendo de la fase de formación en la que se encuentre el planeta, dominará el primero o el segundo término de la ecuación (4.25). Cuando el proceso está controlado fundamentalmente por la acreción de sólidos (la primera etapa), la fuente primordial de luminosidad son los planetesimales ingresantes. Como veremos más adelante, la energía que liberan los planetesimales es la que “sostiene” a las capas de gas, impidiendo que se desplomen contra el núcleo, regulando así la escala de tiempo de la formación. Cuando el protoplaneta incorporó a la mayoría de los sólidos de su vecindad, y la masa de gas ligada alcanza valores aproximadamente iguales a los de la masa del núcleo, la energía que aportan los planetesimales a la presión del gas ya no es suficiente para contrarrestar la fuerza de la gravedad y las capas de gas comienzan a contraerse, entrando el planeta en la segunda etapa. En este caso, la ecuación (4.25) está gobernada por el segundo término. Esta etapa es mucho más rápida que la primera y se la conoce como etapa del crecimiento runaway de gas o simplemente etapa runaway. 62 Por otra parte, la ecuación de equilibrio hidrostático en simetría esférica (y tomando como variables a r y a t) tiene la siguiente forma: ∂φ ∂P = −ρ ∂ r ∂ r donde, para el problema de dos cuerpos, el potencial gravitatorio φ es: φ = − GMr r , (4.32) (4.33) que es el caso considerado en la ecuación (4.24) (los “dos cuerpos” son el protoplaneta por un lado y un elemento de gas por el otro). Sin embargo, si analizamos el problema a resolver, nos damos cuenta que no tenemos exactamente un problema de dos cuerpos ya que hay una influencia gravitatoria importante: la estrella central. Estrictamente hablando, el tratamiento debería involucrar el potencial gravitatorio para el problema restringido de los tres cuerpos. Por simplicidad para el tratamiento numérico, en cambio, conviene utilizar un potencial aproximado (Benvenuto & Brunini 2005). La idea es considerar una corrección al potencial de los dos cuerpos que establezca que el límite para la atracción gravitatoria del planeta termina en el radio de Hill1 (RH ), con lo cual ∂φ ∂ r deberá tender a cero cuando r tienda a RH . Para ello proponemos: , (4.34) = con x = ∂φ GMr r r2 f (x) ∂ r RH donde f (x) = 1 − x3 . Que f (x) tenga esta forma no es arbitrario sino que surge como la primera corrección al potencial de los dos cuerpos cuando se hace un promedio sobre el po- tencial correspondiente al problema restringido de los tres cuerpos (o sea, cuando se calcula 4π R4π φR dΩ con φR el potencial del problema restringido). Si bien considerar esta correc- 1 ción es analíticamente correcto, esta expresión no resulta conveniente en el tratamiento numérico cuando x → 1. Por las características del problema que se quiere abordar, cuan- do las ecuaciones diferenciales se llevan a su forma en diferencias finitas (ver § 5.1) resulta conveniente extender el grillado más allá del borde del planeta (Rp ). Ahora bien, como hemos dicho, el potencial corregido tiende a cero cuando r tiende a RH y, de este modo, ∂φ cuando r >RH tendremos que < 0. Pero, de la ecuación (4.24), vemos que éste hecho ∂ r implicaría un gradiente de presión positivo en la nebulosa. Dado que el gradiente de presión está conectado con el gradiente de temperatura, ésto entraría en contradicción con nuestra hipótesis de temperatura constante para la nebulosa. Para que esto no ocurra, y para que las funciones sean bien comportadas, será conveniente que f (x) tienda suavemente a cero cuando x → ∞. Resulta entonces adecuado empalmar 1 − x3 con la función de Fermi A , a justando los parámetros A y B para que f (x) resulte continua en x0 (que eB (x−x0 ) + 1 1El radio de Hill corresponde al radio equivalente de una esfera de volumen igual al lóbulo de Roche del planeta. 63 es un número cercano a 1). De este modo, obtenemos: f (x) =   6x2 0 con A = 2(1 − x3 0 ) y B = . En nuestras simulaciones hemos considerado general- 1 − x3 0 mente x0 = 0,9, aunque es importante destacar que el valor que tome x0 no repercute significativamente en los resultados, siempre y cuando, x0 sea cercano a 1. 1 − x3 A eB (x−x0 ) + 1 si x ≤ x0 si x > x0 4.4.2. La ecuación de estado del gas Para resolver las ecuaciones (4.23)-(4.26) es necesario establecer relaciones constituti- vas, esto es, dar los valores de ρg , ∇ad , S y κ en función de la temperatura y la presión. Las propiedades termodinámicas de un fluido están caracterizadas por la ecuación de estado (EOS), la cual determina unívocamente la densidad y el resto de las cantidades termodi- námicas en función de la presión y la temperatura. En esta sección discutiremos la EOS considerada para el gas y en la próxima sección discutiremos sobre las tablas utilizadas para la opacidad. La EOS más ampliamente utilizada en los cálculos de estructura de la componente gaseosa de los planetas gigantes es la de Saumon, Chabrier & Van Horn (1995, de ahora en más SCVH) y es la que adoptaremos para nuestro modelo. SCVH calcularon numéricamente la EOS para gases de hidrógeno y helio puros, considerando detalladamente los efectos no ideales, ba jo las condiciones usuales de presión y temperatura presentes en los objetos subestelares (estrellas de masa inferior a 1 M⊙ , enanas marrones y planetas gigantes). Los resultados se agrupan en tablas que cubren el rango de temperatura de 2, 10 ≤ log T ≤ 7, 06 [K] y de presión 4 ≤ log P ≤ 19 [dinas cm−2 ]. Un esquema del diagrama de fases del hidrógeno de esta EOS se esquematiza en la figura 4.4. En la región de ba ja densidad y ba ja temperatura, el hidrógeno se encuentra en estado neutro en forma de átomos y moléculas. Las moléculas dominan a ba jas tem- peraturas (log T ≤ 3, 5), las cuales comienzan a disociarse a medida que la temperatura aumenta. A temperaturas más altas, los átomos se ionizan formando un plasma de proto- nes y electrones. En la figura 4.4, la curva de rayas cortas delimita las tres regiones. Para densidades superiores a log ρ ∼ −2, los átomos y las moléculas interactúan fuertemente y forman un fluido no-ideal. A densidades cercanas a log ρ = 0, la distancia entre los átomos de hidrógeno es comparable al radio de Bohr, con lo cual las funciones de onda de dos electrones cercanos se superponen. Los electrones son entonces forzados a pasar a estados no-ligados y el fluido se ioniza por presión. Según los cálculos de SCVH, esta ionización no es gradual, sino que ocurriría en forma discontinua a través de una transición de fase 64 de primer orden, conocida como “Plasma Phase-Transition” (PPT). En una transición de fase de primer orden todas las cantidades termodinámicas son discontinuas, exceptuando la presión, la temperatura y los potenciales químicos. El valor de la temperatura a par- tir del cual la ionización comienza a producirse en forma continua corresponde al punto crítico, log Tc = 4,185, cuyos valores de la presión y la densidad son: log Pc = 11, 788 y log ρc = −0,456. La PPT separa la fase no ionizada (constituida por moléculas de H2) de la fase ionizada a densidades más altas formada por hidrógeno metálico (H+ ). Respecto de la PPT es importante aclarar que esta transición de fase no está muy comprendida y que no existen evidencias en el laboratorio que sostengan su existencia. Este hecho es extremadamente relevante puesto que, como vemos en la figura 4.4, el interior de Júpiter estaría afectado por la PPT. De hecho, una de las principales fuentes de incerteza a la hora de estimar el núcleo sólido de los planetas gigantes son los problemas que existen en torno al cálculo de la EOS. Afortunadamente, en los cálculos de formación de un planeta de la masa de Júpiter, la estructura del protoplaneta no estaría afectada por esta discontinuidad (si es que, de existir la PPT, ésta se ubica donde actualmente se cree que está), ya que durante la formación la densidad de las capas de la envoltura es mucho menor que en el caso de un planeta totalmente formado, que tuvo que sufrir, para llegar a ese estado, la contracción abrupta de sus capas de gas (Benvenuto & Brunini 2005). En cuanto a la ecuación de estado para el helio, se identifican regímenes dominados por el He atómico y por sus iones, He+ y H++ (ver figura 4.5). En el caso del helio, los autores no resuelven el problema de la ionización por presión. Téngase en cuenta, sin embargo que en un planeta joviano, la envoltura está compuesta mayormente por hidrógeno. Tanto en las envolturas de los planetas gigantes como en la mayoría de los problemas astrofísicos, el hidrógeno y el helio no se encuentran como sustancias puras sino mezclados entre sí. De este modo, resulta necesario tener una EOS para este tipo de mezclas. Lo ideal sería poder calcular esta EOS para una composición cualquiera en forma autoconsisten- te. Sin embargo, la complejidad del problema hace que, en la actualidad, deban utilizarse ciertas simplificaciones. SCVH sugieren que la EOS de la mezcla se puede obtener inter- polando adecuadamente las tablas de las EOS para el hidrógeno y el helio. De entre los modelos de interpolación posibles, SCVH proponen que la interpolación se haga haciendo uso de la regla de adición volumétrica (“additive-volume rule”). Esta regla está basada en las propiedades termodinámicas de las variables. Las variables como la temperatura y la presión son variables intensivas, lo cual significa que son uniformes en un sistema en equilibrio. Las variables extensivas, como el volumen, la entropía y la energía interna, son aditivas en sistemas idénticos, en el límite del régimen ideal. En esencia, la regla de adición lo que considera es que dado un sistema formado por dos o más subsistemas, tendrá como ecuación de estado a aquella que surja como combinación de las EOS de cada subsistema. Claramente, esta regla no es exacta puesto que ignora las interacciones entre las diferentes especies de hidrógeno y helio. Además, el estado de equilibrio de ionización de una mez- cla de hidrógeno y helio estará acoplado a la densidad electrónica. Esto no está tenido en cuenta en la regla de adición volumétrica. Este punto es especialmente importante en las regiones de ionización por presión, donde las interacciones mutuas son fuertes. De este 65 Figura 4.4. Diagrama de fases para el hidrógeno. La región limitada por la línea a rayas es la que cubre la EOS de SCVH. Las curvas etiquetadas con a, b, c, d, y e co- rresponden a los interiores de diversos ob jetos ricos en hidrógeno. La curva a representa el interior de Júpiter. La envoltura está prácticamente dominada en su totalidad por el hidrógeno molecular. Como se ve, su interior más profundo atraviesa la PPT. Justo debajo de la PPT, el fluido se aparta fuertemente del régimen ideal debido a las fuer- zas intermoleculares repulsivas. Las curvas b, c, y e representan el interior de estrellas de secuencia principal, cuyas masas son 0,3, 1, y 15 M⊙ respectivamente. La curva d corresponde a una enana blanca DA estratificada en una capa rica en hidrógeno y otra rica en helio, alrededor de un núcleo de carbono. (Figura perteneciente al artículo de Saumon, Chabrier & Van Horn (1995)) 66 Figura 4.5. Diagrama de fases para el helio correspondiente a la EOS de SCVH. Aquí la línea de puntos, etiquetada por la letra a, representa el estado termodinámico de la envoltura de una enana blanca DB compuesta solamente por helio. (Figura perteneciente al artículo de Saumon, Chabrier & Van Horn (1995)) 67 modo, la naturaleza, la localización y la mera existencia de la PPT estará afectada por la presencia del helio. En este régimen, la regla de adición volumétrica preserva a la PPT y a su curva de coexistencia. En su artículo, SCVH derivan las fórmulas para la obtención de los valores interpo- lados de las cantidades termodinámicas. Desafortunadamente, en el artículo original de SCVH existen errores de tipeo en las fórmulas finales de interpolación, hecho que nos llevó a calcular erróneamente la EOS de la mezcla en nuestros primeros traba jo, y que tuvo importantes consecuencias en nuestros resultados (volveremos sobre este tema en el capí- tulo 6). Por este motivo, derivaremos a continuación las ecuaciones para la interpolación, escribiendo en negrita los términos con errores de tipeo del manuscrito original. Dada una variable extensiva general, W (P , T ), la regla de volúmenes aditivos establece que, si Wi (P , T ) caracteriza a esta cantidad en el sistema i, para un sistema constituido por todos los subsistemas i tendremos que: W (P , T ) = Xi donde Xi es la fracción del sistema total ocupado por el subsistema i. En el caso de una mezcla de hidrógeno y helio, esto estará representado por la fracción en masa, X , para el hidrógeno, e Y , para el helio, tal que se satisface que: Xi W i(P , T ), (4.35) X = 1 − Y . Para el volumen específico (1/ρg ), y la energía interna tendremos, respectivamente: (4.36) 1 ρg (P , T ) = 1 − Y g (P , T ) ρH + Y g (P , T ) ρHe (4.37) U (P , T ) = (1 − Y ) U H (P , T ) + Y U He (P , T ). En el caso de la entropía, la interpolación se hace de la misma manera pero se agrega un término que tiene en cuenta la corrección para la entropía de mezcla del gas ideal. Con este término se recupera el límite en el caso del gas ideal. Escribimos entonces para la entropía por unidad de masa: (4.38) S (P , T ) = (1 − Y ) S H(P , T ) + Y S He(P , T ) + Smix (P , T ). Para calcular las derivadas de ρg (P , T ) y de S , diferenciamos las ecuaciones (4.37) y (4.39). Para ello resulta conveniente definir las derivadas logarítmicas: ∂ log ρg ∂ log ρg ∂ log P (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T ∂ log T (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)P , ρP = (4.39) 68 ρT = . (4.40) Similarmente, con lo cual se obtiene , SP = ST = ∂ log S ∂ log T (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)P ρg ρT = (1 − Y ) ρH g ∂ log S ∂ log P (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T ρg ρH ρHe T + Y T , ρHe g ρP = (1 − Y ) ρg ρH g ρH P + Y ρg ρHe g ρHe P , En el caso de la entropía2 , , (4.41) (4.42) (4.43) , . (4.44) (4.45) SH S SH S S He T + S He P + SHe S SHe S S H T + Y S H P + Y Smix S Smix S ST = (1 − Y ) ∂ log Smix ∂ log T (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)P ∂ log Smix ∂ log P (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T SP = (1 − Y ) Las derivadas de la entropía son importantes puesto que definen el gradiente adiabático, ∂ log T ∂ log P (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)S = − ∇ad = Para evaluar las derivadas hay que calcular la entropía de mezcla, la cual satisface 1 − Y 2 Smix = kB 1 + XH + 3XH2 × { ln(1 + β γ ) + mH − X H e ln(1 + δ) + β γ [ln(1 + 1/β γ ) + − X He ln(1 + 1/δ)]} e donde mH es la masa del hidrógeno y XA es la concentración de partículas A. Finalmente, las constantes β , γ y δ 3 se definen como: (4.47) SP ST . (4.46) , β = mH mHe Y 1 − Y (1 + XH + 3XH2 ) 3 (1 + 2XHe + XHe+ ) 2 2 (2 − 2XHe − XHe+ ) 3 (1 − XH2 − XH ) Dadas estas fórmulas, se pueden interpolar las tablas de las ecuaciones de estado del hi- drógeno y del helio para obtener la EOS de la mezcla para una dada composición. En los cálculos de esta Tesis se utilizó que la abundancia de helio era Y = 0,3. (4.50) (4.49) (4.48) γ = δ = β γ . , 2Los cocientes en negrita están invertidos en el artículo de SCVH. 3En lugar del factor 2/3, la ecuación para δ tenía en factor 3/2. 69 4.4.3. Opacidades La opacidad, al igual que la EOS, juega un papel determinante en el cálculo de la for- mación de planetas gaseosos. Fundamentalmente, es la opacidad media de Rosseland de los granos un factor importantísimo a la hora de estimar las escalas de tiempo y la masa del núcleo para la cual comienza la etapa del runaway gaseoso. En nuestras simulaciones he- mos considerado las tablas de opacidades de los granos de Pollack, McKay & Christofferson (1985), las cuales dominan para ba jas densidades y temperaturas, mientras que para tem- peraturas superiores a los 1000 K utilizamos las tablas de Alexander & Ferguson (1994), las cuales están calculadas hasta temperaturas del orden de los 104 K. Eventualmente, si algún cálculo lo requiriera, para temperaturas superiores se consideran las opacidades de Rogers & Iglesias (1992). Respecto de la opacidad de los granos, es relevante hacer algunos comentarios sobre su influencia en los cálculos de formación de planetas gigantes, debido a que la tasa de contracción de las capas de gas de la envoltura es fuertemente dependiente de la opacidad. Como mencionamos en el párrafo anterior, la opacidad media de Rosseland utilizada es la publicada por Pollack, McKay & Christofferson (1985), la cual fue calculada para un conjunto de diferentes especies de granos de polvo que estarían presentes en la nebulosa solar primordial. Los resultados abarcan un rango de temperatura de entre 10 y 2.500 K (la opacidad debida a los granos de polvo domina hasta, aproximadamente, 1.500 K), con la densidad del gas nebular variando entre 10−14 y 1 g cm−3 . Las integraciones se hicieron considerando numerosas distribuciones de tamaños, incluyendo la correspondiente al medio interestelar la cual se consideró como caso nominal. Los autores encontraron que la opacidad no es demasiado sensible a la distribución considerada, siempre y cuando haya partículas cuyo tamaño supere las decenas de micrones. Los sencillos modelos analíticos de Stevenson (1982) para el cálculo de la formación de los planetas gigantes mostraban ya la dependencia entre el comienzo de la etapa runaway de gas y la opacidad. De todas formas, las simplificaciones utilizadas para elaborar este modelo requieren que sus conclusiones sean tomadas con mucho cuidado. Sin embargo, las simulaciones numéricas de Pollack et al. (1996), donde se considera un modelo mucho más realista y detallado, mostraron que la dependencia con la opacidad era, efectivamente, muy importante. Pollack et al. encontraron que si la opacidad se reduce en un factor 50 (respecto de la correspondiente a la del gas interestelar), el tiempo de formación decrece en forma considerable. Si bien la composición de los granos en la atmósfera protoplanetaria debería ser similar a la presente en la nebulosa solar, la distribución del tamaño de los granos está influenciada por la microfísica asociada a los procesos atmosféricos, con lo cual no necesariamente sería igual a la del gas interestelar. Para evaluar la plausibilidad de una reducción en la opacidad en las atmósferas de los protoplanetas, Podolak (2003) calculó la opacidad de los granos utilizando los modelos de envoltura obtenidos por Pollack et al. (de estos modelos obtuvo la distribución de presión, densidad, temperatura y regiones convectivas que caracterizarían la estructura de la atmósfera de un planeta en formación). Notemos que, en este sentido, los cálculos de Podolak no son autoconsistentes puesto que 70 los perfiles atmosféricos de Pollack et al. fueron obtenidos con las tablas de opacidad para la distribución de tamaños de la nebulosa primordial. Por otra parte, Podolak tuvo en cuenta que los planetesimales ingresantes sufren la ablación en su trayectoria hacia el núcleo, con lo cual fueron también considerados como una fuente de opacidad. De sus cálculos Podolak concluye que la opacidad debida a los granos puede ser, como máximo, del orden de 10−1 cm2 g−1 , muy inferior a los valores estándares. El motivo de esta disminución sería la presencia de corrientes convectivas en la atmósfera, que llevarían a los granos desde regiones frías hasta regiones donde la temperatura es lo suficientemente alta como para que los granos se vaporicen. Basados en estos resultados, Hubickyj, Bodenheimer & Lissauer (2005), hicieron varias simulaciones para estudiar en mayor detalle, entre otras cosas, la dependencia del proceso de formación de Júpiter con la opacidad. Dos de sus simulaciones consideran que los valores de la opacidad de los granos son tan solo el 2 % de la opacidad total correspondiente al gas interestelar. En uno de los casos (10L∞) la reducción se aplica en todo el rango del dominio de los granos, mientras que en el otro caso (10V∞) la reducción es del 2 % para T < 350 K, aumenta linealmente en el rango 350 ≤ T ≤ 500 K, y toma el valor interestelar para T > 500 K. Respecto del caso en el que no se aplica ninguna reducción (10H∞), el tiempo de formación se reduce en un factor 3 (de, aproximadamente 6 × 106 años para el caso 10H∞ a ∼ 2 × 106 para 10L∞ ). Por otra parte, la diferencia en el tiempo de formación entre 10L∞ y 10V∞ es de tan solo 5 × 105 años. Estos resultados muestran la sensibilidad de este tipo de cálculos a la opacidad, y, además, que son las capas más externas de la atmósfera del protoplaneta las que dominan la contracción de la envoltura. Sin embargo, recientemente, Dodson-Robinson et al. (2008) realizaron simulaciones de la formación de Saturno, con y sin reducción de la opacidad. A diferencia de los resultados anteriores, estos autores no encuentran que la reducción en la opacidad afecte significati- vamente la escala de tiempo de formación estimada para Saturno. En función de las incertezas en torno al impacto real de las variaciones de la opacidad de los granos, y en vistas de que no existen tablas de opacidades calculadas en forma autoconsistente que avalen determinados valores para la opacidad, en nuestras simulaciones se utilizarán las tablas mencionadas más arriba en su forma original. Es, sin embargo, importante tener presente que es probable que los valores de la opacidad involucrados en este tipo de cálculos sean menores que los tabulados para el medio interestelar. El efecto de una reducción en la opacidad iría en la dirección de acortar los tiempos de formación, lo cual afectaría positivamente a nuestros resultados. 4.4.4. Condiciones de borde Para resolver las ecuaciones diferenciales que describen el problema (ecuaciones (4.23) a (4.26)) se deben aplicar condiciones de contorno adecuadas. Éstas estarán dadas tanto en el límite interior de la envoltura gaseosa como en el exterior. 71 El borde interno de la envoltura está en contacto con el núcleo sólido, por cuanto allí tendremos: Mr = Mc , r = Rc y Lr = Lc = 0. En nuestro traba jo, el núcleo se considera inerte, estando solo caracterizado por la densidad, la cual se fija constante (ρc ≃ 3 g cm−3 ). De este modo, a medida que el núcleo aumenta su masa por la acreción de planetesimales, su radio crecerá según: . (4.51) Rc = 3s 3Mc 4πρc Durante la formación, el planeta se encuentra acretando material de la nebulosa solar, por cuanto las condiciones de contorno externas dependerán del estado de la misma. Ellas estarán dadas por la presión y la temperatura, P = Pnebulosa y T = Tnebulosa . Estas mag- nitudes dependen, en principio, del tiempo y del semieje del planeta. En nuestro modelo consideraremos que tanto la temperatura como la densidad de la nebulosa se mantienen constantes durante todo el proceso de formación, donde los valores correspondientes a la ubicación del planeta se obtienen de las ecuaciones (4.1) y (4.2). Consideraremos sola- mente que la densidad de sólidos en el disco disminuye debido a la acreción por parte del protoplaneta (ver §5.3). Mg , queda determinada por la La tasa de acreción de gas para formar la envoltura, evolución del límite externo del planeta. El borde externo del planeta deberá encontrarse en el radio de acreción o en el radio tidal, según cuál sea el menor de ellos. El radio de acreción para un objeto de masa Mr se define como , Ra = (4.52) GMr c2 s donde cs es la velocidad del sonido en la nebulosa. El radio de acreción marca el límite a partir del cual las partículas tienen una energía térmica mayor que la energía gravitatoria que las liga al planeta, con lo cual no formarían parte del mismo. El radio tidal (o radio de Hill), se define según la ecuación (4.13) y representa el límite para el cual la influencia gravitatoria del planeta es dominante. Más allá del radio de Hill, una partícula estará dominada por el campo gravitatorio de la estrella central. En general, las capas de gas de la envoltura se encuentran en contracción. Con lo cual, cuando el punto más externo del planeta, correspondiente a la masa MR , se contrae hasta un radio menor a Ra (o a RH ), se adiciona masa para satisfacer la condición de que R = Ra (o R =RH , según cuál sea el mínimo). De este modo transcurre la acreción de gas, la cual se produce en forma contínua. Los radios Ra y RH se incrementan debido a que el planeta aumenta su masa, tanto de gas como de sólidos, lo cual agranda la esfera de influencia del planeta. 4.4.5. Interacción entre las capas de gas y los planetesimales in- gresantes En su trayectoria hacia el núcleo, los planetesimales acretados interactúan con la envol- tura gaseosa del protoplaneta intercambiando energía. Para esta Tesis hemos implementado 72 un modelo sencillo que contempla la interacción entre los planetesimales y la envoltura (pa- ra modelos más detallados ver Podolak et al. 1988; Pollack et al. 1996; Alibert et al. 2005). Para simplificar la situación, aceptaremos que los planetesimales que se acercan al pro- toplaneta lo hacen desde el infinito e ingresan a la envoltura con una velocidad vrel (Ec. (4.16)), describiendo luego una trayectoria recta hacia el núcleo. El considerar una trayec- toria recta parece inconsistente con el hecho que para definir Reff calculamos las órbitas de los planetesimales ingresantes. Sin embargo, vale la pena mencionar que el cálculo orbital anteriormente citado se realiza solo hasta que los planetesimales acretados completan la primera revolución dentro de la esfera de Hill. Para el cálculo completo de la interacción energética sería necesario integrar toda la trayectoria hacia el núcleo, en conjunto con una descripción autoconsistente de la energía depositada por los planetesimales en las capas de la envoltura. Esto requiere de modelos mucho más complejos que, por ejemplo, consideren la ablación de los planetesimales. Si bien es cierto que la distribución de la energía es im- portante en los cálculos de la formación planetaria (Benvenuto & Brunini 2008), incorporar este tipo de modelos requiere de idear una buena estrategia numérica para que los tiempos computacionales no se tornen demasiado largos. Como el objetivo de esta Tesis no está focalizado en este punto, nos hemos conformado con un tratamiento simplificado que tiene en cuenta la existencia del intercambio energético entre los planetesimales y el gas, pero a sabiendas que este modelo debe ser profundizado en el futuro. La energía de un planetesimal de masa m en el borde externo de la envoltura es: E = m v 2 rel . 1 2 El arrastre gaseoso y la aceleración de la gravedad son, según nuestro modelo, las dos fuerzas principales responsables de cambiar la energía de los planetesimales una vez que ingresan a la envoltura (ver figura 4.6). La fuerza viscosa que actúa sobre un planetesimal esférico de radio rm , desplazándose a una velocidad constante v , está dada por la ecua- ción Ec. (4.21). La fuerza gravitatoria se calcula de la manera usual (Ec. (4.20)). Ambas fuerzas actúan simultáneamente y modifican la velocidad de los planetesimales capturados. Entonces, la fuerza total sobre un planetesimal es: (4.53) y ~f = ~fD + ~fG = −m dv dv dt dr con lo cual, la variación de la velocidad de un planetesimal es: = v r , dv dt (4.54) (4.55) = − Ahora bien, queremos calcular la energía que intercambian los planetesimales con el gas que atraviesan. La energía que las capas de gas absorben, en forma de calor, corresponde (4.56) + . CD π r2 m ρg v 2m GMr r2 v dv dr 73 a la pérdida de energía cinética de los planetesimales por fricción. Entonces, si la variación de la velocidad debido a la viscosidad del medio es: dv dt 1 2 CDπ r2 m m = − ρg v 2 , (4.57) la variación de energía cinética será: = − Pero más que la variación por unidad de tiempo, nos interesa cuantificar la pérdida de energía en las capas de gas: CDπ r2 m ρg v 3 . (4.58) = mv dEk,m dt dv dt 1 2 dEk,m dt dt dr dr dt = dEk,m dr v , (4.59) entonces, 1 dEk,m dEk,m dr dt v Con lo cual, teniendo en cuenta que la energía cinética perdida por un planetesimal debido a la fricción con el gas se transforma en calor, tenemos: (4.60) = . dE dr dEk,m dr 1 2 = − dr es el calor absorbido por las capas de la envoltura. Esta cantidad ingresa, trans- donde dE formada en las unidades adecuadas, en las ecuaciones de estructura a través de ǫac . El tratamiento numérico de este problema se detallará en el siguiente capítulo. CD πr2 m ρg v 2 , (4.61) = 74 Figura 4.6. Esquema de un planetesimal que interactúa con la envoltura gaseosa del planeta. En nuestro modelo consideramos que la velocidad del planetesimal se ve afec- tada por la acción de la gravedad y la viscosidad del gas. La energía que pierde el planetesimal se transforma en calor absorbido por la envoltura. 75 76 Capítulo 5 Tratamiento numérico del modelo de formación de planetas gigantes Las ecuaciones diferenciales introducidas en el capítulo anterior no tienen solución ana- lítica, por cuanto hay que recurrir a métodos numéricos para resolverlas. La idea básica al hacer ésto es generar una secuencia de modelos: cada uno de ellos representa el estado de la estructura del planeta para un determinado instante de tiempo t. En la práctica, la estructura gaseosa se representa dividida en cáscaras concéntricas, donde se calcula para un tiempo t y para cada una de estas cáscaras, las cantidades termodinámicas P , T , L, etc. Finalizado el cálculo del modelo, nos valemos de estos resultados para estimar el próximo modelo correspondiente a un ∆t posterior. La implementación de este tipo de resolución está basada en el método de Henyey (Henyey et al. 1959, 1964) según la propuesta de Kippenhahn, Weigert y Hofmeister (1967). 5.1. Llevando las ecuaciones diferenciales a ecuacio- nes en diferencias Benvenuto & Brunini (2005) desarrollaron y publicaron las ideas básicas concernientes al código utilizado en esta Tesis para el cálculo de la formación de planetas gigantes. Respecto de su versión original, el código fue varias veces actualizado, en la medida en que se fueron incorporando ingredientes nuevos al modelo físico. Sin embargo, el esquema básico y la metodología utilizada siguen siendo los originales. El planteo del problema puede ser visualizado de la siguiente manera: el planeta está formado por un núcleo sólido y una envoltura gaseosa, y circundado por la nebulosa solar. Cerca del núcleo, la densidad del gas estará afectada por el campo gravitatorio del núcleo y por la energía liberada por los planetesimales acretados, con lo cual la densidad de la envoltura será distinta a la de la nebulosa. Además, el límite exterior del planeta no 77 existe físicamente, sino que se lo define como el mínimo entre el radio de acreción y el radio de Hill para “separar” lo que se considera como parte del planeta de lo que forma parte de la nebulosa, ya que la componente gaseosa del planeta y el gas nebular forman un continuo. Debido a la hipótesis de simetría esférica, el problema puede ser descripto en forma unidimensional solo con la coordenada radial r (o, en realidad, con Mr ya que estamos haciendo un tratamiento lagrangiano). La coordenada r tiene por origen al radio del núcleo, Rc , que es donde comienza la envoltura del planeta y se extiende hasta el infinito, o sea que r supera el límite del planeta. Es conveniente que esto sea así porque el límite externo del planeta se modifica modelo a modelo, en general aumentando, lo que significa que el planeta crece porque incorpora gas nebular. Para que el tratamiento sea suave, es conveniente que las variables termodinámicas especifican el estado del gas incorporado (gas que en el modelo anterior “pertenecía” a la nebulosa y ahora “pertenece” al planeta) sean continuas y diferenciables cerca del límite del planeta. Es por esto que las ecuaciones diferenciales (Ecs. 4.23–4.26) se integran más allá del borde del planeta. Por razones de estabilidad numérico, la variable independiente Mr y las variables de- pendientes r , P , y T son inconvenientes para la integración de las ecuaciones diferenciales que describen la estructura gaseosa y resulta necesario realizar un cambio de variables. El cambio de variables crucial para la resolución numérica de las ecuaciones resulta de reemplazar a la variable independiente Mr por: ξ ≡ ln (cid:18) Mr Mc − 1(cid:19) = ln (cid:18) Mg Mc (cid:19) , donde Mg es la masa de gas del planeta. Esta transformación resulta vital para lograr la convergencia del código en todo momento dado que elimina la necesidad de hacer in- terpolaciones para considerar la migración del gas desde la nebulosa hacia el interior del planeta. Es importante destacar que, a diferencia del caso estelar donde usualmente la masa es constante y la estructura de gas se extiende desde el centro de la estrella hasta el borde, la situación de un planeta en formación es bien diferente. Por empezar, la com- ponente gaseosa comienza en la superficie del núcleo y no en el centro del planeta, la cual esta condición de borde no es estática sino que se desplaza hacia afuera a medida que el núcleo aumenta de tamaño, empujando con ella a las capas de gas. Pero además, la masa del planeta no es constante sino que aumenta (aunque pueden ocurrir eventos de expulsión de masa, como veremos en el capítulo 6) conforme pasa el tiempo. Estas características del problema (borde interno no fijo más masa total no constante) hacen que el cambio de variables adecuado para la masa no sea trivial. (5.1) Notemos que, por su definición, ξ no es independiente de t. La coordenada ξi representa dos elementos de masa diferentes en dos modelos consecutivos dado que la masa del núcleo creció en ese intervalo del tiempo. De esta manera, las derivadas temporales (como la que aparece en la ecuación de balance energético), que son las que conectan el estado termodinámico del gas entre dos modelos, deben “seguir” adecuadamente a un mismo elemento de masa de gas. Entonces, al reescribir las ecuaciones diferenciales hay que tener 78 , . en cuenta que: = + ∂ t (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ξ ∂ ∂ = − ∂ t ∂ t (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Mg ∂ ξ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)t ∂ ∂ ξ Mc ln Mc = − Mc ∂ t (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Mg ∂ donde de la ecuación 5.1 se obtiene que: ∂ t (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Mg ∂ ξ Si bien es cierto que al hacer este cambio de variables se gana una singularidad en Mr = Mc , (ξ → −∞), ésta puede ser salvada imponiendo que la integración llegue sólo hasta un valor pequeño para ξ que nos permita acercarnos a Mc tanto como queramos, siendo el error cometido en esta aproximación despreciable (en nuestras simulaciones hemos adoptado ξ = −19). Consideraremos que ξ se extiende hacia afuera hasta, aproximadamente, ξ = 20, lo cual supera ampliamente el valor de ξ correspondiente a la masa final del planeta para todos los casos considerados en esta Tesis (observemos que el borde externo del planeta yace en algún punto ξ < 20). Los otros cambios de variables utilizados son: x ≡ ln r θ ≡ ln T p ≡ ln P L se sigue considerando una variable lineal. Para hacer la integración debemos reemplazar las ecuaciones diferenciales por ecuacio- nes en diferencias. Para ello dividimos a la componente gaseosa en m cáscaras concéntricas de espesor variable, que van desde Mc hasta algún valor M (o desde ξm hasta ξ0). Tendre- mos m − 2 capas interiores (cuyos bordes corresponden a ξ1 , ξ2 · · · ξm−1) y las dos capas restantes en los extremos: una linda con el borde interior del planeta (la capa que va desde ξm hasta ξm−1 ), y la otra corresponde a algún punto exterior al planeta, en la nebulosa solar (la que va desde ξ1 hasta ξ0). Los puntos en los cuales calculamos las soluciones de las ecuaciones constituyen la grilla de integración. Luego, podemos expresar las ecuaciones (4.23–4.26) como ecuaciones en diferencias para estas capas o puntos de la grilla. Mostraremos un ejemplo de cómo se hace esto, utilizando para ello la ecuación de continuidad. Primero expresamos a la Ec. (4.23) en función de las nuevas variables: ∂x ∂ ξ = 1 4π Mc ρg e−3x eξ . Luego, esta ecuación escrita en diferencias, resulta en: xj+1 − xj ξj+1 − ξj −3xj+ 1 ξj+ 1 e 2 e 2 Mc ρg j+ 1 2 1 4π = 1 ≤ j ≤ m − 2 79 donde j representa un punto de la grilla y las magnitudes con subíndice j + 1 2 son calculadas de la siguiente manera: = 1 bj+ 1 2 2 con bj+1 y bj correspondientes al modelo que se está iterando (vemos que los bj+ 1 2 ser un promedio de la cantidad b para la capa considerada). (bj+1 + bj ), resultan La evolución temporal se calcula de la siguiente manera. Sea Y alguna de las magnitudes termodinámicas, definimos: para magnitudes logarítmicas (y ≡ ln Y , por ejemplo x ≡ ln r), Yprev ! = yprev + ln(1 + δy ) y ≡ ln Y = ln(Yprev + δY ) = ln Yprev + ln 1 + δY y, para magnitudes lineales (como L) y ≡ Y = Yprev + δY , donde el subíndice “prev” se refiere al valor de la variable en el paso temporal previo. Se tiene, entonces, el modelo correspondiente al tiempo t y se quiere calcular el modelo para t + ∆t. El modelo al tiempo t estará caracterizado por las magnitudes yprev . Con estos valores y las actualizaciones de las condiciones de contorno, y de otros parámetros del modelo que no dependen de las ecuaciones de estructura, se calcula δy . El valor de δy se mejora iterando las ecuaciones tantas veces como sea necesario hasta alcanzar una dada precisión. Concluido este proceso, se define el modelo que caracteriza el estado en t + ∆t como yprev + δy (o, yprev + ln(1 + δy ), según corresponda). De este modo, las cantidades que se iteran son δx, δθ, δp y δL, y no x, θ, p y L ya que esto facilita el cálculo preciso de las derivadas ∂ ∂ t ξ y provee estabilidad numérica en todo momento. Repitiendo el procedimiento para las cuatro ecuaciones tendremos, para cada uno de los m − 2 puntos de la grilla, cuatro ecuaciones en diferencias. Matemáticamente, podemos escribir lo anterior de la siguiente manera: Gj i (pj , θj , xj , Lj , ξj ; pj+1 , θj+1 , xj+1 , Lj+1 , ξj+1) = 0 (5.2) i = 1, · · · , 4 j = 1, · · · , m − 2. Ahora bien, las incógnitas son pj , θj , xj , Lj para 1 ≤ j ≤ m − 1, y pm , θm (ya que Lm y xm son las condiciones de borde internas), con lo cual tenemos 4(m − 1) + 2 incógnitas y 4(m − 2) ecuaciones; luego faltan seis ecuaciones más. 80 Dos de ellas provienen de las condiciones de borde externas (p1 = ln(Pnebulosa) y θ1 = ln(Tnebulosa)). Simbólicamente escribimos esto como: B1 (p1 , θ1 , x1 , L1 ) = 0 B2 (p1 , θ1 , x1 , L1 ) = 0. (5.3) (5.4) las cuatro ecuaciones restantes (que se aplican en la capa delimitada por ξm y ξm−1 ) son las mismas ecuaciones Gi pero que ahora solo dependen de las incógnitas pm−1 , θm−1 , xm−1 , Lm−1 , pm , θm , y que renombramos: Ci (pm−1 , θm−1 , xm−1 , Lm−1 ; pm , θm ) = 0 (5.5) i = 1, · · · , 4. Queda entonces planteado el sistema de 4(m− 1) + 2 ecuaciones con 4(m− 1) + 2 incógnitas. Notemos que hemos definido originalmente una grilla con m capas pero solo hemos dado las condiciones que deben satisfacer m de sus puntos (con lo cual faltaría especificar uno más, que corresponde al límite exterior). Sin embargo dado que las ecuaciones que caracterizan las condiciones de borde externas son por definición ecuaciones en diferencias (y no ecuaciones diferenciales discretizadas), las condiciones de borde pueden aplicarse a los puntos de la grilla correspondientes a ξ1 , ya que en ξ0 tomarán el mismo valor. Condiciones de borde externas Como ya mencionamos, las condiciones de borde externas (P = Pnebulosa, T = Tnebulosa) no se aplican en el límite del planeta sino en algún punto de la nebulosa protoplanetaria exterior al planeta. Desarrollaremos con un poco más de detalle este punto. El estado de la nebulosa solar está perturbado por la presencia del planeta hasta, aproximadamente, su radio de Hill (el cual, durante gran parte del proceso de formación, define el borde del planeta). Más allá del radio de Hill consideraremos que el gas nebular está caracterizado por dos parámetros constantes en el tiempo: la presión y la temperatura. Por otra parte, como mencionamos en § 4.4.1, el potencial gravitatorio considerado no es el correspondiente al problema de dos cuerpos sino que se encuentra ablandado por el RH (cid:17)3 , con el objetivo de considerar la influencia solar en el campo gravitatorio término 1 − (cid:16) r del planeta. La introducción de este ablandamiento hace que, para r > RH , el potencial cambie de signo y entonces deje de tener sentido dentro de las hipótesis que estamos considerando. Por esto es que para radios que superan RH el potencial se hace tender suavemente a cero (Ec. 4.4.1). De la ecuación de equilibrio hidrostático (Ec. 4.24, o en su forma correspondiente a la ecuación 4.32) vemos que si el potencial es estrictamente cero, el gradiente de presión también lo es, con lo cual más allá del radio de Hill la presión es constante. De combinar las ecuaciones de equilibrio hidrostático y de transporte vemos, además, que el gradiente de presión es proporcional al gradiente de temperatura, 1 T ∂ ln T ∂Mr = 1 P ∂ ln P ∂Mr ∇. 81 (5.6) De este modo, el gradiente de temperatura será cero cuando lo sea el gradiente de presión, y la temperatura también será constante más allá del radio de Hill. Así, las condiciones de borde externas se pueden imponer en cualquier punto de la grilla bien entrada la nebulosa y serán satisfechas automáticamente en la posición del radio de Hill. Este procedimiento facilita la integración numérica, ya que a medida que el planeta crece y aumenta el radio de Hill las variables termodinámicas no experimentan ningún cambio abrupto, sino que lo hacen suavemente y con continuidad. 5.2. Resolución de las ecuaciones Una vez que las ecuaciones diferenciales son transformadas en ecuaciones en diferen- cias deben ser resueltas mediante algún método numérico. Es importante notar que las ecuaciones no son lineales. Para la resolución, el programa de computadora que calcula la formación de planetas gigantes utiliza el método de Henyey, de amplio uso en el cálculo de evolución estelar. El método de Henyey utiliza el procedimiento de resolución del método de Newton y, en función de la estructura final del sistema de ecuaciones, propone una metodología para encontrar la solución haciendo uso de la forma particular que tiene la matriz que representa al sistema. A continuación daremos una explicación esquemática del mismo siguiendo a Kippenhahn, Weigert & Hofmeister (1967). El método de Newton Sean y1 , y2 , · · · , yn n incógnitas que satisfacen un sistema de n ecuaciones no lineales: E1 (y1 , · · · , yn) = 0 ... En (y1 , · · · , yn) = 0. Supongamos que conocemos una solución aproximada que llamaremos y 0 n . Dado 2 , · · · , y 0 1 , y 0 n) 6= 0, con i son aproximaciones tendremos, en general, Ek (y 0 que, justamente, los y 0 1 , · · · , y 0 k = 1, · · · , n. Se quiere mejorar la aproximación inicial, obteniendo una corrección de la i . De este modo, a primer orden, misma, que llamaremos δy 0 n ) = Ek (y1 , · · · , yn) + δE 0 n + δy 0 1 , · · · , y 0 1 + δy 0 Ek (y 0 k , (5.7) con ∂yi !0 k = Xi ∂Ek δE 0 y, como se busca perfeccionar la solución, se pide que ella satisfaga el sistema original, con lo cual: (5.8) δy 0 i k + δE 0 E 0 k = 0 82 (5.9) entonces, k = −E 0 δE 0 k , (5.10) y, usando la expresión para δE 0 k , ∂yi !0 Xi ∂Ek k y ( ∂Ek De este modo, como E 0 )0 son cantidades conocidas, se puede despejar δy 0 i . Luego, ∂ yi la solución mejorada es: k = 1, · · · , n. i = −E 0 δy 0 k (5.11) i + δy 0 i = y 0 y 1 i = 1, · · · , n. (5.12) i Este procedimiento se repite nuevamente con y 1 i (que pasa a tomar el lugar de y 0 i ), y así sucesivamente, hasta que δE s k (para todo k) sea menor que una tolerancia dada. Aplicación del método de Newton a las ecuaciones de estructura. El sistema que tenemos que despejar es: B1 (p1 , θ1 , x1 , L1 ) = 0 B2 (p1 , θ1 , x1 , L1 ) = 0 Gj i (pj , θj , xj , Lj ξj ; pj+1 , θj+1 , xj+1 , Lj+1 , ξj+1) = 0 i = 1, · · · , 4 j = 1, · · · , m − 2 (5.13) (5.14) (5.15) Ci (pm−1 , θm−1 , xm−1 , Lm−1 ; pm , θm ) = 0 i = 1, · · · , 4. Como mencionamos, para poder resolverlo, hay que dar una aproximación inicial a la solución p0 j , a partir de la cual se buscarán las correcciones δp0 j . En j , δL0 j , δx0 j , δθ0 j , L0 j , x0 j , θ0 la práctica, se genera un modelo inicial muy sencillo que surge de las condiciones de borde que impone la nebulosa, y del modelo del gas ideal1 . El sistema lineal de ecuaciones queda, entonces, planteado de esta manera: ∂p1 !0 ∂L1 !0 ∂Bi 1 + ∂Bi δL0 ∂x1 !0 1 + ∂Bi δp0 δθ0 1 = −(Bi )0 (5.17) (5.16) ∂θ1 !0 1 + ∂Bi δx0 i = 1, 2 1No es necesario, en la mayoría de los casos, que el modelo inicial sea muy preciso ya que el algoritmo empleado no es muy sensible a las condiciones iniciales, y al cabo del cálculo de algunos modelos se consigue la independencia de las mismas. Es mucho más importante generar un modelo inicial que garantice la convergencia del código. 83 δx0 j + ∂Lj !0 ∂pj !0 ∂xj !0 ∂Gj j + ∂Gj j + ∂Gj i i i δL0 δp0 ∂θj !0 ∂Lj+1 !0 ∂pj+1 !0 j + ∂Gj δLj+1 + ∂Gj + ∂Gj i i i δp0 δθ0 j+1 + ∂xj+1 !0 ∂θj+1 !0 j+1 + ∂Gj + ∂Gj i i δx0 i = 1, ..., 4 y j = 1, ..., m − 2 j+1 = −(Gj i )0 δθ0 (5.18) ∂Lm−1 !0 ∂Ci ∂pm−1 !0 m−1 + ∂Ci δL0 ∂xm−1 !0 ∂θm−1 !0 m−1 + ∂Ci m−1 + ∂Ci δp0 δx0 ∂pm !0 ∂θm !0 m + ∂Ci + ∂Ci δθ0 δp0 m = −(Ci )0 (5.19) δθ0 m−1 + i = 1, ..., 4. Podemos escribir este sistema en forma matricial: )0 · · · ( ∂B1 ∂x1 ...     La matriz de coeficientes tiene, esquemáticamente, la siguiente forma: =   −B1 ...     δx0 1... δθ0 m   ( ∂C4 ∂ θm )0 −C4 84 (5.20) . . . c1 c2 c3 c4 b1 b1 b1 b1 b2 b2 b2 b2 g1 g1 g1 g1 g1 g1 g1 g1 g2 g2 g2 g2 g2 g2 g2 g2 g3 g3 g3 g3 g3 g3 g3 g3 g4 g4 g4 g4 g4 g4 g4 g4 . . . . . .     Básicamente, la matriz es una matriz diagonal por bloques. Se observan tres estructuras de bloques diferentes: la primera formada por un bloque de seis filas constituido por dos subestructuras (de elementos representados esquemáticamente por bi y gi ); la segunda estructura se compone por m − 3 bloques como el correspondiente a gi solamente, y la tercera es el bloque de los elementos representados por ci . Hemos representado con bi las derivadas de Bi , con gi las derivadas de Gi , y con ci a las de Ci . Henyey propuso un método para resolver el sistema, aprovechando la forma particular de la matriz. El detalle del método puede encontrarse en el artículo de Kippenhahn, Weigert & Hofmeister (1967). En resumen, podemos decir que las correcciones se obtienen de atrás m−1 · · · hasta obtener, por último, m , δp0 m , δθ0 hacia adelante, o sea, primero se encuentra δp0 1 . Este procedimiento se repite hasta que las correcciones sean menores que una δp0 1 , · · · , δx0 tolerancia dada o hasta que lo sea δEk para todo k . . . c1 c2 c3 c4 . c1 c2 c3 c4 c1 c2 c3 c4 c1 c2 c3 c4 c1 c2 c3 c4 5.3. Tratamiento de la acreción de sólidos En las secciones previas mostramos el método de resolución de las ecuaciones diferen- ciales que describen el estado de la estructura gaseosa del planeta. Como ya mencionamos, el núcleo sólido se considera inerte, solo caracterizado por su densidad constante. Obvia- mente, con esta simplificación se están despreciando, entre otras cosas, cualquier proceso de liberación de energía que pueda ocurrir en el propio núcleo. Luego, la presencia del núcleo tiene solo un efecto gravitatorio en este problema. Sin embargo, los planetesimales acretados, que incrementarán la masa del núcleo a través del tiempo, juegan un papel fun- damental en la energética del problema. Por este motivo, describiremos aquí el tratamiento numérico de la acreción de sólidos. 85 El protoplaneta incorpora material de la nebulosa solar. La región de la cual el proto- planeta puede acretar material se conoce como zona de alimentación y corresponde a un anillo centrado en la órbita del planeta. La extensión considerada para la zona de alimen- tación varía dependiendo de los autores, ya que no se calcula analíticamente sino que se estima en forma numérica. En general suele considerarse que se extiende entre 3 y 5 RH a cada uno de los costados de la órbita del planeta. Para este traba jo hemos considerado un valor de 4 RH , por cuanto la zona de alimentación forma un anillo cuyo ancho total es de 8 RH , centrado en la circunferencia de radio igual al radio orbital del planeta. La tasa de acreción de sólidos está dada por la ecuación (4.14). Esta ecuación depende de varios factores, entre ellos de la densidad superficial de sólidos. La densidad en la nebulosa sigue una ley de potencias con la distancia, por cuanto calcularemos su valor medio en la zona de alimentación para poder obtener la tasa de acreción. El valor medio de la densidad superficial de sólidos es: hΣi = R aint aout Σ(a)2πad a π(a2 int) out − a2 donde aout y aint representan los límites externo e interno de la zona de alimentación. Como tenemos que (5.21) , (5.22) Σ(a) = Σ0 a−p , el valor medio para la densidad superficial de sólidos es2 : out − a2−p 2 − p a2−p 2 Σ0 int ! . int hΣi = a2 out − a2 Notemos que estamos considerando que, en cada instante, la nebulosa está suficien- temente rela jada como para que el material disponible en la zona de alimentación esté distribuido de manera uniforme. Ahora bien, el núcleo del planeta crece a expensas de acretar material de su zona de alimentación, por cuanto esto modifica el valor de hΣi. Lla- memos ∆A al área de la superficie de la zona de alimentación en el modelo correspondiente al tiempo t, (5.23) out − a2 ∆A = π(a2 int ), (5.24) y sea ∆ A el área correspondiente al modelo anterior (en t − ∆t). Entre ambos modelos el planeta incorporó masa (tanto en forma de sólidos como de gas), por cuanto el radio de Hill creció y los límites de la zona de alimentación se expandieron. Sea δa el incremento en el ancho del anillo, tanto hacia el interior como hacia el exterior (ver figura 5.1. Notemos 2Notemos que acá estamos suponiendo que Σ es una función continua. Esto no es cierto si la zona de alimentación cruza la línea de hielo. De todos modos, en las simulaciones consideradas para este trabajo, siempre nos mantendremos bastante lejos de esta situación. 86 δr ∆ R rδ Planeta Figura 5.1. Representación esquemática de la zona de alimentación del planeta. En el instante t − ∆t el ancho de la zona de alimentación es ∆R. Para un instante posterior t, dado que el planeta incrementó su masa y que la zona de alimentación es proporcional a RH , sus límites se incrementan en δr. que lo que estamos llamando δa en el texto se representa con δr en la figura). Asociado a δa hay un aumento en el área del anillo que llamaremos δAout y δAint , según corresponda al incremento exterior o interior. En cada uno de estos pequeños anillos que se incorporan a la zona de alimentación, la densidad de la nebulosa es la original para esa posición. Notar que no se considera la migración de los planetesimales por fricción con el gas nebular. Consideramos que el material en la zona de alimentación del planeta se modifica por la acreción y fuera de ella se mantiene intacto ya que no consideramos ningún tipo de evolución nebular. Luego, si la densidad en t − ∆t era hΣ(t − ∆t)i, la densidad superficial de sólidos media en la zona de alimentación al tiempo t será: ∆ A δAint δAout hΣi = ∆A hΣint i. ∆A hΣout i + ∆A hΣ(t − ∆t)i + Si llamamos ∆M a la masa de sólidos que acretó el planeta en el intervalo ∆t, el valor final de la densidad media en la zona de alimentación para el modelo al tiempo t será: (5.25) ∆M hΣ(t)i = hΣi − ∆A Repitiendo este procedimiento con cada nuevo modelo se va modificando la densidad superficial de sólidos, tanto por la acreción de material por parte del planeta, lo cual provoca (5.26) . 87 una disminución en la densidad, como por la expansión de la zona de alimentación, que permite la incorporación de sólidos disponibles para ser acretados. Si, por algún motivo, el planeta perdiera material (que solo podría ser gas de la envoltura), el esquema anterior dejaría de ser válido. Para evitar esta inconsistencia, supondremos que si el planeta pierde masa, la zona de alimentación no modifica su tamaño, con lo cual estamos suponiendo que la zona de alimentación no puede decrecer. Esto es, si la masa del planeta disminuye, si bien su radio de Hill se reduce, consideraremos que la zona de alimentación mantendrá el tamaño previo. Esto se puede justificar de la siguiente manera: los planetesimales que se encuentran en la zona de alimentación son objetos cuyas órbitas sufren las perturbaciones debidas a la presencia del planeta de modo que, tarde o temprano, terminarán cruzando la órbita del planeta, lo cual les da la posibilidad de ser acretados. Si el RH se reduce, eso no implica que la zona de alimentación deba estrecharse puesto que los planetesimales que en algún momento formaron parte de la zona de alimentación ya fueron excitados por el protoplaneta y seguirán teniendo probabilidad de ser acretados por el mismo. Otro factor que influye en el cálculo de la tasa de acreción es el radio efectivo del pro- toplaneta. El radio efectivo es el máximo valor que puede tomar el parámetro de impacto de un planetesimal para ser capturado por el protoplaneta. Por otra parte, el valor mínimo posible para Reff es el debido al enfocamiento gravitatorio (Ec. 4.15), donde no se conside- ra la viscosidad del gas. Sea, entonces, un planetesimal que viene desde el infinito en una trayectoria recta, de la cual se ve deflectado por la presencia del protoplaneta. Suponga- mos que, en su trayectoria, el planetesimal ingresa en la atmósfera del protoplaneta. Para determinar el radio efectivo de captura debemos encontrar cuál es el parámetro de impacto más grande para el cual ésto ocurre. En el tratamiento numérico, la búsqueda del radio efectivo se hace de la siguiente manera: dado el Rprev del modelo anterior, proponemos eff como valor inicial para la búsqueda del radio efectivo del modelo que queremos calcular (Reff ) una fracción del valor anterior (Reff = α Rprev eff ). Si bien lo correcto sería comenzar la búsqueda a partir del radio mínimo, dado solo por el enfocamiento gravitatorio (R∗ eff ), estos dos valores son tan diferentes (R∗ eff ≪ Reff , ver figura 4.3), que el costo computacional sería enorme. Por otra parte, a menos que haya algún cambio abrupto en la densidad del gas (cosa que en general no ocurre), el radio efectivo siempre aumenta o, si disminuye (como en el caso de los planetas que pierden masa), el decrecimiento siempre es menor que el propuesto como valor inicial para el cálculo. Entonces, tomamos el valor previo del radio efectivo, Rprev eff , lo reducimos en una fracción (generalmente entre un 5 y un 10 %), y bus- camos el nuevo valor de Reff para el modelo que estamos calculando. Para eso, integramos la trayectoria de un planetesimal desde el radio de Hill del planeta, moviéndose con una velocidad inicial dada por la ecuación (4.16). La integración de la ecuación de movimiento, considerando la aceleración de la gravedad y la fricción producida por el gas, se hace con un integrador Runge Kutta Fehlberg (Press et al. 1992). Si el planetesimal describe una trayectoria cerrada dentro de la esfera de Hill del planeta, se interrumpe la integración y se lo considera acretado. Se aumenta el valor del parámetro de impacto (en un factor que, usualmente, tomamos en 5 × 10−4 del radio efectivo) y se procede nuevamente. Esto se repi- te hasta que las trayectorias que describen los planetesimales dejen de estar contenidas en 88 la esfera de Hill del planeta. El mayor valor que toma el parámetro de impacto es el que se asigna como radio efectivo. Por la naturaleza del problema, y dado que el cálculo del radio efectivo se basa en considerar el frenado producido por el gas, cuyo estado está dominado por la evolución del planeta, el parámetro de impacto máximo posible es el dado por el radio de Hill. Si estamos en una fase de la formación en la cual el radio del planeta está definido por el radio de acreción (Rp = Ra ) puede ocurrir que Reff > Rp . Dado que estos planetesimales terminan cayendo siempre hacia el centro del planeta no hay contradicción en este punto, siendo perfectamente posible que en estos casos el radio efectivo sea mayor que el del planeta. Una vez que queda establecida la tasa de acreción de sólidos, debemos calcular cuánta energía depositan los planetesimales en las capas de gas de la envoltura. Como dijimos en § 4.4.5, consideramos que los planetesimales acretados describen una trayectoria recta entre el borde del planeta y el núcleo. A lo largo de este camino, la velocidad de los planetesimales varía por la acción de la viscosidad del gas, que los frena, y de la fuerza gravitatoria que los acelera hacia el centro. La variación de la velocidad con el radio está dada por la ecuación (4.56) que, si la multiplicamos por la velocidad v , nos da: v dv dr CD π r2 m ρg v 2 2m = − + GMr r2 . (5.27) Sean ri y ri+1 los valores de los radios correspondientes a dos puntos adyacentes de la grilla y supongamos que, entre estos puntos pueden considerarse constantes todas las variables de la ecuación anterior que no involucren a la velocidad. Esto es, si y C1 = CD π r2 m ρg 2m C2 = GMr r2 son consideradas constantes, podemos escribir, = −C1 v 2 + C2 . Esta ecuación tiene solución analítica, lo cual nos permite expresar v dv dr (5.28) (5.29) (5.30) i+1 = v 2 v 2 i e−2 C1 ∆ri + C2 C1 (cid:16)1 − e−2 C1 ∆ri (cid:17) , donde definimos vi ≡ v (ri ) y ∆ri ≡ ri+1 − ri . Así, obtenemos la expresión para la velocidad con la cual un planetesimal atraviesa una capa de gas. Por efecto de la fricción con el gas de la envoltura, el planetesimal pierde parte de su energía cinética en forma de calor. La ecuación (4.61) corresponde al intercambio de energía entre un planetesimal con el gas que lo rodea. Si tenemos en cuenta todos los planetesimales ingresantes por unidad de tiempo, (5.31) 89 Mc , y transformamos las ecuaciones diferenciales en ecuaciones en diferencias, la variación total de la energía ∆Ei de la capa de gas i en el intervalo de tiempo ∆t es: ∆Ei = 1 2 CD π r2 m ρg Mc m v 2 i ∆ri ∆t. (5.32) Finalmente, cuando un planetesimal llega al núcleo, toda su energía cinética remanente se deposita en la capa de gas adyacente a él. 5.4. Cálculo de la acreción de gas La tasa de acreción de gas es una consecuencia de la condición de borde que define el límite del planeta. Como mencionamos en § 4.4.4, no existe una separación física entre el planeta y la nebulosa, por cuanto el radio del planeta debe ser definido. Es una norma general tomar como definición: Rp = min[RH , Ra ]. (5.33) La búsqueda de Rp no es trivial ya que tanto RH como Ra dependen de la masa del planeta, la cual no se conoce hasta tanto no se sabe donde está el borde externo del mismo. De la resolución de las ecuaciones de estructura se obtiene, para cada punto de la grilla, el valor del radio r asociado a la masa contenida hasta r , Mr . Para explicar el procedimiento, supongamos que R generaliza el concepto del radio del planeta (con lo cual R puede ser Ra o RH ). Entonces, comenzando por el borde externo de la grilla, calculamos para cada punto de la grilla (que tiene asociado un valor de la masa Mr ), el valor de R. Recordemos que tanto Ra como RHdependen de la masa total del planeta. En general, el valor de R será distinto del de r correspondiente a ese punto. Se calcula para cada punto la diferencia entre ambos radios. Dado que tanto r como R son monótonamente crecientes con la masa, la cantidad r − R tendrá una única raíz en toda la grilla. Esa raíz estará entre dos puntos consecutivos de la grilla para los cuales r − R cambie de signo. Luego, el valor de R se obtiene de interpolar al punto medio entre ambos. Este procedimiento se realiza tanto para el radio de acreción, Ra , como para el radio de Hill, RH . El mínimo entre ambos definirá el radio del planeta, Rp , para el modelo que estemos calculando. La variación, modelo a modelo, del borde del planeta es lo que determina la cantidad de gas por unidad de tiempo que es acretado por el planeta. El paso de tiempo de integración y la cantidad de puntos de la grilla no son constantes sino que varían a lo largo de todo el proceso, dependiendo de la precisión temporal o espacial que requiera el cálculo para un determinado estado del planeta. En las regiones en que las cantidades termodinámicas tienen un comportamiento suave muchas veces no es necesario que el grillado sea muy fino. En esos casos, se eliminan capas y se reestructura el grillado, reasignando las cantidades. Si, por el contrario, deben agregarse puntos, se calculan los valores de las cantidades termodinámicas en los puntos nuevos haciendo una 90 interpolación lineal. En cuanto al paso de tiempo, su valor depende de cuán dificultosa sea la integración de un modelo. El algoritmo que determina cómo debe actualizarse el paso de tiempo depende del cambio máximo permitido para la luminosidad, la temperatura, la presión y el radio entre dos modelos, y el cambio que efectivamente fue calculado. Si el cambio ocurrido es menor que el permitido, el paso de tiempo aumenta; en el otro caso disminuye. 91 92 Capítulo 6 Resultados En este capítulo presentaremos los resultados obtenidos con nuestro programa de for- mación de planetas gigantes. Durante el desarrollo de esta Tesis se realizaron numerosas simulaciones, considerando diversas situaciones que involucran: la posición del planeta en el disco, la densidad superficial de sólidos en la zona de alimentación del planeta y el tamaño de los planetesimales acretados. Los resultados obtenidos al variar estos parámetros moti- varon que el modelo siguiera evolucionando en el sentido de la incorporación de una ley de potencias para la distribución de tamaños de los planetesimales acretados. El considerar que la población de planetesimales no sea uniforme hace que el aborda je del problema sea más realista pero, a la vez, el tratamiento numérico se vuelve más complejo y el tiempo de cálculo aumenta considerablemente. Esto reduce la posibilidad de hacer simulaciones que consideren una gran diversidad de casos. Sin embargo, los resultados pusieron de manifiesto que introducir una distribución de tamaños era un ingrediente fundamental en el cálculo de formación de planetas gigantes, algo que hasta el momento no se había tenido en cuenta. Finalmente, se consideró la arquitectura original del Sistema Solar según lo propuesto por el modelo de Niza (Tsiganis et al. 2005), y se calculó la formación de Júpiter, Saturno, Urano y Neptuno. Los resultados muestran una concordancia muy buena entre las masas de los núcleo obtenidas y las estimadas a partir de los datos observacionales, como así también entre los tiempos de formación y las cotas observacionales dadas por las escalas de vida de los discos circumestelares. Es importante destacar que, en algunas simulaciones, se encontraron eventos cuasi-periódicos de expulsión de masa gaseosa. Este fenómeno no tiene antecedentes en la bibliografía y es por demás interesante puesto que su aparición puede retrasar o, directamente interrumpir la formación de un planeta. 93 6.1. Simulaciones considerando una población de pla- netesimales de tamaño único Todas las simulaciones que presentaremos en este capítulo tienen en común que fueron realizadas ba jo las hipótesis del modelo que presentamos en los capítulos anteriores. A con- tinuación sintetizaremos sus características principales. El embrión planetario se encuentra en órbita circular alrededor del Sol, siendo el Sol el único cuerpo que tiene influencia gravi- tatoria sobre el planeta en formación. No se consideran efectos magnéticos, de rotación o la presencia de otros embriones creciendo en la vecindad. El disco protoplanetario es estático, así que no se contempla la posibilidad de migración de los planetesimales, ni los procesos que operan para la disipación del disco (como, por ejemplo, la fotoevaporación). Tampo- co se consideró la interacción gravitatoria entre el disco y el protoplaneta, por cuanto el protoplaneta permanece fijo en su órbita (el planeta no migra). La densidad superficial de sólidos en la zona de alimentación del planeta solo varía por la acreción por parte del propio planeta, mientras que la densidad de gas queda inalterada (esto último no tiene consecuen- cias relevantes puesto que la cantidad de gas que rodea al planeta es varias veces la masa del mismo). La tasa de crecimiento del núcleo sólido se calcula de acuerdo al modelo de crecimiento oligárquico: dado que el crecimiento runaway del embrión ocurre en una escala de tiempo despreciable respecto de la correspondiente a la formación del planeta (siendo la primera del orden de 104 años, mientras que la segunda es del orden de 107 años), igno- raremos la etapa runaway y comenzaremos las simulaciones con un embrión de masa lunar (aproximadamente del orden de 10−2 M⊕ ). La masa de gas inicial corresponde a la masa de gas ligada gravitatoriamente a ese embrión, que es del orden de 10−10 M⊕ . La tasa de acreción de gas se calcula en forma autoconsistente resolviendo las ecuaciones diferenciales que describen la estructura del interior del planeta, las cuales están acopladas al proceso de acreción de sólidos. La finalización del proceso de formación se estipula en el momento en que el planeta alcanza la masa final, la cual es introducida como parámetro. Este criterio es arbitrario, puesto que no se están considerando los mecanismos reales que regulan la finalización de la acreción de gas, como la disipación del gas nebular o la apertura de una brecha en el disco. De hecho, el estadío final de la formación de un planeta gigante debería ser objeto de un estudio profundo, lo cual está fuera del alcance de esta Tesis. De todos modos, el estudio que presentaremos no se focaliza en esta última etapa, sino en el análisis de la formación hasta una vez iniciado el runaway de gas. Para los resultados que presentaremos en esta sección, las simulaciones se hicieron con- siderando que los planetesimales acretados tienen un único tamaño. Los parámetros que se tuvieron en cuenta para estudiar la respuesta del proceso de formación frente a una variación de los mismos fueron: el radio de la órbita del embrión, la densidad del disco protoplanetario y el radio de los planetesimales acretados. La elección de estos tres pará- metros tiene que ver con que ellos afectan directamente a la tasa de acreción de sólidos (Ec. (4.14) y siguientes) y esto repercute fuertemente en las dos cantidades que se considerarán fundamentales a la hora de evaluar globalmente el proceso que estamos estudiando: la masa del núcleo y la escala de tiempo de formación. 94 Dado que en lo que sigue utilizaremos en repetidas oportunidades el concepto de “masa de aislación”, resulta oportuno aquí mostrar cómo puede calcularse esta cantidad. Recor- demos que la masa de aislación es la masa total de sólidos en la zona de alimentación de un protoplaneta. Σ(a) 2 π ad a. Sea a el semieje orbital del planeta y sea 2 ∆a el ancho de la zona de alimentación. Definimos la masa de aislación, Miso , como la masa total de sólidos presente en la zona de alimentación. Entones tendremos que: Miso = Z a+∆a a−∆a Dado que la densidad superficial de sólidos satisface una ley de potencias con la distancia (Σ(a) ∝ a−p ) la masa de aislación satisface: a+∆a Miso ∝ a2−p (cid:12)(cid:12)(cid:12) a−∆a ∆a = 4 RH ∝ a M 1/3 p Haciendo un desarrollo a primer orden se obtiene para la masa de aislación: Ahora bien, (6.3) (6.1) (6.2) . . Miso ∝ a2−pM 1/3 p Dado que en nuestro modelo todos los planetesimales acretados son incorporados a la masa del núcleo estamos interesados en conocer la masa del núcleo, cuando se alcanza la masa de aislación para una dada masa total del planeta. Como Mp = Mc + Mg tendremos que: (6.4) . c (cid:18)1 + Mc ∝ a2−p M 1/3 1/3 Mg Mc (cid:19) , (6.5) con lo cual, 1/2 Mg Mc (cid:19) (cid:18)1 + Para el caso de un perfil de densidad como el de la nebulosa estándar, p = 3/2 y la masa de aislación resulta ser: Mc ∝ a (6.6) 3 (2−p) 2 . 1/2 1 UA (cid:19)3/4 (cid:18)1 + d (cid:18) a Mg Mc (cid:19) Miso = 0, 65 f 3/2 donde fd representa el aumento en la densidad de sólidos respecto de los valores estándar (fd = 1 si se considera la nebulosa solar estándar). (6.7) M⊕ . 95 6.1.1. Ejemplo y análisis de la formación de un planeta gigante Dado que, frente a la variación de los parámetros que mencionamos antes (radio de los planetesimales, masa del disco, semieje del planeta, etc.), el proceso de formación so- lo cambia cuantitativamente, tomaremos aquí los resultados de una de las simulaciones para hacer el análisis detallado del proceso completo de formación. Los otros casos que presentaremos en esta sección son cualitativamente similares al siguiente ejemplo. Consideremos el caso de un planeta ubicado en la posición actual de Júpiter (a = 5,2 UA), siendo la densidad superficial de sólidos en su zona de alimentación Σ = 10 g cm−2 (correspondiente a, aproximadamente, 3 veces la Nebulosa Solar de Masa Mínima, NSMM), la densidad del gas nebular ρ = 1, 5×10−11 g cm−3 , y la temperatura T = 150 K. El radio de los planetesimales acretados es √10 km (rm = 3, 16 km), siendo la densidad de los mismos 1, 39 g cm−3 (según Pollack et al. 1996). La densidad del núcleo se fijó en 3, 2 g cm−3 . El hecho que la densidad del núcleo se considere mucho mayor a la de los planetesimales acretados tiene que ver con que el núcleo está sometido a la alta presión que ejerce la masa del planeta, lo cual hace que sea mucho más compacto. Denominaremos a este caso que estamos analizando a modo de ejemplo como J 1√10. De acuerdo con los resultados de nuestra simulación, para el caso J 1√10, la formación del planeta se extiende por 7 × 106 años, como se puede observar de la figura 6.1. En alrededor de la mitad de este tiempo el núcleo alcanza una masa de, aproximadamente, 10 M⊕ , mientras que en este proceso, la masa de gas ligada puede considerarse desprecia- ble. Vemos, entonces, que el tiempo que insume la formación del núcleo regula la escala de tiempo que demanda la primera etapa de la formación de un planeta. En el régimen oligárquico, el tiempo característico de crecimiento de un embrión aumenta con su masa (Ec. 4.12), como se muestra en la figura 6.2. En el transcurso de este intervalo de tiempo, para su formación, el núcleo consumió gran parte de los sólidos del disco correspondiente a su región de influencia. Cuando el material en la zona de alimentación comienza a agotarse (último panel de la figura 6.1), la tasa de acreción de sólidos disminuye (tercer panel de la misma figura). En lo que resta para terminar la formación, el núcleo crecerá poco y muy lentamente. En cambio, la tasa de acreción de gas es la que comienza a dominar el proceso. Esto se debe fundamentalmente a dos cosas: por un lado, el embrión ya tiene una masa lo suficientemente grande como para que su influencia gravitatoria sea importante, lo cual le permite mantener ligado mayores cantidades de gas; pero por otro lado, y quizá más importante aún, son las consecuencias del vaciamiento de la zona de alimentación. En el tiempo transcurrido, el embrión utilizó casi todo el material sólido de su vecindad para la formación del núcleo, lo que hace que la acreción de planetesimales disminuya, lo cual, a su vez, significa que el aporte energético que estos cuerpos hacen al ingresar al planeta sea mucho menor. La energía que depositan los planetesimales en las capas de gas de la envoltura aumenta la presión del gas, que tiende a poner resistencia al colapso gravitato- rio. Cuando el número de planetesimales acretados es pequeño, esta energía disminuye y la compresión de las capas de gas aumenta, lo cual da lugar a que la acreción de gas sea cada vez más alta. De hecho, durante la primera parte de la formación, la luminosidad del 96 80 60 40 20 0 0 −10 ] ⊕ M [ a s a M ] s o ñ a / ⊕ M [ ) t d / M d ( g o l Mp Mc Mg ] S L [ ) L ( g o l −4 −6 −8 −10 log(LT) log(LPl) 0 1 2 3 4 5 6 7 Tiempo [106 años] 0 1 2 3 4 5 6 7 Tiempo [106 años] log(dMc/dt) log(dMg/dt) ] 2 m c / g [ Σ 10 5 0 0 1 2 3 4 5 6 7 6 años] Tiempo [10 0 1 2 3 4 5 6 7 6 años] Tiempo [10 Figura 6.1. Los paneles de esta figura muestran la evolución de distintas cantidades en función del tiempo durante el proceso de formación de un planeta gigante, cuyo radio orbital es a = 5,2 UA. El disco protoplanetario es tres veces más masivo que la NSMM, y rm = √10 km. De izquierda a derecha y de arriba hacia abajo, el primer panel muestra la masa total del planeta (Mp ), la masa del núcleo (Mc ) y la masa de la envoltura gaseosa (Mg ) en función del tiempo (Caso J 1√10). Notemos que la mitad del tiempo de formación está destinada a la formación del núcleo sólido el cual, cuando alcanza una masa de alrededor de 10 M⊕ , comienza a acretar gas de una manera más significativa. Después de que la masa del núcleo iguala a la de la envoltura comienza el crecimiento runaway del gas, donde el planeta acreta la mayor parte de su masa en muy poco tiempo. El segundo panel muestra la evolución de la luminosidad en función del tiempo, donde LT representa la luminosidad total y LPl la que proviene de la acreción de los planetesimales. En el tercer panel se muestran la tasa de acreción de sólidos y de gas, log(d Mc/dt) y log(d Mg /dt), respectivamente; y en el cuarto se representa la evolución de la densidad superficial de sólidos, Σ, en la zona de alimentación del planeta. 97 planeta proviene casi exclusivamente de la acreción de planetesimales, como se puede ver del segundo panel de la figura 6.1. Pero cuando los planetesimales ingresantes empiezan a menguar, la luminosidad proveniente de la contracción de las capas de gas comienza a ser del mismo orden que la debida a la de los planetesimales. Luego, el núcleo masivo y el decrecimiento de la energía proveniente de la acreción de los planetesimales, quienes apor- tan la resistencia a la contracción de las capas de la envoltura, le dan paso a un aumento en la acreción de gas. En un intervalo de tiempo comparable al que tomó la formación del núcleo, la masa de la envoltura alcanza un valor crítico: llega a ser igual a la masa del núcleo. Diremos que en este momento se alcanzó la masa de cruce, Mcross , que se define como el valor de la masa del núcleo (o de la envoltura) en el momento en que la masa de la envoltura iguala a la masa del núcleo. El tiempo de cruce corresponde al valor de la coordenada temporal cuando se alcanza la masa de cruce. Matemáticamente, esto es: cuando Mg = Mc (6.8) (6.9) Mcross ≡ Mc tcross ≡ t(Mcross ). Para el caso que estamos analizando, Mcross = 16, 4 M⊕ y tcross = 6, 6 × 106 años. Cuando se supera la masa de cruce, el crecimiento de la envoltura se exacerba y comienza el llamado “runaway de gas”, que lleva a que el planeta adquiera la mayor parte de su masa en un corto intervalo de tiempo. Siendo la masa del planeta cada vez más grande, la autogravedad del planeta no puede ser balanceada por la energía térmica, lo cual provoca el colapso de las capas de gas. Es importante notar que, a pesar del marcado cambio en la tasa de acreción gaseosa, este proceso sigue siendo de carácter hidrostático. Durante el crecimiento runaway del gas, el planeta aumenta su masa muy rápidamente, lo cual hace que la zona de alimentación se expanda (recordemos que la zona de alimentación es proporcional al radio de Hill el cual, a su vez, es proporcional a M 1/3 ). Esto incrementa p notablemente la cantidad de sólidos disponibles, aumentando la tasa de acreción de pla- netesimales y, consecuentemente, la masa del núcleo. Sin embargo, la energía que aportan estos planetesimales no es suficiente para contener el crecimiento de la envoltura. La última etapa de la formación del planeta, que involucra la finalización de la acre- ción, no está incluida en nuestro modelo, ni tampoco ha sido estudiada en profundidad por otros autores en el marco del modelo de inestabilidad nucleada. De acuerdo con Lissauer & Stevenson (2007), de las simulaciones hidrodinámicas con modelos tridimensionales rea- lizadas para el estudio de la interacción entre un planeta que se encuentra acretando gas y el disco protoplanetario, se encuentra que, mientras la masa del planeta es relativamente pequeña (Mp ∼< 10 M⊕), el disco permanece inalterado y puede suministrar gas al planeta sin restricciones. Sin embargo, cuando la masa del planeta aumenta se producen torques en el disco que alejan al gas nebular del planeta, abriéndose una brecha entre ambos. Los límites hidrodinámicos para la tasa de acreción de gas son del orden de 10−2 M⊕/año para planetas de entre 50 y 100 M⊕ . En cuanto a la masa final del núcleo, por otra parte, hay que tener en cuenta que una vez iniciado el runaway de gas, la expansión de la zona de 98 ] s o ñ a 6 0 1 [ o p m e i T 7 6 5 4 3 2 1 0 Mc f(M)= C M1/3 0 5 10 Mc [M⊕] 15 20 Figura 6.2. J 1√10. Esta figura muestra la masa del núcleo (Mc ) en la abscisa y el tiempo en la ordenada. Como se mencionó en § 4.2.2, el tiempo característico del cre- cimiento oligárquico satisface: Tgrow ∝ M 1/3 . En línea roja se muestra el resultado de nuestra simulación, mientras que la línea punteada representa una función proporcio- nal a M 1/3 . La diferencia entre ambas curvas es ínfima hasta que el núcleo alcanza aproximadamente 10 M⊕ . Luego ambas curvas comienzan a separarse debido a que el crecimiento del núcleo se acelera por la presencia de la envoltura. Cuando comienza el runaway de gas (Mg = Mc ≃ 16 M⊕ ), el núcleo aumenta su masa en forma notable porque la zona de alimentación del planeta se expande considerablemente y se puebla, nuevamente, de gran cantidad de planetesimales. 99 alimentación hacia regiones del disco que, en nuestro modelo, conservan los valores de la densidad inicial, sobredimensiona la tasa de acreción de sólidos. Tengamos presente que nuestro modelo no considera la migración de los planetesimales en el disco ni la presencia de otros embriones que también pueden estar acretando material. De hecho, la migración de los planetesimales puede reducir considerablemente la cantidad de material disponible para ser acretado (Thommes, Duncan & Levison 2003). Además, cuando el protoplaneta se vuelve muy masivo provoca también la eyección de parte de los planetesimales hacia afuera de su zona de alimentación. Por estos motivos, de nuestras simulaciones no se pueden obtener los valores finales precisos para la masa del núcleo ni para el tiempo de formación. Resultan sí , mucho más representativas, las magnitudes de la masa y del tiempo de cruce. Tomaremos entonces, en lo que sigue, a ambas variables como referencia a la hora de comparar los resultados de distintas simulaciones. Siguiendo con el análisis de nuestro ejemplo J 1√10, notamos de la figura 6.3 que el radio de acreción define el borde del planeta desde el inicio de la formación y hasta después de comenzado el runaway de gas. Dado que tanto el radio de acreción como el radio de Hill dependen solo de la masa total, y no de la masa de la envoltura, el momento en el cual se produce el cambio entre uno y otro para definir el borde no está relacionado con la etapa del proceso que esté atravesando el planeta. El hecho que el radio de acreción domine al principio de la formación tiene que ver con la ubicación del planeta en el disco y con las condiciones de la nebulosa (que definen la velocidad del sonido para el radio de acreción). Después de que ambos radios se igualan, el radio de Hill resulta ser el menor entre los dos por su dependencia con la raíz cúbica de la masa, lo que hace que el crecimiento del radio de Hill sea mucho más lento que el del radio de acreción, el cual crece linealmente con la masa. Es por esto que, una vez que el radio de Hill satisface la condición para ser el borde del planeta, lo seguirá siendo hasta el final de la formación, siempre y cuando, claro está, no haya eventos de pérdida de masa. En la bibliografía pueden encontrarse autores que definen el borde del planeta, en todo momento, como igual al radio de Hill (por ejemplo, Wuchterl 1995). El argumento para no considerar al radio de acreción en la definición del radio del planeta es que, dado que el planeta está rodeado del gas nebular, por más que las moléculas tengan la energía térmica como para escapar del campo gravitatorio del planeta, la distribución aleatoria de veloci- dades hará que otras moléculas se muevan en la dirección contraria, no habiendo entonces un efecto neto de material que “escape” de sus límites gravitatorios. De todos modos, dado que las capas externas del planeta contienen muy poca masa, en las simulaciones que he- mos hecho no encontramos cambios significativos en los resultados si consideramos, o no, el radio de acreción en la definición del radio del planeta. Hemos adoptado la definición clásica para el borde del planeta porque nos permite comparar más limpiamente nuestros resultados con los de otros autores y porque, como dijimos, la inclusión del radio de acreción no afecta sustancialmente los resultados. Para estudiar lo que ocurre en el interior del planeta durante su formación elegiremos 100 ] m c [ ) R ( g o l 12.5 12 11.5 11 10.5 10 9.5 9 RH Ra Rp 0 10 20 30 40 Mp [M⊕] 50 60 70 80 Figura 6.3. J 1√10. Evolución del radio con la masa del planeta. En línea verde se muestra el radio de Hill, en azul el radio de acreción y en rojo el radio del planeta, el cual se define como: Rp = min[Ra , RH ]. El radio de acreción representa el límite externo durante casi todo el proceso de formación del planeta, salvo cuando éste se vuelve muy masivo y provoca que el borde pase a estar definido por el radio de Hill. Esto ocurre después de comenzado el runaway, cuando la masa de la envoltura domina la masa total del planeta. 101 A Mp= 47 M⊕ 9 10 11 12 13 14 log(R) [cm] C log(LT) log(LPl) ] g [ ) r M ( g o l ] S L [ ) L ( g o l 32 30 28 26 24 22 20 −5 −6 −7 −8 −9 ] g / g r e [ ) l P E ( g o l ] g e s / m c [ ) l P v ( g o l 0.5 0.4 0.3 0.2 0.1 0 6.5 6 5.5 5 4.5 B 9 10 11 12 log(R) [cm] D 9 10 11 12 9 10 11 12 log(R) [cm] log(R) [cm] Figura 6.4. Los cuatro paneles que componen esta figura muestran el perfil radial de la masa (Mr ), la energía que aportan los planetesimales acretados, la luminosidad total y la velocidad de los planetesimales en el interior del modelo correspondiente a Mp = 47 M⊕ , que forma parte del ejemplo J 1√10. La abscisa representa el logaritmo de la coordenada radial. En el panel A vemos la distribución de la masa de la envoltura en función del radio. La línea vertical establece la ubicación del radio del planeta. Para radios mayores, el gas no forma parte del planeta sino de la nebulosa protoplanetaria. Notemos que la masa del planeta se mantiene casi constante desde el borde del planeta hacia el centro en una extensión radial que involucra una variación de R de dos órdenes de magnitud. Entonces, si bien el planeta presenta una estructura gaseosa muy extendida, la mayor parte de la masa se concentra en las capas interiores. El panel B representa el logaritmo de la energía que liberan los planetesimales a medida que atraviesan las capas de la envoltura en su recorrido hacia el núcleo del planeta (la energía liberada cuando los planetesimales llegan a la superficie del núcleo no se muestra). Desde el exterior hacia el centro vemos que la energía tiene dos máximos, el primero para log(R)=10,34 y el segundo sobre la base del núcleo. El primer máximo está relacionado con el aumento de la velocidad (panel D): dado que la velocidad se incrementa, de la Ec. (4.61) vemos que también lo hace la energía que pierden los planetesimales por efecto de la viscosidad. Pero la velocidad luego decrece, lo cual hace que también lo haga la energía. Sin embar- go, en las cercanías del núcleo, la velocidad se mantiene casi constante pero la densidad sigue en aumento y tomando los valores más altos, con lo cual la energía liberada por los planetesimales se incrementa nuevamente. El panel C muestra el logaritmo de la lu- minosidad total, log(LT ), y la luminosidad proveniente de la acreción de planetesimales, log(LPl ), ambas en unidades de la luminosidad solar. Notar que son las capas externas las que contribuyen al aumento de la luminosidad respecto de la debida a la acreción de los planetesimales. Por último, el panel D muestra los cambios en la evolución de la 102 velocidad de los planetesimales, gobernada por la Ec. (4.56). uno de los modelos de la secuencia de esta simulación para hacer el análisis (llamamos modelo al conjunto de variables que describen, para un determinado instante t, el estado de la envoltura del planeta, desde el borde externo hasta la base del núcleo sólido). Tomamos el modelo que corresponde a una masa total de 47 M⊕ (t ≃ 6, 8 × 106 años), donde la masa de la envoltura es, aproximadamente, 30 M⊕ , el radio del planeta es log(Rp ) = 12, 4, en centímetros, y el radio del núcleo es log(Rp ) = 9, 3. En la figura 6.4 se grafican cuatro paneles que muestran la masa, la luminosidad, la energía depositada por los planetesimales y la velocidad de los planetesimales en función del radio (en realidad, en todos los casos se consideró el logaritmo de estas cantidades). Se eligió el radio como referencia porque, dado que el planeta presenta una estructura muy extendida, esta coordenada permite apreciar en detalle las variaciones de las cantidades que se quieren estudiar. Del primer panel podemos ver que prácticamente la totalidad de la masa está contenida en log(R) < 10, 4. Entre este valor y el borde del planeta la densidad del gas es muy ba ja y el aporte que hacen estas capas a la masa total es inferior al 10 %. Sin embargo, en parte de esas capas (10, 5 < log(R) < 11) está el aporte más importante a la luminosidad debido a la contracción del gas. En el panel C esto se ve reflejado en la diferencia entre la luminosidad total y la correspondiente a la acreción de los planetesimales. La energía cinética que los planetesimales pierden en forma de calor por el efecto de la viscosidad del gas es: (6.10) dE dr ∝ ρg v 2 , como lo establece la ecuación (4.61). Luego, la energía que absorba cada capa de gas depen- derá de la densidad ρg y de la velocidad v . Cuando los planetesimales ingresan al planeta lo hacen con velocidad vrel (Ec. (4.16)). La densidad del gas en las capas exteriores del planeta es ba ja, y la ecuación de movimiento (Ec. 4.56) está dominada por la aceleración gravitatoria, lo que hace que la velocidad de los planetesimales aumente (panel D). Como la energía que pierden los planetesimales depende de la velocidad de los mismos, si ésta aumenta, también lo hace la energía que ceden a la envoltura. Cuando los planetesimales ingresan en regiones más densas, la fricción es mayor. Pero, además, la fracción de masa del planeta que atraviesan disminuye, por lo tanto la aceleración gravitatoria es menor. En- tonces, la viscosidad comienza a jugar un papel importante en la ecuación de movimiento, con lo cual la velocidad decrece y la energía liberada también. Sin embargo, llegando a la base de la envoltura, la energía comienza a aumentar nuevamente. Esto ocurre cuando el frenado producido por la viscosidad compensa a la aceleración gravitatoria y la velocidad se vuelve casi constante. La energía liberada aumenta ahora por su dependencia con la densidad, ya que el planetesimal atraviesa las regiones donde el gas es más denso. Para estudiar la evolución de las variables termodinámicas en el interior del planeta, en la figura 6.5 se muestran los perfiles de temperatura, presión, luminosidad y densidad para cinco modelos (Mp = 15, 32, 47, 57 y 65 M⊕), donde el de menor masa está dominado por la masa del núcleo, el segundo corresponde al momento de cruce (cuando Mg = Mc ), y los tres restantes se encuentran en la etapa de crecimiento runaway del gas. Mientras que la temperatura, la presión y la densidad son monótonamente decrecientes con el ra- dio, la luminosidad es creciente. Los valores de las tres primeras en el borde del planeta 103 corresponden a las condiciones de borde que caracterizan el estado de la nebulosa solar. A medida que la masa de la envoltura aumenta, cada una de estas variables también au- menta a radio fijo. Notemos, sin embargo, que un valor de radio constante no representa al mismo elemento de masa en dos modelos distintos. Las curvas que muestran el perfil de temperatura presentan una región achatada (en el caso, por ejemplo, de Mp = 15 M⊕ corresponde a 10 < log(Rc) < 10, 5) que se corresponde con la aparición de la primera zona de transporte radiativo en el interior (figura 6.6). El gradiente de temperatura (4.26) depende del tipo de transporte que tenga lugar en al región, si el transporte es radiativo el gradiente será proporcional a ∇rad (Ec. 4.28), mientras que si es convectivo se considera proporcional al gradiente adiabático (∇ad ). El tipo de transporte de energía que domina la estructura del planeta depende de la masa que tenga la envoltura. En todos los casos, la región más interna es convectiva. Luego aparece la región radiativa que recién mencionába- mos. Siguiendo hacia afuera se desarrollan pequeñas zonas convectivas que crecen a medida que la masa del planeta crece. En el caso del modelo de menor masa considerado aquí , el borde externo es radiativo mientras que cuando la masa de la envoltura se encuentra en pleno crecimiento runaway, la región exterior se vuelve completamente convectiva. Ésto tiene que ver con que el gradiente de temperatura se vuelve cada vez más agudo a medida que la masa del planeta aumenta, superando al gradiente adiabático, lo cual hace que se instale el transporte convectivo. En cuanto a la luminosidad, ésta aumenta con la masa. La luminosidad impuesta por la condición de borde en la superficie del núcleo es que ésta sea cero. Sin embargo, la luminosidad en la primera capa de gas no es cero debido a la energía que dejan los planetesimales. En el caso de los modelos menos masivos, como el correspon- diente a Mp = 15 M⊕ , la luminosidad proviene casi exclusivamente de la acreción de los planetesimales. A medida que la masa de la envoltura aumenta, comienza a aparecer un punto de quiebre en el perfil de la luminosidad. Éste está relacionado con la contracción de las capas externas que, como mencionamos en un párrafo anterior, frente al decrecimiento de la acreción de sólidos y al aumento de la masa del planeta, incrementan el valor de la luminosidad. Por último, la figura 6.7 muestra la evolución de los valores de la densidad y la tempe- ratura de la última capa de gas (que se encuentra sobre la superficie del núcleo) durante toda la formación del planeta. Si bien la figura se extiende sobre diversos valores de la tem- peratura y la densidad, es importante destacar que la zona verdaderamente representativa para el análisis del estado del gas en la adyacencia con el núcleo del planeta, comienza para log(T ) > 3, 8 (Mg > 0, 01 M⊕ ). Para valores “centrales” menores, la masa de gas ligado es despreciable. Mencionaremos, sin embargo, que en la región donde 3, 1 < log(T ) < 3, 4, la opacidad de los granos disminuye abruptamente, lo que hace que en esa zona del interior del planeta el transporte sea siempre radiativo. Por eso es que aparece, en esa región, el cambio de pendiente en la curva. Cuando los valores centrales se encuentran en este rango de temperaturas, se libera, relativamente, mucha energía y las capas se contraen conside- rablemente, hasta que las condiciones en región más interior del planeta salen de la zona de ba ja opacidad. Sin embargo, dado que cuando estas son las condiciones presentes en el planeta la masa de la envoltura es ínfima, este hecho no tiene una repercusión impor- 104 tante en la formación del planeta. En la figura se incluyó también el diagrama de fases del hidrógeno, correspondiente a la EOS de SCVH, y la curva que representa el estado del interior del modelo de 47 M⊕ . Dada la monotonía de la temperatura y la densidad, los valores “centrales” son los más altos que se registran para cada modelo. Vemos entonces que, como la curva no atraviesa la PPT, ningún modelo de los que componen la secuencia de esta simulación atravesará esta discontinuidad. Sin embargo, el interior del planeta sí atravesará dos transiciones: del hidrógeno molecular al atómico, y del atómico al ionizado, como puede observarse en la curva de 47 M⊕ . 6.1.2. Variación de los resultados frente al cambio de las condi- ciones de borde externas Cualitativamente, el proceso de formación de un planeta gigante, en base a nuestras simulaciones, puede ser descripto como lo hicimos en la sección previa, sin verse notoria- mente afectado por variaciones en las condiciones de borde impuestas por la nebulosa, por cambios en su ubicación en el disco protoplanetario o por el tamaño de los planetesimales acretados. Sin embargo, tanto la masa como el tiempo de cruce, son muy sensibles a la elección de estos valores. Consideremos primero el caso de un planeta ubicado en la posición actual de Júpiter (a = 5, 2 UA). En la tabla 6.1 se muestran los resultados de nuestras simulaciones cuando la densidad del disco protoplanetario corresponde al rango de 6 a 10 veces NSMM. Para cada uno de estos cinco casos se consideraron dos posibles valores para el radio de los Tabla 6.1. Masa y tiempo de cruce para diez simulaciones en las cuales se variaron la densidad del disco en la ubicación del planeta (a = 5, 2 UA), y el radio de los plane- tesimales acretados (rm = 10, 100 km). En la primera columna se muestra la densidad del disco protoplanetario en función del número de veces que representa respecto de la nebulosa estándar. La segunda y cuarta columna listan los resultados para el tiempo de cruce, en millones de años, y en la tercera y quinta columna se muestran las masas de cruce en unidades de masas de la Tierra. Masa del disco [MMSN] 6 7 8 9 10 Radio de los planetesimales rm = 100 km Tiempo de cruce [My] 7, 20 5, 58 4, 60 3, 78 3, 20 Masa de cruce [M⊕ ] 25, 50 27, 90 29, 30 30, 90 31, 90 Radio de los planetesimales rm = 10 km Tiempo de cruce [My] 2,43 1,97 1,49 1,32 1,10 Masa de cruce [M⊕ ] 31, 80 34, 80 36, 80 38, 60 39, 90 105 ] K [ ) T ( g o l ] S L [ ) L ( g o l 4.5 4 3.5 3 2.5 2 −4 −5 −6 −7 −8 −9 −10 9 10 11 12 log(R) [cm] 15 M⊕ 32 M⊕ 47 M⊕ 57 M⊕ 65 M⊕ 9 10 11 12 log(R) [cm] ] 3 m c / g r e [ ) P ( g o l ] 3 m c / g [ ) ρ ( g o l 12 10 8 6 4 2 0 −2 0 −2 −4 −6 −8 −10 −12 9 10 11 12 9 10 11 12 log(R) [cm] log(R) [cm] Figura 6.5. J 1√10. Perfiles de distintas variables termodinámicas en el interior del planeta en formación cuando su masa total es: 15 M⊕ , 32 M⊕ , 47 M⊕ , 57 M⊕ , y 65 M⊕ . La abscisa representa el logaritmo de la coordenada radial. El primer panel muestra el perfil de temperatura, el segundo el de la presión, el tercero la luminosidad (en unidades solares) y el cuarto el de la densidad. El límite exterior del planeta se encuentra, según el orden creciente en la masa, en: 11,9, 12,2, 12,4, 12,45, 12,5; en todos los casos este valor corresponde al log(Rp ). 106 15 M⊕ 32 M⊕ 47 M⊕ 57 M⊕ 65 M⊕ e t r o p s n a r T 2 1.5 1 0.5 0 9 9.5 10 10.5 11 11.5 12 12.5 log(R) [cm] Figura 6.6. J 1√10. Para los mismos casos que el la figura 6.5, mostramos aquí el tipo de transporte que se establece en las diferentes capas del interior del planeta. Los valores entre 0 y 0,5 corresponden a regiones de transporte radiativo, mientras que entre 1 y 1,5 al transporte convectivo (los distintos casos fueron espaciados en 0,1 unidades para su mejor visualización). El transporte radiativo domina la estructura en los estadíos de menor masa. Cuando la masa de gas ligada supera a la masa del núcleo, la convección se hace más importante, tanto en el interior profundo como en las capas externas. 107 ] 3 m c / g [ ) ρ ( g o l 0 −2 −4 −6 −8 −10 −12 PPT H+ H2 H H+ 47 M⊕ 2 2.5 3 3.5 4 4.5 5 log(T) [K] Figura 6.7. J 1√10. El eje x corresponde al logaritmo de la temperatura y el eje y al logaritmo de la densidad. La curva roja muestra la evolución de estas dos cantidades en la base de la envoltura, durante toda la formación del planeta. A medida que el planeta incrementa su masa, las capas de gas del interior se contraen, aumentando su densidad y su temperatura. La curva en línea de puntos azules representa el interior del modelo de 47 M⊕ . El valor mínimo responde a las condiciones nebulares, mientras que el máximo corresponde a uno de los puntos de la curva roja. En este gráfico se añadió el diagrama de fases del hidrógeno de la EOS de SCVH. En línea punteada gris se muestra la ubicación teórica de la Plasma Phase Transition (PPT), y las regiones de hidrógeno molecular, atómico e ionizado. 108 planetesimales acretados: 10 y 100 km. Como ya mencionamos, tomaremos como valores representativos para cada una de las simulaciones los resultados de Mcross y tcross . Los valores de la densidad del disco fueron seleccionados de modo que el tiempo de cruce, en todos los casos, fuera inferior a 107 años (este valor en general es el que se toma como característico para la vida media de los discos, ver Haisch 2001). Notemos que considerar la densidad del disco como la correspondiente a n veces la NSMM no necesariamente significa que la masa total del disco sea n veces la masa estándar de la nebulosa solar. Dado que nos estamos focalizando en una determinada órbita, la cual es además fija, los valores de la densidad altos pueden deberse a enriquecimientos locales (Dodson-Robinson 2008a) y no necesariamente extenderse uniformemente en todo el disco. Entonces, el asociar la densidad para una dada posición a un valor de la NSMM es solo a modo de referencia y no implica que estemos considerando un disco que se corresponda con esa masa. En cuanto al tamaño de los planetesimales, ambos valores son los que frecuentemente se encuentran en la literatura cuando se estudia este tipo de problemas. Hay dos características fundamentales que surgen de analizar los datos a radio fijo: cuando aumenta la densidad del disco, el tiempo de cruce disminuye y la masa de cruce aumenta. De la ecuación (4.14), se ve que la tasa de acreción es proporcional a la densidad del disco, por lo que alcanzar una determinada masa para el núcleo será más rápido cuanto más sólidos haya disponibles para ser acretados. El hecho que la masa de cruce no sea en todos los casos la misma sino que aumente con la densidad tiene que ver con que, a mayor acreción, mayor es también la energía que suministran los planetesimales a la envoltura, lo cual retarda el inicio del colapso de las capas de gas. Sin embargo, la reducción porcentual del tiempo de cruce es mucho mayor que el aumento en la masa de cruce. Tanto para el caso de rm = 100 km, como para rm = 10 km, si comparamos los dos casos de los extremos (nos referimos a las simulaciones correspondientes a 6 y 10 NMMS), al aumentar la densidad el tiempo de cruce se reduce en un factor 2,25 mientras que la masa de cruce solo aumenta un 25 %. Un resultado interesante es que, si nos restringimos al rango de densidades que estamos considerando, y pensamos a la masa de cruce como función del tiempo de cruce, se puede a justar una recta entre los puntos para cada uno de los casos considerados (ver figura 6.8). Esto es, si: (6.11) Mcross = a tcross + b, obtenemos que para planetesimales de radio rm = 100 km, a = -1,6 y b = 36,9; mientras que para rm = 10 km, a = -5,95 y b = 46,25. Es importante mencionar que este ajuste no puede extrapolarse para densidades mucho mayores o menores a las consideradas para estos casos. Claramente, para densidades suficientemente grandes nos encontraríamos con tiempo negativos, y para densidades suficientemente pequeñas con masas de cruce también negativas. Sin embargo, por otra parte, el rango de densidades donde vale el a juste es bastante amplio. Discos con densidades superiores a 10 NSMM serían muy masivos y, pro- bablemente, poco frecuentes. Por otra parte, para densidades ba jas el tiempo de formación sería demasiado largo. Si ahora comparamos los resultados para una misma densidad del disco pero para los diferentes radios de los planetesimales considerados encontramos que para los planetesi- 109 ] ⊕ M [ s s o r c M 32 30 28 26 24 rm= 100 km ] ⊕ M [ s s o r c M 40 38 36 34 32 30 rm= 10 km 3 4 6 5 tcross [106 años] 7 8 1 1.5 2 tcross [106 años] 2.5 Figura 6.8. Ajuste de mínimos cuadrados para los datos de la tabla 6.1. El primer panel corresponde a los casos donde el radio de los planetesimales acretados es de 100 km, y el segundo cuando el radio de 10 km. En los dos casos se ajustó una recta (línea a rayas) entre los valores de la masa de cruce y el tiempo de cruce (cruces rojas) para las condiciones del disco consideradas. males más pequeños el tiempo de cruce se reduce notablemente pero la masa de cruce aumenta. En este caso, lo que influye en la tasa de crecimiento del núcleo (Ec. (4.14)) para que esto sea así es el radio efectivo de captura, que depende de la velocidad rela- tiva entre los planetesimales y el embrión. De la ecuación (4.16), se ve que vrel depende de la excentricidad, la cual depende a su vez de la masa (y por ende del radio) de los planetesimales (ver Ec. (4.19)). Luego, la ecuación (4.14) depende, aproximadamente, del inverso de la velocidad relativa (esto se ve de la Ec. (4.15), que define el radio efectivo en ausencia de gas). Los planetesimales más pequeños tienen una velocidad relativa más ba ja puesto que se ven más afectados por la viscosidad nebular, lo cual implica un aumento del radio efectivo de captura del planeta. Esto lleva a que la tasa de acreción aumente, lo que conlleva una reducción en las escalas de tiempo. Al igual que en el caso anterior, el hecho de incorporar más planetesimales por unidad de tiempo implica que la energía que liberan en la envoltura produce que el colapso de las capas de gas ocurra para núcleos más masivos. La reducción en el tiempo de formación al considerar planetesimales de menor radio es inclusive más marcada que cuando se considera un aumento en la densidad nebular para planetesimales de 100 km. Para una misma densidad nebular el tiempo de formación disminuye en un factor 3 si comparamos los resultados para planetesimales de 10 y 100 km. En cuanto al aumento en la masa de cruce para esta misma comparación encontramos que es de un 25 %. Esto implica que si los planetesimales presentes en la nebulosa son en su mayoría de radio menor a los 100 km (o, incluso, a los 10 km), se podrían considerar densidades más ba jas para el disco, lo cual llevaría a una reducción en la masa del núcleo, con tiempos de formación todavía dentro de los límites impuestos por las observaciones. Consideremos ahora dos posibles valores para la densidad del disco (3 NSMM y 7 NSMM), dos radios para los planetesimales acretados (rm = 10, 100 km) y cinco semiejes (9,55, 8, 6, 5,2, y 4 UA), entre los cuales se incluyen los correspondientes a Júpiter y 110 Saturno. En la tabla 6.2 se listan los resultados para la masa y el tiempo de cruce para los casos donde el tiempo de cruce sea inferior a 20 millones de años. Como era esperable, el tiempo de cruce disminuye cuando nos acercamos al Sol, puesto que la densidad aumenta en esa dirección. Notemos que ninguno de los casos considerados se ve afectado por la discontinuidad en la línea de hielo. Similarmente, la masa de cruce aumenta en dirección al Sol. Resultados análogos a los anteriores se repiten al aumentar la densidad del disco y al considerar que el tamaño de los planetesimales se reduce. En todos los casos se puede ver el amplio espectro de los tiempos de cruce, el cual se ve seriamente afectado por la distancia al Sol (y siempre teniendo en cuenta la restricción establecida por la vida media de los discos). Como se ve de estos resultados, parece difícil la formación de todos los planetas gigantes del Sistema Solar ba jo las condiciones de una nebulosa solar estándar. Tengamos en cuenta que en estas simulaciones no se consideraron los semiejes correspondientes a Urano y Neptuno, mucho más alejados del Sol que Saturno. A pesar de eso, es evidente que el tiempo de formación resultaría extremadamente largo, aún para discos muy masivos. Notas relativas a la ecuación de estado y su impacto en los resultados En nuestro traba jo, Fortier, Benvenuto & Brunini (2007), presentamos gran parte de los resultados mencionados arriba. Sin embargo, meses después de su publicación, descubrimos errores de tipeo en el artículo original de Saumon, Chabrier & Van Horn (1995). Estos errores involucran a las fórmulas de interpolación de la entropía de mezcla, necesarias para calcular la ecuación de estado de un gas compuesto por hidrógeno y helio. En § 4.4.2 mencionábamos cuáles eran los errores en las ecuaciones. Aquí presentaremos los resultados obtenidos con la EOS sin corregir y los compararemos con la EOS corregida (Fortier, Benvenuto & Brunini 2009). Es interesante analizar el impacto que estos errores tienen en los resultados y sirven para poner de manifiesto la sensibilidad del proceso de formación de un planeta gigante a variaciones en la EOS. Tabla 6.2. Masa y tiempo de cruce variando la ubicación del planeta, la densidad del disco y el radio de los planetesimales acretados. La masa de cruce, Mcross , está en unidades de masas de la Tierra (M⊕ ) y el tiempo de cruce, tcross , en millones de años (My). a [UA] rm = 100 km tcross Mcross 3 NSMM rm = 10 km tcross Mcross rm = 100 km tcross Mcross 7 NSMM rm = 10 km tcross Mcross 9, 55 8, 00 6, 00 5, 20 4, 00 > 20 > 20 > 20 > 20 18, 2 − − − − 13, 0 > 20 > 20 12, 3 10, 3 9, 2 > 20 16, 1 8, 3 5, 6 3, 3 − 24, 5 27, 0 27, 9 29, 9 8, 7 5, 6 3, 2 2, 0 1, 0 29, 7 31, 0 33, 3 34, 8 35, 5 − − 17, 0 15, 9 13, 5 111 En la tabla 6.3 mostramos los resultados correspondientes a la EOS sin corregir (tiempo y masa de cruce) para densidades del disco entre 6 y 10 NSMM, y para planetesimales con radios de 100 y 10 km. De comparar estos resultados con los presentados en la tabla 6.1 vemos que, con la EOS sin corregir los tiempos de cruce son más largos pero, fundamen- talmente, las masas de cruce son considerablemente mayores, en la mayoría de los casos más del doble que cuando se considera la EOS corregida. Los errores en la EOS afectan a la entropía de la mezcla (Ec. 4.39), pero fundamental- mente a las derivadas logarítmicas de la misma, ST y SP (Ecs. 4.44, 4.45). Estos valores son muy importantes puesto que su cociente define el gradiente adiabático (Ec. 4.46), el cual, a su vez, es fundamental para determinar el tipo de transporte. Cuando se comparan los valores del gradiente adiabático entre ambas tablas de la EOS, se ve que en la tabla corregida, ∇ad es siempre menor que en la tabla sin corregir, y ba jo las condiciones pre- sentes en el interior del protoplaneta en algunas regiones la diferencia puede llegar hasta un 50 %. Esto implica que nuestros primeros resultados fueron obtenidos sobreestimando los valores del gradiente adiabático. De acuerdo con el criterio de Schwarzschild (4.4.1), cuando ∇ > ∇ad el transporte se vuelve convectivo. Cuanto mayor sea ∇ad , más difícil se vuelve el desarrollo de regiones convectivas que permitan un transporte más eficiente de la energía hacia el exterior. Esto frena la gran contracción de las capas de gas, lo cual demora el comienzo del crecimiento runaway. Se tiene como consecuencia la prolongación del tiempo para alcanzar la masa de cruce que, en la mayoría de los casos considerados resulta ser cercana a la masa de aislación (esto último es mucho más notorio en el caso de los planetesimales de 10 km que en los de 100 km). La masa de aislación de un protoplaneta es la masa total de sólidos presente en su zona de alimentación. En las simulaciones que que estamos mostrando a modo de ejemplo, la masa de cruce (con la EOS sin corregir) resulta ser muy cercana a la masa de aislación, lo cual implica que la contracción es significativa recién cuando el planeta está cercano a agotar su fuente de planetesimales. Esto quiere decir que el desplome de las capas de gas recién ocurre cuando deja de ingresar la energía suficiente para impedir el colapso, sin que tenga mucha incidencia en esto la capacidad del gas para transportar hacia afuera la energía aportada por los planetesimales. El error en la fórmula de interpolación solo modificó el valor del gradiente adiabático dado por la EOS (también afectó, pero en forma despreciable, al calor específico a presión constante). En las regiones de interés para el planeta, la diferencia entre ambos valores puede tener máximos de hasta un 50 %, pero en promedio queda acotada a alrededor de un 15 %. Esta diferencia, sin embargo, repercute significativamente en los valores de la masa y el tiempo de cruce. 112 Tabla 6.3. Masa de cruce y tiempo de cruce para los mismos casos presentados 6.1, pero con la EOS sin corregir. Masa del disco [MMSN] Resultados con la EOS sin corregir (rm = 100 km) Masa de Tiempo de cruce [My] cruce [M⊕ ] Resultados con la EOS sin corregir (rm = 10 km) Masa de Tiempo de cruce [My] cruce [M⊕ ] Masa de ais- lación [M⊕ ] 6 7 8 9 10 11, 40 8, 75 7, 00 5, 70 4, 80 44, 70 55, 40 66, 60 77, 60 88, 50 4, 95 3, 53 2, 65 2, 00 1, 65 46, 70 58, 70 71, 40 84, 85 99, 00 46, 52 58, 60 71, 60 85, 00 100, 00 6.2. Simulaciones considerando una población de pla- netesimales que sigue una distribución de tama- ños Como vimos en la sección anterior, al aumentar la densidad en el disco se reduce el tiempo de formación pero también aumenta la masa del núcleo. El mismo efecto se obtiene manteniendo fija la densidad pero disminuyendo el tamaño de los planetesimales acretados. Considerar densidades muy altas implica invocar mecanismos que permitan que ese enri- quecimiento ocurra. Sin embargo, considerar planetesimales pequeños es, por el contrario, una hipótesis que se condice completamente con la realidad. De hecho, surge naturalmente pensar que los planetesimales que pueblan el disco no tienen una masa uniforme sino que siguen alguna ley de distribución de masas. Por ejemplo, para el caso de un planeta en la ubicación de Júpiter, se suele aceptar que un valor plausible para la densidad del disco sea Σ = 10 g cm−2 (aproximadamente el correspondiente a 3 NSMM). La figura 6.9 muestra el proceso de formación del planeta considerando que los planetesimales acretados tienen un radio uniforme de 10 y 100 km respectivamente. Mientras que en el caso en el que los planetesimales tienen un radio de 100 km el tiempo de formación es de alrededor de 24 millones de años, superando ampliamente la cota impuesta por las observaciones, en el caso de planetesimales de 10 km el tiempo de formación es de 10 millones de años. En pocas palabras, el radio que se considere para los planetesimales acretados hace que, ba jo las mismas condiciones nebulares, los resultados de una simulación puedan ser categorizados como “aceptables” o no en función del tiempo que demande su formación. Kokubo & Ida (2000) estudiaron mediante simulaciones de N -cuerpos tridimensionales la ley de distribución de masas entre los planetesimales del disco protoplanetario. Sea nc el 113 Mp Mc Mg rp= 10 km rp= 100 km ] ⊕ M [ a s a M 80 70 60 50 40 30 20 10 0 0 5 10 15 20 25 6 Tiempo [10 años] Figura 6.9. Evolución de la masa en función del tiempo para un planeta ubicado en a = 5, 2 UA, bajo las condiciones de un disco protoplanetario de 3 NSMM. En línea roja se muestra la masa total, en azul la masa del núcleo y en fucsia la de la envoltura. Se consideraron dos casos: uno para planetesimales acretados de radio 10 km, y otro para 100 km. Notar que el tiempo que involucra este proceso en cada uno de los casos hace que, cuando rm = 100 km se superen los 107 años impuestos por la vida media de los discos. El caso de planetesimales de 10 km queda dentro de esta cota. 114 número acumulativo de planetesimales de masa m,1 y sea n(m) el número de planetesimales con masa en el intervalo (m, m + dm). Kokubo & Ida encontraron que la distribución de masa entre los planetesimales resulta bien aproximada por una ley de potencias, d nc dm ∝ n ∝ mα . (6.12) Dado que la masa total es: (6.13) Z n m dm ∝ mα+2 , si α < −2, los planetesimales más pequeños contendrán la mayor parte de la masa del sistema. Que α sea menor a -2 caracteriza al crecimiento runaway y post-runaway (creci- miento oligárquico), donde los planetesimales dominan la masa total del sistema mientras que los embriones son los únicos que continúan en crecimiento. En base a sus simulaciones, Kokubo & Ida encontraron que α decrece con el tiempo, y que toma valores ente -3 y -2, siendo el valor de referencia α = −2, 5. En función de los resultados previos, donde la masa y el tiempo de cruce son significati- vamente dependientes del tamaño de los planetesimales acretados, un estudio más realista del problema lleva, entonces, a la incorporación de una distribución de tamaños para los planetesimales. Dado que en nuestro modelo la densidad de los planetesimales es constante, independientemente de su radio, transformaremos la ley de potencias para la masa en una ley de potencias para el radio. Como m ∝ r3 , tenemos: n(m) m dm ∝ r3 α+5d r, que representa la masa contenida en planetesimales cuyo radio se encuentra en el intervalo (r, r + d r). En la práctica, un intervalo de tamaños que nos sea de interés tendrá que ser discreti- zado. Sea rmax y rmin los dos extremos de la distribución que estamos considerando y sea η el número de especies en las cuales estará discretizada la distribución. Tomaremos a los valores del radio equiespaciados en escala logarítmica: rj = rmin (cid:18) rmax rmin (cid:19) donde, claramente, r1 = rmin y rη = rmax . Sea f (r) = C rβ la ley de potencias para el radio, donde f (r) d r es el número de planetesimales con radio en (r, r + d r). Para satisfacer la ecuación (6.14), β = 3 α + 5. La masa total contenida en el intervalo (rmin , rmax ) será: 2(η−1)  − 1 β+1 rβ+1 β + 1 (cid:18) rη (cid:18) rη Z rη+1/2 r1 (cid:19)− r1 (cid:19) 1  . r1−1/2 1nc representa el número total de cuerpos de masa mayor a m. f (r)d r = C (6.14) j = 1, ..., η (6.15) η(β+1) η−1 (6.16) j−1 η−1 115 1 η+1 , Notemos que los límites de la integral no son r1 y rη , sino que se toman por fuera del intervalo. Ésto es así porque se está considerando que los planetesimales con radio en (rj−1/2 , rj+1/2) están representados por una especie de radio rj . Luego, si definimos ∆ = (cid:18) rmax rmin (cid:19) la fracción F , respecto del total, de planetesimales con radios en (rj−1/2 , rj+1/2) es: ∆(j−1)(β+1) [∆β+1 − 1] ∆η(β+1) − 1 En el tratamiento numérico, cada bin de tamaño se traba ja en forma independiente. Se calcula el radio efectivo para cada una de las especies de planetesimales, se establece la tasa de acreción de sólidos en cada caso y se elimina la masa de planetesimales acretada por cada uno de los bines. En cada capa de la envoltura se deposita la energía total que dejan los planetesimales, que es la suma de la energía que pierde por viscosidad cada una de las especies. Hechos estos cálculos, se procede a la resolución del modelo de la manera usual. F = (6.17) . (6.18) 6.2.1. El modelo de Niza Las teorías de formación del Sistema Solar indican que los planetas gigantes se formaron en órbitas circulares y coplanares. Sin embargo, esa no es, estrictamente hablando, la configuración que presentan actualmente los planetas gigantes del Sistema Solar (ver tabla 1.1). Tsiganis et al. (2005) mostraron que un sistema planetario donde las órbitas iniciales sean coplanares y cuasi circulares puede evolucionar hasta el estado actual, siempre y cuando, Júpiter y Saturno crucen la resonancia de movimientos medios2 1:2. Una vez que los planetas gigantes culminaron su formación y se disipó la nebulosa solar, los planetas migraron debido al intercambio de momento angular con el disco de plane- tesimales. En este proceso, Júpiter debió migrar hacia el interior mientras que Saturno, Urano y Neptuno lo hicieron en el sentido contrario. Durante la migración, las excentrici- dades e inclinaciones de los planetas se amortiguaron debido a la fricción dinámica con las partículas del disco. Si las órbitas originales de los planetas estaban lo suficientemente pró- ximas entre sí , es probable que hayan tenido que atravesar una resonancia de movimientos medios de ba jo orden, lo cual excitaría las excentricidades de los planetas involucrados. Tsiganis et al. (2005) realizaron simulaciones numéricas considerando los siguientes semi- ejes iniciales: aJ = 5, 45 UA, aS = 8, 65 UA (en el interior de la resonancia 1:2 con Júpiter), y aN , aU ∈ [11, 17] UA (con aJ , aS , aN y aU los semiejes de Júpiter, Saturno, Neptuno y Urano respectivamente), y la presencia de un disco de planetesimales (de entre 30 y 50 2Dos planetas se encuentran en resonancia de movimientos medios si existe una relación de conmen- surabilidad entre sus períodos orbitales. 116 M⊕ ), que se extiende hasta 30-35 UA. Tsiganis et al. encontraron que Júpiter y Saturno cruzan la resonancia 1:2 durante su migración, lo cual excita sus excentricidades hasta valores compatibles con los que se miden actualmente. Este cambio en las excentricidades tiene un efecto drástico en todo el sistema planetario, ya que fuerza el incremento de las excentricidades de Urano y Neptuno. Las órbitas de los planetas se intersectan, lo que provoca cambios en las inclinaciones y que los gigantes de hielo migren hacia el exterior, adentrándose en el disco. De hecho, en la mitad de sus simulaciones, Tsiganis et al. ob- tienen que Urano y Neptuno intercambian posiciones. Que Neptuno se haya formado en una órbita interior a la de Urano explicaría la mayor abundancia de sólidos en su inte- rior. Luego, la fricción dinámica estabiliza al sistema y la migración termina cuando los planetesimales son eyectados del disco. La configuración final del sistema, en la mayoría de sus simulaciones, muestra un excelente acuerdo con los valores actuales. Este modelo se conoce como modelo de Niza. El modelo de Niza consigue explicar las características orbitales más importantes de los planetas gigantes del Sistema Solar; esto es, los semiejes, excentricidades e inclinaciones. Además, es consistente con la existencia de los satélites regulares, de los Troyanos de Júpiter (y, probablemente también, con los de Neptuno), y no contradice la distribución que presenta el Cinturón Principal de Asteroides. 6.2.2. Resultados En general, las simulaciones de formación de planetas gigantes dentro del marco de la inestabilidad nucleada, pueden dar cuenta de Júpiter y Saturno en tiempos razonablemente cortos (Pollack et al.1996, Hubickyj, Bodenheimer & Lissauer 2005, Dodson-Robinson et al. 2008b). Sin embargo, tanto las estimaciones analíticas como las numéricas fracasan a la hora de explicar la formación de Urano y Neptuno en escalas de tiempo acordes a la vida media de la nebulosa (Lissauer 1987, Thommes, Duncan & Levison 2003, Hillenbrand 2005, Lissauer & Stevenson 2007). La clave está en que, en las ubicaciones actuales de estos planetas, la densidad superficial de sólidos de la nebulosa primordial tendría que haber sido muy ba ja, lo cual extendería el proceso de acreción. De hecho, que la masa de gas de Urano y Neptuno sea bastante inferior a la masa de su núcleo podría tomarse como un indicador de que estos planetas no hayan alcanzado nunca el runaway de gas, probablemente porque la nebulosa se habría disipado antes de que esto ocurriera. Sin embargo, de acuerdo con el modelo de Niza, Urano y Neptuno se formaron mucho más cerca del Sol de lo que se encuentran actualmente. De haber sido esto así , estos planetas habrían tenido mayor cantidad material disponible, lo cual habría facilitado la acreción, acelerando el proceso. Es importante notar que, si la configuración inicial del Sistema Solar exterior era más compacta que la actual, la nebulosa solar estándar, por construcción, resulta inconsistente con este modelo. Basándose en argumentos similares a los de Weidenschilling (1977) y Hayashi (1981), pero considerando una arquitectura primordial compatible con el modelo de Niza, Desch (2007) calculó el perfil de densidad nebular correspondiente a este caso. Al igual que para la nebulosa de masa mínima, Desch dividió al disco protoplanetario en anillos centrados en las órbitas de cada uno de los planetas (pero ahora, según lo establecido por 117 el modelo de Niza) y distribuyó uniformemente la masa total de sólidos que ellos contienen para poder estimar la densidad del disco protoplanetario. Dado que no todo el material sólido presente en el disco es incorporado por los planetas, ya que diversos mecanismos lo remueven de la zona de alimentación de los planetas antes de ser acretados (migración, fotoevaporación, etc.), la nebulosa solar estará enriquecida respecto de la masa total de sólidos estimada hoy en día. De acuerdo a estas hipótesis, Desch encontró que el perfil de densidad de la nebulosa compatible con el modelo de Niza decrece con la distancia en forma más abrupta que en el caso de la nebulosa estándar, aunque siguiendo también una ley de potencias, análogo a la ecuación (4.2). Entonces, Σ = Σ0 a−p , con p = 2,168. La relación entre la masa de sólidos y de gas se establece en 1 en 100, en correspondencia con la abundancia solar. En base al modelo de Niza y al perfil nebular calculado por Desch, realizamos simulacio- nes para estudiar la formación de Júpiter, Saturno, Urano y Neptuno. Los radios orbitales adoptados fueron: aJ = 5, 5 UA, aS = 8, 3 UA, aU = 14 UA y aN = 11 UA. Notemos que estamos aceptando que, después de su formación y durante su migración hacia el exterior del Sistema Solar, Urano y Neptuno intercambiaron lugares. El estudio de la formación de cada planeta se hizo en forma independiente, lo cual significa que no están contempladas las interacciones entre ellos, ni cómo la presencia de cada uno podría afectar la formación del resto. Como ya mencionamos en el Capítulo 4, nuestro modelo no considera la migra- ción planetaria, sino que la formación es calculada in situ. Thommes, Matsumura & Rasio (2008) calcularon la evolución de sistemas planetarios compatibles con el modelo de Niza y encontraron que, de sus simulaciones, surgen análogos al Sistema Solar siempre y cuando, durante su formación, los planetas gigantes no hayan sufrido una migración empinada y se mantuvieran en órbitas circulares. Esto le da validez a nuestra hipótesis. Es importante hacer énfasis en que la migración de los planetas gigantes predicha por el modelo de Niza ocurriría una vez disipada la nebulosa solar, y después de que los planetas han alcanzado su masa final. Hemos considerado que en el disco de sólidos la densidad superficial dada por: 5,5 !−2,168 Σ = 11 a La densidad superficial de gas nebular sigue el mismo perfil de la densidad de sólidos, y ambas reproducen la abundancia solar. La temperatura de la nebulosa, en el caso de Júpiter, se tomó en ∼ 130 K, mientras que para los tres planetas restantes se adoptó la mínima compatible con las tablas de la EOS, T = 125 K. En cuanto al tamaño de los planetesimales, aceptaremos que siguen una distribución de masa según lo propuesto por Kokubo & Ida (2000), con n(m) ∝ mα . Tomaremos como valor de referencia α = −2, 5. Consideraremos que la población de planetesimales está representada por 9 especies, con rmin = 0, 03 km y rmax = 100 km. La distribución de tamaños la hacemos según lo explicado en la ecuación (6.18). Dado que los planetesimales sufren un decaimiento orbital debido a la viscosidad de la nebulosa, no todos podrán g cm−2 . (6.19) 118 ser acretados en forma eficiente. En particular, los más pequeños son los que se verán más afectados por este efecto. Si se estiman las escalas de tiempo de migración de los planetesimales y de acreción por parte del protoplaneta para diversos tamaños, se puede obtener en límite inferior para el radio de los planetesimales a partir del cual son válidas las hipótesis de nuestro planteo. De esta estimación es que surge como valor adecuado un radio mínimo de 30 m. Para el caso de los resultados que presentaremos a continuación, hemos simulado la presencia de otros embriones creciendo en la vecindad del planeta del cual estamos calculando la formación, poniéndole una restricción a la tasa de acreción de sólidos. Al igual que hicieron Hubickyj, Bodenheimer & Polack (2005), luego que se alcanza el máximo de la acreción de sólidos, inhibimos el aumento de la masa de sólidos que puede acretar el planeta. Esto es, se considera que la zona de alimentación no se repuebla, sino que a partir de este máximo el planeta solo tiene posibilidad de acretar el material que había presente hasta ese momento en su zona de alimentación. Sobre la base del modelo que acabamos de delinear realizamos las simulaciones para la formación de Júpiter, Saturno, Urano y Neptuno. Las condiciones iniciales nebulares en las ubicaciones de los cuatro planetas gigantes y los resultados de nuestras simulaciones se listan en la tabla 6.4. En el caso de Júpiter y Saturno, las simulaciones involucran el cálculo del runaway de gas, mientras que para Urano y Neptuno los cálculos se terminan artificialmente cuando alcanzan su masa actual. Como se puede observar, en todos los casos el tiempo de formación es inferior a los 10 millones de años que imponen las observaciones de los discos circumestelares. En la figura 6.10 se muestra la evolución de la masa total y de la masa del núcleo en cada uno de los casos. Es evidentemente notorio el acuerdo entre las masas finales de los núcleos que surgen de las simulaciones y las estimaciones más recientes que surgen de los modelos del interior planetario. En el Capítulo 2 hicimos Tabla 6.4. Condiciones nebulares y resultados para la formación de los planetas gi- gantes del Sistema Solar. El radio orbital, a, y la densidad se sólidos inicial, Σ0 , surgen como consecuencia del modelo de Niza. La masa de aislación para cada uno de los ca- sos, Miso , se lista con fines comparativos. La masa del núcleo y el tiempo de formación que surge de nuestras simulaciones se encuentran en muy buen acuerdo con los límites observaciones en los cuatro casos. Dado que Urano y Neptuno no llegan al runaway de gas, pues alcanzan su masa final antes, presentamos aquí los resultados finales para la masa del núcleo y el tiempo de formación, y no la masa y el tiempo de cruce como en los ejemplos anteriores. Planeta a [UA] Σ0 [g cm−2 ] Miso [M⊕ ] Mc [M⊕ ] t [106 años] Júpiter Saturno Neptuno Urano 5, 5 8, 3 11, 0 14, 0 11, 0 4, 5 2, 4 1, 4 15, 3 13, 9 12, 8 12, 3 13, 34 12, 24 11, 48 11, 00 0, 52 1, 71 4, 00 6, 65 119 mención del estado actual del conocimiento de la estructura de los planetas gigantes. Las incertezas en la EOS, y la falta de datos observacionales más precisos (sobre todo en el caso de Urano y Neptuno), hacen que la masa del núcleo de estos planetas pueda ser estimada muy pobremente. En el caso de Júpiter, Saumon & Guillot (2004), utilizando la EOS de SCVH, encontraron que sus modelos de estructura son compatibles con una masa para el núcleo que se encuentre entre 0 y 12 M⊕ . Recientemente, Nettelmann et al. (2008) y Militzer et al. (2008) hicieron nuevas estimaciones, independientes entre sí , utilizando ecuaciones de estado derivadas a partir de primeros principios. Sin embargo, sus estimaciones no son concordantes. Mientras que Nettelmann et al. ubican a la masa del núcleo en el intervalo [0,7] M⊕ , Militzer et al. lo hacen en [14,18] M⊕ . Militzer & Hubbard (2008) argumentaron que esta diferencia se debe al tratamiento que ambos grupos le dieron a la transición entre el hidrógeno molecular e ionizado. Si bien nuestras simulaciones fueron realizadas utilizando la EOS de SCVH, por cuanto lo más coherente sería limitarnos a comparar nuestros resultados con los de Saumon & Guillot (2004), la controversia que existe sobre cuál es la masa de los núcleos de los planetas gigantes pone de manifiesto que los valores que se manejan actualmente solo pueden tomarse como una referencia aproximada. En el caso de Saturno, las estimaciones más recientes son las de Saumon & Guillot (2004), que establecen que la masa del núcleo podría tomar valores entre 9 y 22 M⊕ . Los casos de Urano y Neptuno son los más inciertos debido a la falta de datos precisos provenientes de sondas espaciales. Las cotas superiores para la masa del núcleo, de acuerdo con Guillot (2005) son 14 y 16,6 M⊕ para Urano y Neptuno respectivamente, mientras que las cotas inferiores son 9 y 12,4 M⊕ según las estimaciones de Podolak, Podolak & Marley (2000). El logro de formar a los cuatro planetas gigantes en tiempos razonablemente cortos y con masas para los núcleos compatibles con las estimaciones actuales se debe al hecho de considerar que los planetesimales del disco siguen una distribución donde los más pequeños son los más abundantes. Como vimos en la sección previa, considerar planetesimales chicos produce una aceleración significativa del proceso de formación. Esto nos permite mantener ba ja la densidad superficial de sólidos, dado que la masa del núcleo no superará, en general, la masa de aislación. El hecho quizá más relevante es que, con mismo perfil de densidad nebular y sin la necesidad de hipótesis extra, se logra dar cuenta de la formación de los cuatro planetas en menos 107 años. Este tipo de resultados no habían sido obtenidos hasta el momento y forman parte de un artículo recientemente aceptado para su publicación (Benvenuto, Fortier & Brunini 2009). El efecto de frenado por la viscosidad del gas favorece la captura de los planetesimales más pequeños. Esto provoca que la tasa de acreción sea más alta para los planetesimales en el extremo inferior de la distribución, pero también implica que son los primeros en agotarse. En la figura 6.11 vemos la evolución de la densidad superficial de sólidos correspondiente a cada una de las especies consideradas. Los planetesimales de 30 m de radio se agotan en 2 × 105 años, mientras que la población de los más grandes, de 100 km, prácticamente no varía en ese intervalo de tiempo. Notemos, además, que al principio de la formación, la densidad superficial se mantiene constante y que, al cabo de cierto tiempo (distinto para cada una de las especies), la pendiente de la curva comienza a variar rápidamente. Esto 120 10 ] ⊕ M [ a s a M 1 0.1 J S N U 1 6 Tiempo [10 años] 10 Figura 6.10. Evolución de la masa durante la formación de Júpiter (J), Saturno (S), Urano (U) y Neptuno (N). En línea roja se muestra la masa total y en azul la masa del núcleo. Las simulaciones se hicieron en base a las condiciones de borde que surgen del modelo de Niza, y considerando que los planetesimales acretados siguen una distribución de tamaños. Las líneas verticales representan distintas estimaciones de la masa del núcleo para los cuatro planetas: en verde la de Militzer et al. (2008), en turquesa la de Nettelmann et al. (2008), en fucsia la de Saumon & Guillot (2004) y en amarillo la de Guillot (2005) para los límites superiores y Podolak, Podolak & Marley (2000) para los inferiores. 121 ocurre cuando la viscosidad comienza a ser efectiva, y aumenta el radio de captura (Ec. 4.15). Dado que el efecto que tiene la fricción gaseosa en el frenado de los planetesimales es inversamente proporcional a su radio, los más pequeños son los primeros en experimentarlo, por cuanto el decaimiento en la densidad superficial de sólidos ocurre antes en estos casos. El valor de α fue elegido igual a -2,5, pero en realidad la restricción para α, se- gún Kokubo & Ida (2000), es que esté acotado al intervalo [−3, −2]. Para estudiar la sensibilidad de los resultados con la potencia de la distribución realizamos las simula- ciones para los casos de Júpiter, Saturno, Urano y Neptuno para otros 4 valores de α (α = −2, 17, −2, 33, −2, 67, −2, 83). En la tabla 6.5 mostramos la fracción, respecto de la masa total, asociada a cada una de las 9 especies de planetesimales consideradas. Cuanto más chico es el valor de α, más pronunciada es la distribución, con lo cual la fracción de planetesimales pequeños es mayor. Eso hace que la formación sea más rápida y que la masa ] 2 m c / g [ Σ 10 1 0.1 0.01 0.001 0.0001 1e−05 ΣT Σ1 Σ2 Σ3 Σ4 Σ5 Σ6 Σ7 Σ8 Σ9 0 0.1 0.2 0.3 0.4 0.5 0.6 6 Tiempo [10 años] Figura 6.11. Evolución de la densidad superficial de sólidos durante la formación de Júpiter. La densidad total, ΣT , es la suma de las densidades correspondientes a cada uno de los tamaños de los planetesimales considerados. La población de planetesimales está representada por nueve especies cuyos radios, de menor a mayor, son: r1 = 0, 03 km, r2 = 0, 08 km, r3 = 0,23 km, r4 = 0, 63 km, r5 = 1, 73 km, r6 = 4, 77 km, r7 = 13, 16, r8 = 36, 27 km y r9 = 100, 00 km. En el gráfico, Σi = Σ(ri ). 122 del núcleo sea más grande (6.6). A medida que α se acerca a -3, la variación de los resultados es menos importante. Esto se debe a que los planetesimales más chicos de la distribución contienen casi el total de la masa. Notemos que, mientras que para α = −2, 17 solo el 40 % de la masa está contenida en planetesimales de un radio de 30 m, para α = −2, 83 estos mismos planetesimales contienen el 92 %. Solo en el caso de α = −2, 17, el tiempo de formación de Urano supera los 107 años. Para el resto de las simulaciones, los tiempos de formación de todos los planetas son inferiores a esta cota. Por último, vamos a analizar la dependencia del tiempo de formación del planeta con el tamaño del planetesimal más pequeño de la distribución. Consideramos nuevamente que α = −2, 5, pero ahora la población de planetesimales estará compuesta por 7 especies, con rmin = 0, 1 km y rmax = 100 km. Los resultados son, para Júpiter, tf = 0, 89 My; para Saturno, tf = 3, 01 My; para Neptuno, tf = 6, 39 My; y para Urano, tf = 12, 6 My. El tiempo de formación se extiende en un 70 % para Júpiter y en casi un 90 % para Urano comparado con los casos de 9 especies. Esto pone de manifiesto, una vez más, la importancia del tamaño de los planetesimales acretados, y cómo la especie más pequeña de la distribución regula la escala del tiempo de formación. 6.3. Resultados inesperados: eventos cuasi periódicos de variación de la masa de la envoltura Durante el desarrollo de todo nuestro traba jo nos hemos encontrado con simulaciones que mostraban alteraciones en el proceso de acreción de gas respecto de la evolución espe- Tabla 6.5. Fracción respecto de la masa total para cada especie de planetesimales, con α entre -2 y -3. La primera columna muestra los valores de los radios representantes de cada una de las 9 especies de la distribución. Para 5 valores de α, las entradas de la tabla corresponden a la fracción de planetesimales asociada a cada una de las especies. r [km] α = −2, 17 α = −2, 33 α = −2, 5 α = −2, 67 α = −2, 83 0, 03 0, 08 0, 23 0, 63 1, 73 4, 77 13, 16 36, 27 100, 00 4, 01 × 10−1 2, 42 × 10−1 1, 45 × 10−1 8, 78 × 10−2 5, 28 × 10−2 3, 18 × 10−2 1, 91 × 10−2 1, 15 × 10−2 6, 96 × 10−3 6, 37 × 10−1 2, 31 × 10−1 8, 38 × 10−2 3, 04 × 10−2 1, 10 × 10−2 4, 00 × 10−3 1, 45 × 10−3 5, 27 × 10−4 1, 91 × 10−4 7, 81 × 10−1 1, 70 × 10−1 3, 73 × 10−2 8, 15 × 10−3 1, 78 × 10−3 3, 89 × 10−4 8, 50 × 10−5 1, 85 × 10−5 4, 06 × 10−6 8, 68 × 10−1 1, 14 × 10−1 1, 50 × 10−2 1, 97 × 10−3 2, 60 × 10−4 3, 42 × 10−5 4, 51 × 10−6 5, 93 × 10−7 7, 81 × 10−8 9, 20 × 10−1 7, 29 × 10−1 5, 78 × 10−3 4, 58 × 10−4 3, 63 × 10−5 2, 88 × 10−6 2, 28 × 10−7 1, 81 × 10−8 1, 43 × 10−9 123 Tabla 6.6. Tiempo de formación y masa del núcleo en simulaciones considerando 9 especies de planetesimales con radios entre 0,03 y 100 km. Los resultados corresponden a cada uno de los 4 valores para α adoptados en la distribución de tamaños. El caso para α = −2, 5 aparece en la tabla 6.4. α = −2, 33 α = −2, 17 tf Mc tf Mc [My] [M⊕ ] [My] [M⊕ ] α = −2, 67 tf Mc [My] [M⊕ ] Planeta α = −2, 83 tf Mc [My] [M⊕ ] Júpiter Saturno Neptuno Urano 1, 70 3, 93 7, 16 13, 70 11, 78 10, 90 10, 00 9, 55 0, 80 2, 03 4, 15 8, 15 12, 90 12, 00 11, 00 10, 60 0, 44 1, 52 3, 11 6, 73 13, 61 12, 42 11, 67 11, 32 0, 38 1, 45 2, 95 6, 35 13, 73 12, 52 11, 72 11, 30 rada, que es la que hemos descripto en las secciones previas. En los casos que presentamos antes, la masa de la envoltura crece monótonamente durante la formación del planeta. Sin embargo, esto no siempre es así , como podemos observar de la figura 6.12. Existen situaciones donde la envoltura se vuelve inestable, provocando eventos cuasi periódicos de variación de su masa, involucrando en cada uno de los “períodos” la expulsión y subsi- guiente acreción de gas (Benvenuto, Brunini & Fortier 2007). La figura 6.12 muestra dos casos, a priori cualitativamente distintos, de estos eventos y de su impacto en la formación de un planeta. El primer panel corresponde a la formación de un planeta cuyo radio orbital es a = 5, 2 UA, considerando la densidad superficial para un disco de 3 NSMM y donde los planetesimales son todos del mismo tamaño, en este caso, de 1 km de radio. Como se puede observar, las inestabilidades en la masa de gas ocurren cuando el núcleo ha, prácticamente, completado su formación, y la envoltura comienza a ser una fracción no despreciable de la masa total. La amplitud de la variación en la masa de gas aumenta conforme pasa el tiempo. La finalización abrupta de esta simulación tiene que ver con la convergencia de nuestro código, la cual se ve dificultada o impedida después de algunos “períodos” para este tipo de casos. Sin embargo, la simulación presentada en el segundo panel muestra una situación diferente. Si bien también la envoltura presenta sucesivos eventos de pérdida de gas, la inestabilidad termina siendo amortiguada y la evolución del planeta continúa de la forma habitual. En lo que sigue llamaremos al primer ejemplo de inestabilidad InC (Inestabilidad Continua), y al segundo, InD (Inestabilidad Dampeada). En general, en la mayoría de las simulaciones en que encontramos inestabilidades de la envoltura, corresponden a casos análogos al InC. Lo que tienen en común todos los casos inestables es que ocurren cuando la masa del núcleo alcanza valores muy cercanos a la masa de aislación. Si bien no es una condición suficiente haber agotado el material que rodea al planeta para que se produzca la inestabilidad (como lo hemos comprobado en casos previos), sí parece ser una condición necesaria. 124 a= 5,2 UA Σ= 10 g cm−2 rm= 1 km ] ⊕ M [ a s a M 15 10 5 0 a= 3,5 UA Σ= 12 g cm−2 rm= 100 km ] ⊕ M [ a s a M 15 10 5 0 0 1 3 2 Tiempo [106 años] 4 0 20 60 40 Tiempo [106 años] 80 Figura 6.12. Evolución de la masa del núcleo (azul) y de la envoltura (fucsia) para dos casos donde se producen inestabilidades en la envoltura. El primer panel representa el caso de un protoplaneta ubicado en la posición de Júpiter (a = 5, 2 UA), donde la densidad superficial de sólidos corresponde a 3 NSMM y los planetesimales acretados tienen un radio rm = 1 km (caso InC). El segundo panel muestra la formación de un protoplaneta cuyo radio orbital es a = 3, 5 UA, tomando la densidad de un disco de 2 NSMM y planetesimales de radio rm = 100 km. En este caso (InD) el tiempo de formación supera ampliamente la vida media de un disco protoplanetario. Sin embargo, dejando de lado este hecho y analizando la curva en forma cualitativa, notamos que, a diferencia del caso del primer panel, la inestabilidad se amortigua y el planeta continúa su formación en el modo usual. Las inestabilidades están, por cierto, asociadas a la tasa de acreción de sólidos. Cuan- do los planetesimales de la zona de alimentación comienzan a agotarse, el planeta pierde una fuente fundamental de energía para compensar la contracción de las capas de gas. De este modo, la tasa de acreción de gas aumenta. Si este proceso lleva a que se incorporen grandes cantidades de gas en poco tiempo, la masa total del planeta aumenta y la zona de alimentación sufre una expansión brusca, poblándose con planetesimales de las regiones del disco vecinas al planeta. Esto provoca un gran aumento en la tasa de acreción de só- lidos, con el consiguiente aporte de energía por parte de los planetesimales. Sin embargo, la envoltura no logra transportar eficientemente esta energía hacia la superficie, y ésta resulta absorbida por el interior, provocando una expansión en las capas de gas (ver figura 6.13). Eventualmente, como consecuencia de esta dilatación, la masa de gas contenida en la esfera de radio Rp disminuye, lo cual significa que la masa total del planeta disminuye. Esto impide el crecimiento de la zona de alimentación que, mientras la masa de gas dismi- nuye, se mantiene constante con lo cual, los sólidos disponibles para ser acretados entran nuevamente en una escala descendente. El proceso se retroalimenta hasta que la tasa de acreción de planetesimales alcanza de nuevo valores suficientemente ba jos como para per- mitir que la tasa de acreción de gas cambie de signo, y vuelva a ser positiva. Es importante mencionar que en todo momento, tanto durante la acreción como durante la eyección de 125 masa, la envoltura se encuentra en equilibrio hidrostático, por cuanto estos eventos no son de carácter hidrodinámico. Notablemente, la inestabilidad no es una pulsación libre radial como la que se encuentra en algunas estrellas variables, como las Cefeidas. Mientras que la escala de tiempo de una pulsación radial es del orden de Rp/cs , que en estas casos sería de tan solo unas horas, los períodos que encontramos aquí son varios órdenes de magnitud mayores. Esto se debe a que estas inestabilidades son oscilaciones forzadas cuya escala de tiempo está asociada al proceso de acreción. Como mencionamos más arriba, las inestabilidades de la envoltura no son periódicas, con lo cual hay que ser cuidadoso a la hora de estudiar los mecanismos de excitación y amortiguamiento de las oscilaciones. Cuando se estudian pulsaciones es habitual calcular la integral del traba jo que realizan las capas de gas al cabo de un ciclo (Clayton 1968), W = Z dMr I P d V . En el plano P − V , si un elemento de masa a lo largo de un ciclo evoluciona en el sentido de las agujas del relo j, excita la inestabilidad. A la inversa, si va en sentido contrario a las agujas del relo j, la amortigua. En estos casos encontramos que, mientras que las capas más internas tienden a excitar la inestabilidad, las externas tienden a amortiguarla. Que las capas internas exciten la inestabilidad es esperable puesto que allí es donde liberan los planetesimales la mayor cantidad de energía. Sin embargo, cuando se calcula el traba jo total sobre todas las capas de gas del planeta se encuentra que es negativo, con lo cual el efecto neto es que el planeta tiende a amortiguar la inestabilidad. Si bien las curvas que caracterizan cada uno de los casos muestran diferencias en su forma (sobre todo en el caso de la luminosidad), no estamos seguros de que los dos tipos de (6.20) 2e−06 1e−06 0 ] S L [ L a= 5,2 UA LPl LT 2e−08 1e−08 0 ] S L [ L a= 3,5 UA LT LPl 3 3.1 Tiempo [10 3.3 3.2 6 años] −1e−08 50 52 54 56 58 60 6 años] Tiempo [10 Figura 6.13. Para cada uno de los casos presentados en la figura 6.12 (InC e InD), mostramos la luminosidad total del planeta (rojo) y la producida por la acreción de planetesimales (azul), para cuatro “períodos” en la acreción de gas. Notemos como ambas cantidades están en contrafase. 126 inestabilidad sean esencialmente distintos. Puede que, en los casos como el InC, simplemen- te ocurra que no lo hemos podido dejar evolucionar completamente, y que si lo hiciéramos, en algún momento llegaría a alcanzar las condiciones para amortiguar la inestabilidad. Lo que sí es claro en ambos casos es que este fenómeno, cuando se desencadena, tiene como principal efecto demorar la formación de un planeta. El fenómeno de la inestabilidad de la envoltura es sumamente complejo y todavía no lo hemos comprendido en profundidad. Más traba jo en relación a este tema queda por delante. 127 128 Capítulo 7 Discusión En el capítulo anterior presentamos los resultados más importantes obtenidos con nues- tro modelo de formación de planetas gigantes. Otros autores (Bodenheimer & Pollack 1986, Wuchterl 1991, Pollack et al. 1996, Hubickyj, Bodenheimer & Lissauer 2005, Alibert et al. 2005, Dodson-Robinson et al. 2008) han realizado cálculos ba jo hipótesis similares a las adoptadas para este traba jo. En lo que sigue, discutiremos nuestros resultados en el marco del estado actual del entendimiento de este problema. Bodenheimer & Pollack (1986) fueron los primeros en realizar modelos numéricos de formación de planetas gigantes, de acuerdo a la hipótesis de inestabilidad nucleada, resol- viendo en forma autoconsistente las ecuaciones diferenciales de la estructura gaseosa. Sus simulaciones, si bien con muchas simplificaciones, formaron parte del primer traba jo que estudiaba en forma integral, y considerando los aspectos físicos más importantes, el proceso de formación de un planeta con un núcleo sólido y una extendida envoltura gaseosa. Diez años más tarde, Pollack et al. (1996) actualizaron el código de Bodenheimer & Pollack, pre- sentando un modelo mucho más detallado de formación de planetas gigantes, que hoy sigue teniendo vigencia, y que es considerado referente en el área por el grado de aproximación con el que se trataron ciertos aspectos físicos del problema. A continuación, describiremos sus resultados para poder realizar una posterior comparación con los obtenidos con nuestro modelo. El modelo de Pollack et al. (1996), en adelante P96, para el cálculo de la formación de planetas gigantes contiene tres ingredientes fundamentales: un modelo para el cálculo de la componente gaseosa del planeta, la incorporación de una tasa de acreción de sólidos dependiente del tiempo y un modelo para el intercambio energético entre los planetesimales ingresantes y las capas de la envoltura, el cual tiene en cuenta la ablación que sufren los planetesimales al atravesar la envoltura. Las diferencias más importantes entre nuestro modelo y el de P96 están en la elección de la tasa de acreción de sólidos y el tratamiento que se hace cuando los planetesimales atraviesan las capas de gas de la envoltura en su trayectoria hacia el núcleo del planeta. Nos focalizaremos primero en el modelo de acreción de planetesimales de P96 ya que 129 este es el ingrediente que marca la diferencia entre los resultados de P96 y los nuestros. En el capítulo 4 hemos discutido largamente el proceso de crecimiento de un embrión sólido, puntualizando que, en el caso de los núcleos de los planetas gigantes, se atraviesan dos etapas: las correspondientes al régimen runaway y al régimen oligárquico. El régimen runaway es de corta duración comparado con las escalas involucradas en la formación de los planetas gigantes (Ida & Makino 1993, Kokubo & Ida 2000), y sólo sería efectivo hasta que los embriones alcanzan una masa comparable a la lunar. Para masas superiores, la dispersión de velocidades que provoca el embrión sobre los planetesimales que lo rodean lentificaría su crecimiento en forma considerable. Sin embargo, P96 incorporaron a su modelo una tasa de acreción de sólidos que favorece el rápido crecimiento de los embriones y que, sin ser estrictamente la correspondiente al caso runaway, tiene características que lo aproximan más a este tipo de crecimiento que al régimen oligárquico. P96 fundamentaron la elección de este tipo de acreción en que les permitiría establecer cotas inferiores para el tiempo de formación de un planeta gigante y, fundamentalmente, en la diferencia sustancial que esta incorporación establecía respecto de los modelos previos donde la tasa de acreción se consideraba constante. El modelo de acreción empleado por P96 es el de Lissauer (1987), y podemos resumirlo de la siguiente manera. De acuerdo con su ecuación (1), la tasa de acreción del núcleo se puede escribir como, dMc = πR2ΣΩkFg dt donde R es el radio de captura, Ωk es la velocidad angular kepleriana, y Fg es el factor que tiene en cuenta el enfocamiento gravitatorio. Fg depende de los valores cuadrático medios de la excentricidad e inclinación reducida de los planetesimales, como también del radio reducido de captura: (7.1) e . a RH i a RH dc ≡ R eh,m ≡ ih,m ≡ RH Las fórmulas analíticas para Fg fueron obtenidas por Greenzweig & Lissauer (1992) a partir de simulaciones numéricas de N -cuerpos. En su modelo, Greenzweig & Lissauer proponen que las inclinaciones de los planetesimales solo dependen de las interacciones planetesimal- planetesimal (no se considera la presencia del embrión). En este caso, la inclinación puede aproximarse por: ve,m√3ΩkRH donde ve,m es la velocidad de escape de la superficie de un planetesimal. Claramente, la inclinación, i = ih,mRH/a, permanece constante en el tiempo mientras que la inclinación reducida, ih,m , decrece (dado que RH aumenta con la masa y a está fijo). En cuanto a la excentricidad, se supone que está afectada tanto por la presencia del embrión como por el resto de los planetesimales. El valor de la excentricidad reducida corresponde a: ih,m = (7.2) eh,m = max(2ih,m , 2). (7.3) 130 Esto significa que, si eh,m = 2ih,m , el embrión crece de acuerdo al régimen runaway, ya que en ese caso la excentricidad y la inclinación de los planetesimales acretados resultan independientes de la masa del embrión. Por otra parte, si eh,m = 2, la excentricidad de los planetesimales estará afectada por la presencia del embrión. Esta condición corresponde a la dispersión que sufren los planetesimales cuando se encuentran dominados por la ciza- lladura kepleriana (shear-dominated regime). Sin embargo, en base a sus simulaciones de N -cuerpos, Ida & Makino (1993) mostraron que la dispersión protoplaneta-planetesimal ba jo las condiciones de este régimen solo dura unos miles de años, luego de los cuales los planetesimales son fuertemente perturbados por la presencia del protoplaneta, con lo cual el régimen de crecimiento post-runaway del embrión pertenece al régimen donde do- mina la dispersión (dispersion-dominated regime). En este régimen, las excentricidades e inclinaciones satisfacen eh,m/ih,m ≃ 2 y eh,m > 2. Luego, cuando se considera que los plane- tesimales se encuentran en el shear-dominated regime, sus excentricidades e inclinaciones se mantienen ba jas, aún en la vecindad del embrión. Adoptar esta hipótesis durante todo el proceso de formación de un planeta significa considerar una tasa de crecimiento para los sólidos muy alta, o sea, un régimen de acreción considerablemente más efectivo que el correspondiente al crecimiento oligárquico. Además de la elección de distintos regímenes de acreción de sólidos, nuestro modelo y el de P96 difieren en que P96 consideran un modelo detallado para el intercambio energético entre los planetesimales acretados y la envoltura del planeta, donde se contempla la ablación que sufren los planetesimales en este proceso. Si bien considerar la ablación es tener en cuenta un efecto importante para el cálculo de la sección eficaz de captura de un planeta (Benvenuto & Brunini 2008), este efecto se hace realmente apreciable cuando el núcleo supera, aproximadamente, las 10 M⊕ . Como veremos a continuación, la diferencia entre nuestros resultados y los de P96 no pueden deberse a la ablación puesto que se manifiestan mucho antes de que el embrión tenga suficiente gas ligado como para que este efecto sea importante. Para evaluar los efectos de adoptar al crecimiento oligárquico como regulador de la formación del núcleo de los planetas gigantes en lugar de un régimen más cercano al runaway, como en el caso de P96, hicimos una serie de simulaciones ba jo las mismas condiciones que estos autores. P96 utilizan como caso de referencia para sus resultados (Caso J1) una simulación caracterizada de la siguiente manera. Un embrión sólido del tamaño de Marte (∼ 0, 1 M⊕ ), ubicado en la posición actual de Júpiter (a = 5, 2 UA), crece acretando planetesimales de 100 km de radio y de 1, 39 g cm−3 de densidad. La densidad del núcleo se acepta constante en 3, 2 g cm−3 . Las condiciones nebulares se caracterizan por: T = 150 K, ρ = 5 × 10−11g cm−3 y Σ = 10 g cm−2 . Los resultados de P96 para la evolución de la masa del planeta y el comportamiento de la tasa de acreción de sólidos y gas pueden observarse en la figura 7.1, la cual corresponde al artículo original de estos autores y se muestra aquí para la mejor comparación de los resultados. Bajo estas circunstancias, P96 encuentran que la masa de cruce es Mcross ≃ 16, 2 M⊕ y, el tiempo de cruce, tcross = 7, 6×106 años. Tanto en este, como en los otros resultados que estos autores muestran en su artículo, ellos distinguen que el proceso de formación atraviesa tres fases. La primera (“fase 1”) 131 involucra el principio del proceso y corresponde básicamente a la formación del núcleo. En el Caso J1, esta etapa abarca los primeros 6 × 105 años, que es el tiempo que le demanda al núcleo alcanzar unas 10 M⊕ . En este período, la tasa de acreción de sólidos alcanza su pico máximo, luego de lo cual decrece sensiblemente. En este caso, el embrión alcanza su masa de aislación casi en ausencia de la envoltura. Debido al vaciamiento de la zona de alimentación, pocos planetesimales ingresan en la envoltura después que esto ocurre, lo que permite el comienzo de la acreción de gas de una manera más significativa. P96 definen la finalización de la “fase 1” y el comienzo de la “fase 2” cuando la tasa de acreción de sólidos se iguala a la de gas. La “fase 2” es la etapa que regula el tiempo de formación del planeta puesto que corresponde al proceso de acreción de gas para la formación de la envoltura, y finaliza con el comienzo del runaway gaseoso. La marcada longitud de la “fase 2” se debe a que, si bien la tasa de acreción de sólidos se reduce considerablemente, la energía que aportan los planetesimales acretados durante esta etapa es lo suficientemente alta como para impedir la rápida contracción de las capas de gas. Entre el inicio y la finalización de la “fase 2” transcurren 7 × 106 años. Durante este tiempo, la relación entre la tasa de acreción de gas y de sólidos es constante, siendo la tasa de acreción de gas superior en un factor entre 2 y 3. La “fase 2” concluye cuando se alcanza la masa de cruce y comienza el runaway de gas, este último define la “fase 3”. Esta última etapa caracteriza la finalización del proceso de formación, que ocurre en menos de 5 × 105 años. Figura 7.1. Esta figura corresponde a los dos primeros paneles de la Fig. 1 del artículo de Pollack et al. (1996) y se muestra aquí para facilitar su comparación con nuestros resultados (ver figura 7.2). La simulación corresponde al Caso J1, donde la densidad superficial de sólidos es σinit = 10 g cm−2 en la posición del planeta (a = 5, 2 UA). El panel a muestra la masa acumulada de gas, MXY , de sólidos, MZ , y la masa total, Mp . El panel b muestra el logaritmo de la tasa de acreción de sólidos (línea llena) y de gas (línea punteada) en función del tiempo. P96 definen tres fases en la formación del planeta: la “fase 1”, que comprende desde el principio de la formación del núcleo MZ = MXY , extendiéndose solo por 6 × 105 años; la “fase 2” que domina hasta que MZ / MXY ≃ constante, y que finaliza cuando se la formación y está caracterizada por alcanza la masa de cruce; y, por último, la “fase 3” que corresponde al runaway de gas. 132 Caso J1 Mc Mg Mp ] ⊕ M [ a s a M 90 80 70 60 50 40 30 20 10 0 ] o ñ a / ⊕ M [ ) t d / M d ( g o l −2 −4 −6 −8 −10 Caso J1 log(dMc/dt) log(dMg/dt) 0 5 10 15 6 Tiempo [10 años] 20 25 0 5 10 15 6 Tiempo [10 años] 20 25 Figura 7.2. Nuestros resultados de la evolución de la masa del planeta con el tiempo (primer panel) y de las tasas de acreción (segundo panel) para las mismas condiciones del caso J1 de P96. Mc representa la masa del núcleo, Mg la masa de la envoltura y Mp la masa total del planeta. Teniendo en cuenta las mismas condiciones que caracterizan el Caso J1 de P96, hicimos una simulación utilizando nuestro código. Como podemos observar en la figura 7.2, nuestros resultados son muy diferentes a los de P96. En nuestro caso, el tiempo de cruce es ∼ 23 millones de años (Tabla 7.1), más de tres veces el obtenido por P96 y superior a la cota de 107 años que surge de las observaciones de discos circumestelares. De hecho, ni siquiera la masa del núcleo llega a alcanzar 10 M⊕ en menos de 107 años (Haisch 2001). Sin embargo, la masa de cruce es muy similar en las dos simulaciones. En ambos casos ésto ocurre cuando el núcleo alcanza aproximadamente la masa de aislación (para la densidad de sólidos considerada, Miso ≃ 16, 45 M⊕ para el caso en que Mg = Mc ). Por otra parte, no resulta evidente en nuestros resultados la necesidad de distinguir entre la “fase 1” y la “fase 2”. El proceso de acreción de gas se presenta en nuestra simulación como monótonamente creciente y con pendiente distinta de cero. En ningún momento encontramos una etapa donde la tasa de acreción de gas y de sólidos sean aproximadamente constantes ni proporcionales entre sí . La diferencia evidente entre ambos resultados son las escalas de tiempo que involucra cada uno. Mientras que la formación del núcleo en el caso de P96 toma alrededor de medio millón de años, en nuestro caso supera los 10 millones. Las largas escalas de tiempo relacionadas con la formación de embriones sólidos está asociada al régimen de crecimiento oligárquico. Si bien P96 contemplan la ablación de los planetesimales, lo cual favorece la aceleración de la formación, este efecto es importante una vez formado el núcleo, cuando el planeta comienza a tener una atmósfera más masiva, por cuanto no opera en forma eficiente en la primera etapa, con lo cual esta diferencia entre ambos modelos no puede ser responsable de la escala de tiempo tan corta que caracteriza la formación del núcleo en el modelo de P96. 133 P96 comienzan sus simulaciones con un embrión sólido de 0, 1 M⊕ . Nosotros lo hacemos con uno que es más de un orden de magnitud menor (0, 0075 M⊕), el cual se encontraría ya en pleno crecimiento oligárquico. De acuerdo a nuestra simulación del Caso J1, a nuestro embrión le toma 2 millones de años alcanzar la masa inicial de P96. Con lo cual, el tiempo que insume el crecimiento oligárquico de nuestro embrión para llegar a la condición inicial de P96 es cuatro veces mayor que el tiempo en que P96 llegan a la masa de cruce con un núcleo de 11 M⊕ . La existencia de la mencionada “fase 2”, en el sentido en que P96 la definen, fue estu- diada por Shiraishi & Ida (2008) con un modelo semi-analítico. Estos autores encuentran que la manifestación de tan largo período de acreción de gas es debida a la sobreestimación de la tasa de acreción de sólidos. En el modelo de referencia de P96 (Caso J1), la tasa de acreción de sólidos tiene un máximo del orden de 10−4 M⊕ / año, y durante la “fase 2” se mantiene en, aproximadamente, 10−6 M⊕/ año. Según los cálculos de Shiraishi & Ida, si la tasa de acreción de sólidos es ineficiente (como en el caso del crecimiento oligárquico), después que el núcleo alcanza la masa de aislación, la tasa de acreción, para este caso, sería del orden de 10−7 M⊕/año. En nuestra simulación, el máximo de la tasa de acreción es 10−6 M⊕ / año, y después de alcanzar la masa de aislación, en lo que podría tomarse como nuestra “fase 2”, si usamos la misma definición que P96, la tasa de acreción es, aproxi- madamente, 7 × 10−7 M⊕/año. El cálculo de Shiraishi & Ida tiene en cuenta que, después que el núcleo llega a unas 10 M⊕ , de abre una brecha con el disco de planetesimales. Esta brecha ocurre por la combinación de dos efectos: la excitación que provoca el embrión en los planetesimales que lo rodean y el decaimiento orbital que sufren por la viscosidad del gas nebular. Esto resulta en un truncamiento en la tasa de acreción de sólidos. En nuestro modelo no está contemplada la posibilidad de la apertura de una brecha de estas carac- terísticas. Sin embargo, aún considerando solo la dispersión que sufren los planetesimales en la zona de alimentación del planeta debido al efecto gravitatorio del embrión, se ve cla- ramente de nuestros resultados que la definición de la “fase 2” no tiene mucha relevancia puesto que no se registra un cambio notorio en el crecimiento del planeta ni en el tipo de acreción después que éste alcanza la masa de aislación. Tabla 7.1. Cuadro comparativo de los resultados obtenidos para tcross y Mcross en los casos J1, J3 and J7 de Pollack et al. (1996). Las columnas 2 y 3 corresponden a nuestros resultados, mientras que las 4 y 5 a los de Pollack et al. U significa que la envoltura es inestable. Caso tcross Mcross [M⊕ ] [My] tP96 cross M P96 cross [M⊕ ] [My] J1 J3 J7 23,25 13,20 U 15,75 21,4 U 7,58 1,51 6,94 16,17 29,61 16,18 134 P96 calcularon un caso donde la densidad superficial de sólidos es superior a la del Caso J1 (Σ = 15 g cm−2 ). Este aumento en la densidad lleva a que el tiempo de cruce se reduzca a 1,5 millones de años, mientras que la masa de cruce resulta, aproximadamente, 30 M⊕ (Caso J3, tabla 7.1). De acuerdo a nuestras simulaciones, el tiempo de cruce sería un orden de magnitud superior (13 millones de años), mientras que la masa de cruce resultaría inferior, aproximadamente 21, 5 M⊕ . En el caso de P96 el aumento de la densidad representa una reducción en el tiempo de formación de un factor 5, sin embargo, en nuestro caso este mismo incremento solo conduce a que el tiempo de formación se reduzca a la mitad. La tasa de acreción de sólidos depende inversamente de la velocidad relativa de los planetesimales respecto del protoplaneta, la cual está regulada por la excentricidad e in- clinación de los planetesimales que serán acretados (Ec. 4.16). Para el Caso J3, la figura 7.4 muestra el cociente entre la excentricidad (inclinación) de nuestro modelo y el de P96. Para este caso, ih,m es desde el comienzo inferior a 1 y, como es una función decreciente de la masa del protoplaneta, eh,m siempre es igual a 2. De este modo, nuestras excentricidades siempre son mayores a las de P96 en un factor 4. Además, la inclinación es constante en el caso de P96 (i ≃ 0, 004), mientras que en nuestro modelo es i ≃ e/2. Esto impacta en las velocidades relativas y en el cálculo de la altura de escala del disco de planetesimales, h. La tasa de acreción de sólidos es inversamente proporcional a h, con lo cual si h es más grande, porque los planetesimales están más dispersos, la tasa de acreción es menor. De este modo, siendo nuestras inclinaciones y excentricidades mucho mayores a las adoptadas en P96, es lógico que nuestros tiempos de acreción sean mucho más largos. Thommes, Duncan & Levison (2003) estimaron el tiempo la formación de un protopla- neta sólido en crecimiento oligárquico (sin envoltura gaseosa) en un disco protoplanetario donde la densidad corresponde a 1 y a 10 NSMM. En el caso de 1 NSMM, para la posición de Júpiter, al cabo de 107 de años solo llega a formarse un embrión de 1 M⊕ . En tanto, para 10 NSMM, un embrión de 10 M⊕ se forma en, aproximadamente, ∼< 5 × 106 años. El Caso J1 corresponde a 3 NSMM, con lo cual no es esperable conseguir la formación de un núcleo de 10 M⊕ en tiempos razonablemente cortos. Dado que previamente habíamos encontrado que la acreción es muy sensible al tamaño de los planetesimales acretados, volvimos a considerar las condiciones del Caso J3, pero para planetesimales de 1 y 10 km de radio (ver figura 7.3). En ambos casos, el tiempo de formación resulta bastante inferior a 107 años. Cuando rm = 10 km, el tiempo de cruce es 3,7 millones de años y la masa de cruce es 25, 5 M⊕ , mientras que para el caso de planetesimales de 1 km de radio, el tiempo de cruce se reduce a 1, 4 millones de años y la masa de cruce a 29 M⊕ . Como podemos notar, considerando planetesimales de 1 km de radio podemos reproducir los resultados de P96 para el Caso J3 (donde el radio de los planetesimales es de 100 km). El tamaño típico de los planetesimales no se conoce todavía. De hecho, como hemos mencionado previamente, la formación y consecuente distribución de tamaños de los planetesimales primordiales está todavía ba jo estudio y es uno de los temas fundamentales que debe resolver las Ciencias Planetarias. Si repetimos este procedimiento de reducir el tamaño de los planetesimales pero ahora 135 Mp Mc Mg rp= 1 km rp= 10 km rp= 100 km ] ⊕ M [ a s a M 80 70 60 50 40 30 20 10 0 0 2 4 6 8 6 Tiempo [10 años] 10 12 14 Figura 7.3. Tres simulaciones de la formación de Júpiter considerando solo variaciones en el tamaño de los planetesimales (rm = 1, 10, 100 km). El caso correspondiente a rm = 100 km fue calculado adoptando los mismos parámetros que Pollack et al. (1996) para su caso J3. para el Caso J1, encontramos que, para rm = 10 km, el tiempo de cruce es inferior a los 10 millones de años. El caso J 1√10 presentado en el capítulo 6 fue realizado ba jo las mismas hipótesis del Caso J1 pero con planetesimales de radio 3,17 km. Sin embargo, no siempre reducir el tamaño de los planetesimales acretados lleva a favorecer la formación de un planeta. Si para el Caso J1 consideramos el radio de los planetesimales en 1 km obtenemos el comportamiento oscilatorio que presentamos en el capítulo anterior (figura 6.12, primer panel). Mientras que P96 vuelven a encontrar una solución estable análoga a las anteriores (Caso J7), nosotros encontramos que la envoltura se vuelve inestable cuando la masa del núcleo es 12 M⊕ y la de la envoltura es 2, 43 M⊕ . Para estos valores, la masa del núcleo es, al orden de la aproximación del cálculo de la masa de aislación, igual a la masa de aislación para un objeto con esa cantidad de gas ligado. Si bien desconocemos el motivo por el cual nosotros encontramos esta inestabilidad y otros autores no, una posibilidad es que este hecho esté ligado al tratamiento de la acreción de sólidos. Dado que nuestra acreción es más lenta, para un mismo valor de la masa del núcleo, la masa de la envoltura asociada a nuestros modelos es mayor que en la de P96. Por ejemplo, en el Caso J7, cuando la masa del núcleo de P96 es 11,4 M⊕ , la masa de la envoltura es 0, 14 M⊕ , mientras que en nuestro 136 80 70 60 50 40 30 20 10 0 i / i* 0 0.1 e / e* ~ 4 0.2 RH [AU] 0.3 0.4 Figura 7.4. El impacto del modelo elegido para el crecimiento del núcleo. En línea llena se muestra el cociente entre las excentricidades e/e∗ , donde e corresponde a la calculada según la ecuación (4.19) y e∗ de acuerdo al modelo adoptado por Pollack et al. (1996). La línea punteada muestra el cociente entre las inclinaciones (i/i∗ ). Mientras que en nuestro modelo i depende de e, según i = e/2, Pollack et al. consideran a i∗ constante durante toda la simulación. En la abscisa consideramos al radio de Hill para poder hacer una comparación que no involucre las escalas de tiempo. caso, para esa misma masa del núcleo, la masa de gas es tres veces mayor, Mg ≃ 0, 46 M⊕ . De este modo, nosotros llegamos a la masa de aislación con una envoltura más masiva, lo cual la vuelve más propensa a desarrollar este tipo de inestabilidad. Es importante enfatizar que los planetesimales de mayor tamaño tienden a favorecer la estabilidad de la envoltura. Hemos ya mencionado que la tasa de acreción de sólidos depende del tamaño de los planetesimales. Dado que en el crecimiento oligárquico las velocidades relativas de los planetesimales más grandes son mayores, el radio efectivo en estos casos es menor, lo cual significa que la luminosidad por acreción es también menor. Sin embargo, son los planetesimales más pequeños los más abundantes en el disco, los cuales son también más eficientemente acretados, con lo cual favorecen el desencadenamiento de la inestabilidad. El hecho que no hayamos registrado este efecto en nuestras simulaciones considerando una distribución de tamaños está relacionado con que, si se dieran las condiciones para el desarrollo de la inestabilidad, la limitación que le hemos puesto a la tasa de acreción de sólidos las inhibirían. Sin embargo, si calculamos el caso J7, sin limitar la acreción y con una distribución de tamaños (análoga a la empleada en los ejemplos del capítulo 6) encontramos que la inestabilidad está igualmente presente. Por otra parte, cuanto más densa es la nebulosa y/o mayor el radio orbital del embrión, menor será la tendencia a 137 que se produzca la inestabilidad. Esto se debe a que en estos casos la masa de aislación es mayor y el runaway de gas se alcanza antes de llegar a las condiciones necesarias para que surjan las oscilaciones. Varios mecanismos podrían inhibir la aparición de la inestabilidad de la envoltura. Esencialmente, la inestabilidad aparece cuando el planeta alcanza la masa de aislación, con lo cual las capas de gas sufren una abrupta contracción, se acreta gas en consecuencia y la zona de alimentación se expande incorporando una gran cantidad de planetesimales. Ahora bien, cuántos planetesimales ingresan en la zona de alimentación cuando ésta se expande dependerá del estado del disco en ese momento. En nuestro modelo no hemos incluido la evolución propia del disco protoplanetario, por cuanto cuando se expande la zona de alimentación, las nuevas regiones que se le suman se encuentran en su estado inicial. Es esperable, sin embargo, que el número de planetesimales sea menor que el original. Debido a la viscosidad de la nebulosa, los planetesimales sufren un decaimiento orbital, espiralando en dirección al Sol. Los más pequeños serán, además, los más afectados. Además, la escala de tiempo de migración de los planetesimales es del mismo orden que la de la formación del planeta (Thommes, Duncan & Levison 2003), por cuanto es esperable que este hecho tenga una incidencia no despreciable. Otro efecto que debería tenerse en cuenta es la eyección de planetesimales fuera de la zona de alimentación. Dado que esto ocurre cuando el planeta es suficientemente masivo no retrasaría la acreción inicial para formar el núcleo, con lo cual no repercutiría en el tiempo de formación. Sin embargo, operaría en la dirección de disminuir la densidad superficial de sólidos en la vecindad del planeta cuando éste tiene una masa apreciable. En la bibliografía relativa a modelos de formación de planetas gigantes no hay registro de inestabilidades como las encontradas en nuestras simulaciones. Wuchterl (1991, 1995), con un código completamente hidrodinámico, encuentra que cuando el planeta está cercano a alcanzar la masa de cruce, atraviesa una inestabilidad dinámica debido a la cual eyecta gran parte de su envoltura. Sin embargo, la inestabilidad presente en nuestras simulaciones no es de carácter hidrodinámico, sino que puede ser descripta por una secuencia de modelos en equilibrio hidrostático debido a que la velocidad del material es suficientemente ba ja. Además, esta inestabilidad no está asociada al transporte de energía, sino a la interacción entre el proceso de acreción y la respuesta de la envoltura frente a cambios en la luminosidad de acreción de sólidos. Otros escenarios se han tenido en cuenta para acelerar la formación de los planetas gigantes. Hubickyj, Bodenheimer & Lissauer (2005) analizaron la posibilidad de que la opacidad de la envoltura fuera mucho menor que la correspondiente al medio interestelar. El modelo de estos autores es esencialmente igual al de P96, con la actualización de la EOS como principal diferencia (P96 no utilizaban la EOS de SCVH sino una tabla pre- via). Recalcularon el Caso J1, obteniendo Mcross ≃ 16, 2 M⊕ (igual que P96), pero con una reducción en el tiempo de cruce (tcross ≃ 6 millones de años). Esta misma simulación fue realizada suponiendo que la opacidad de los granos era tan solo un 2 % de la opacidad interestelar, de acuerdo con los cálculos de Podolak (2003). Sus resultados muestran que el tiempo de cruce se ve significativamente afectado por la opacidad, ya que en este caso 138 tcross ≃ 2, 2 millones de años. La masa de cruce, por otra parte, se mantiene en el mismo va- lor. Actualmente, no existen tablas de opacidades calculadas en forma autoconsistente con la formación del planeta. En cualquier caso, si la opacidad es menor que la considerada en nuestras simulaciones, el efecto que tendría sobre nuestros resultados sería el acortamiento de las escalas de tiempo, lo cual favorecería la formación de los planetas gigantes. Recientemente, Dodson-Robinson et al. (2008b) estudiaron la formación in situ de Jú- piter y Saturno. El modelo y el código empleado para los cálculos es el de Hubickyj, Bodenheimer & Lissauer (2005). El traba jo de Dodson-Robinson et al. se basa en cálcu- los detallados de la densidad superficial de sólidos realizados por Dodson-Robinson et al. (2008a). Para estimar la densidad superficial de sólidos, ellos combinan la evolución vis- cosa de un disco protosolar con un modelo cinético de formación de hielos, lo cual les permite obtener un perfil de densidad en función del tiempo y de la distancia heliocéntri- ca. Entre sus resultados encuentran que en la región trans-saturniana habría abundantes cantidades de amoníaco en forma de hielos, lo cual implicaría un enriquecimiento en un factor 6 respecto de la NSMM en la región de Saturno. Dodson-Robinson et al. (2008b) encontraron que Saturno se forma en 3, 5 × 106 años, sin llegar a agotar nunca su zona de alimentación. Tanto en el caso donde la opacidad se considera reducida como cuando se asumen los valores originales de la tabla, el tiempo de formación les da prácticamente igual. Sin embargo, la masa del núcleo que obtienen es 44 M⊕ en el caso de las opacidades de los granos reducidas, y 54 M⊕ en el otro caso. Estos valores son mucho mayores que los estimados para la masa del núcleo de Saturno, e incluso superarían los valores aceptados para explicar enriquecimiento registrado en su atmósfera. Otra consideración a tener en cuenta es que ellos comienzan sus simulaciones con un embrión de 1 M⊕ . Hemos mostrado ya que, el tiempo de formación de un embrión de esta masa no es para nada despreciable. Por su parte, Alibert et al. (2005a) tuvieron en cuenta la migración del planeta durante su formación y la evolución del gas nebular. Su modelo es de características similares al de P96, con el agregado que consideran que el disco protoplanetario no es estático, y que el planeta puede migrar en el disco debido a su interacción con el gas. Sus simulaciones muestran que el tiempo de formación de los planetas gigantes (al menos hasta que alcanzan el crecimiento runaway), se acelera notablemente cuando de tiene en cuenta la migración del planeta. Esto se debe fundamentalmente a que desaparece la “fase 2” puesto que la zona de alimentación del planeta no llega nunca a vaciarse ya que el planeta atraviesa siempre nuevas regiones en el disco. La migración suprime la necesidad de esperar a que la envoltura alcance una masa significativa para que el planeta llegue a la masa de cruce. Para simulaciones ba jo condiciones análogas a las de P96 en el Caso J1, Alibert et al. encuentran que el planeta llega a la masa de cruce en una escala de tiempo del orden del millón de años. Para conseguir esto, el embrión debe comenzar el proceso de acreción y migración a una distancia del Sol de entre 7 y 15 UA. Como mencionamos en el capítulo 3, las teorías de migración de los planetas en formación debido a su intercambio de momento angular con la nebulosa protoplanetaria tienen su origen a partir de las observaciones de los sistemas extrasolares, donde en muchos casos los planetas se encuentran muy cercanos a su estrella central. Sin embargo, todavía existen muchas dudas sobre cómo opera realmente 139 este mecanismo. De hecho, Alibert et al. deben reducir artificialmente la tasa de migración del embrión en, al menos, un factor 100 para impedir que los protoplanetas sean engullidos por el Sol. En un traba jo posterior, Alibert et al. (2005b), estudiaron la formación de Júpiter y Saturno considerando migración. Calcularon primero una serie de simulaciones, variando ciertos parámetros, para la formación de Júpiter. Una vez encontrados los parámetros para los cuales el resultado final concordaba con las características actuales de Júpiter (masa total, masa de sólidos, distancia al Sol, etc.), fijaban estos parámetros y buscaban la posición inicial y el retardo necesario para el comienzo de la formación de Saturno. Alibert et al. encontraron que, si Júpiter comenzaba su formación en a ∼< 10 UA, y Saturno en 11, 9 UA y 0,2 millones de años más tarde, la configuración final se correspondía con la actual. El tiempo de formación de ambos planetas era inferior a 3 ×106 años. Es importante mencionar que la formación de Júpiter y Saturno no fueron calculadas en forma simultánea, por cuanto no está contemplado en este modelo la interacción entre ambos protoplanetas. Alibert et al. obtuvieron muy buenos a justes entre su resultado final y los datos actualmente aceptados para estos planetas. Sin embargo, el escenario de formación que ellos sugieren contradice al modelo de Niza. Como mencionamos antes, el modelo de Niza consigue explicar numerosas características del Sistema Solar actual, por cuanto se lo considera robusto y confiable. En el traba jo donde Desch (2008) calcula el perfil de densidad de la nebulosa protosolar compatible con el modelo de Niza, también realiza estimaciones para la formación de los núcleos de los cuatro planetas gigantes del Sistema Solar. Considerando el crecimiento oligárquico de los embriones, pero en ausencia de una envoltura de gas (que aumenta el radio efectivo de captura), Desch encuentra que los tiempo de formación de los núcleos son: para Júpiter, ∼ 0, 5 × 106 años, para Saturno, ∼ 1, 5 − 2 × 106 años, para Neptuno ∼ 5, 5 − 6 × 106 años y para Urano, ∼ 10 × 106 años. El tamaño considerado para los planetesimales es rm = 0, 1 km. Si bien el modelo de Desch es mucho más simple que el nuestro, si comparamos estos resultados con los que obtuvimos nosotros para el caso de 7 especies de planetesimales (capítulo 6), donde el radio mínimo considerado para la distribución es 0,1 km, vemos que las escalas de tiempo involucradas en cada caso son del mismo orden. Es importante puntualizar que nuestro modelo no considera la evolución de la nebulosa protoplanetaria. Desch (2008) estimó los tiempos de formación de los núcleos de los planetas si se permite la evolución viscosa de la nebulosa. Mientras que, dada la rápida formación de Júpiter, este hecho no tendría un impacto significativo en este caso, y para Saturno el tiempo de formación de su núcleo sería de 3 millones de años, los tiempos se verían considerablemente prolongados en el caso del resto de los planetas: para Neptuno sería de 15 millones de años y para Urano superaría ampliamente los 20 millones de años. 140 Capítulo 8 Conclusiones y perspectivas El objetivo de esta Tesis es el estudio de la formación in situ de los planetas gigantes del Sistema Solar en el marco del modelo de inestabilidad nucleada. Para ello se traba jó con un modelo teórico y numérico que permite el cálculo de la tasa de acreción de sólidos y de gas en forma autoconsistente, las cuales son necesarias para un estudio realista de este problema. Si bien cálculos de esta naturaleza fueron realizados ya por otros autores (Bodenheimer & Pollack 1986; Wuchterl 1991, 1995; Pollack et al. 1996; Alibert et al. 2005, Hubickyj, Bodenheimer & Lissauer, 2005), son todavía muchas las incertezas entorno a este tema. Existen dos factores derivados de los datos observacionales que imponen restricciones a los resultados teóricos: la vida media de los discos circumestelares (estimada en menos de 107 años) y las masas de los núcleos de los planetas gigantes del Sistema Solar (en general, entre 10 y 20 M⊕ ). Es cierto que los valores de estas cotas no son estrictos y que hay lugar para cierta flexibilidad, pero muy probablemente puedan ser tomados como buenas aproximaciones. Encontrar soluciones a los modelos de formación, que se mantengan dentro de estos límites es una tarea que continúa entre los investigadores del área, ya que no se ha conseguido explicar en forma completamente satisfactoria la formación de Júpiter, Saturno, Urano y Neptuno en el marco de un modelo integral de evolución de un sistema planetario. El problema de la formación de los gigantes gaseosos es tan vasto, que no puede ser estudiado, por ahora, en conjunto y en forma completamente autoconsistente, con la evolución del disco protoplanetario. Sin embargo, puede ser aislado y atacado desde diversos ángulos que permitan la obtención de resultados parciales, los cuales, a su vez, den lugar a aproximaciones realistas de la solución. Teniendo presente este hecho es que hemos formulado nuestro modelo. Los aportes más importantes de nuestro traba jo son la incorporación de la tasa de crecimiento oligárquico para el núcleo sólido y las condiciones nebulares compatibles con el modelo de Niza. En estudios previos, la tasa de acreción de sólidos considerada era cercana a la correspondiente al régimen runaway, mucho más eficiente en la captura de planetesimales pero probablemente altamente sobreestimada. De hecho, en muchas de las 141 simulaciones donde se considera este tipo de crecimiento, la formación del núcleo ocurre en una escala de tiempo despreciable comparada con la correspondiente a la formación completa del planeta. Nuestro modelo es el primero en incorporar una tasa de acreción de sólidos realista. Las conclusiones más importantes, en base a los resultados de nuestras simulaciones y a la comparación con modelos previos, son: El crecimiento oligárquico del núcleo tiene impacto directo en la formación de un planeta gigante dado que regula, no solo la escala de tiempo de formación del núcleo, sino también la tasa de acreción de sólidos. El tiempo involucrado en la formación de un núcleo a partir del cual el planeta comienza a acretar gas en forma significativa representa, aproximadamente, la mitad del tiempo necesario para alcanzar la masa de cruce. En estudios realizados por otros autores, donde la acreción de planetesimales se considera mucho más efectiva, el tiempo de formación del núcleo puede, en la mayoría de los casos, considerarse despreciable. La tasa de acreción de sólidos en el crecimiento oligárquico es mucho más ba ja debido a la dispersión de velocidades de los planetesimales, provocada por la presencia del embrión. Este hecho tiene incidencia, a su vez, en la tasa de acreción de gas. Mientras que asociado al crecimiento runaway (o cuasi runaway) de sólidos está la existencia de la “fase 2” en la formación del planeta, la cual regula la escala de tiempo de toda la formación, esta fase está prácticamente ausente si se considera el crecimiento oligárquico del núcleo. Cuando se adopta este régimen de crecimiento, la tasa de acreción de gas resulta ser siempre estrictamente creciente. Debido a la dependencia de la tasa de acreción de sólidos con la densidad superficial en la zona de alimentación, cuando se aumenta la densidad, los tiempos de formación se acortan y las masas de los núcleos aumentan. Un resultado análogo surge cuando se consideran variaciones en las distancias heliocéntricas. Dado que la densidad decrece hacia el exterior del disco, el tiempo de formación de un planeta aumenta en esa dirección. El tiempo de formación decrece sensiblemente cuando se reduce el tamaño de los planetesimales acretados por el planeta. Dado que la viscosidad del gas que consti- tuye la envoltura es más efectiva sobre los cuerpos más pequeños, los planetesimales que la atraviesan sufren una mayor desaceleración por efecto del frenado viscoso y disminuyen sus velocidades relativas, lo cual facilita su acreción. En otras palabras, la sección eficaz de captura de los planetesimales de menor radio es mayor que la de los planetesimales más grandes, razón por la cual la formación del núcleo ocurre más rápidamente cuando la población de planetesimales en la zona de alimentación del planeta está dominada por cuerpos pequeños. Considerar una distribución de tamaños para los planetesimales acretados, donde los más pequeños contienen la mayor parte la masa del disco favorece la formación de los 142 planetas gigantes. Sin embargo, el tiempo de formación depende significativamente de la elección del radio del planetesimal más pequeño. El proceso de formación de los planetas gigantes del Sistema Solar es compatible con el modelo de Niza. El hecho que los cuatro planetas conformaran originalmente un sistema más compacto, con radios orbitales menores a los actuales, favorece el proceso de acreción. La formación de un planeta gigante se puede ver seriamente afectada si la envoltura se vuelve inestable y desarrolla procesos cuasi periódicos de pérdida de masa. La aparición de estas inestabilidades está asociada a situaciones de acreción donde el núcleo llega a alcanzar la masa de aislación. Las densidades superficiales ba jas y/o los planetesimales pequeños y/o las distancias heliocéntricas cortas favorecen la aparición de la inestabilidad. Variaciones en la ecuación de estado pueden tener serias implicancias en el tiempo de formación y en el valor de la masa del núcleo. El resultado fundamental de nuestro traba jo es haber encontrado que el régimen de crecimiento oligárquico no es incompatible con la formación de los planetas gigantes. De hecho, aún siendo éste un escenario poco favorable, hemos podido demostrar, por primera vez y ba jo hipótesis razonables, que Júpiter, Saturno, Urano y Neptuno pueden formarse en menos de 10 millones de años. Además, la masa de los núcleos resulta consistente con las estimaciones actuales. Por otra parte, la posibilidad de la manifestación de inestabilidades en la envoltura po- dría dificultar la formación de los planetas. Sin embargo, otros mecanismos podrían inhibir el desarrollo de la inestabilidad. En el futuro se debería incluir un modelo de evolución del disco protoplanetario que contemple tanto la dependencia con el tiempo de la densi- dad del gas nebular como el decaimiento orbital de los planetesimales por el efecto de la viscosidad. Además, habría que tener en cuenta la disminución de los planetesimales en la zona de alimentación del planeta por eyección de los mismos como consecuencia de las interacciones gravitatorias con el planeta. La sensibilidad que presentan los modelos, tanto de formación como los correspondientes a la estructura interior de los planetas gigantes, a variaciones en la ecuación de estado pone de manifiesto la necesidad de contar con tablas más detalladas. Resulta fundamental, a la hora de mejorar los modelos, poder contar con una ecuación de estado confiable. En el mismo sentido, el cálculo de la opacidad en la envoltura debería ser autoconsistente con el proceso de acreción. La ablación de los planetesimales y el cálculo detallado de la energía que estos depositan en las capas de gas es uno de los principales objetivos en la continuación de este traba jo. Dos cuestiones interesantes para explorar en relación con este problema están relaciona- das con las condiciones de borde, tanto internas como externas. Por un lado, en los estudios 143 como el presentado en esta Tesis se considera que el núcleo es inerte y que no participa energéticamente en el proceso. Esta hipótesis debería ser evaluada en más detalle. Nosotros hemos considerado que el núcleo no absorbe ni emite energía. Si, por el contrario, existiera un flujo de energía neto en alguna de las dos direcciones, el proceso de formación se vería seriamente afectado. Si hubiera un flujo neto de energía hacia el núcleo la formación sería más rápida dado que el colapso de las capas de gas se produciría antes justamente por no contar éstas con la presión necesaria para contrarrestar el “peso” de las capas exteriores. Si, en cambio, el núcleo fuera una fuente de energía con un flujo neto hacia el exterior el proceso de colapso se retrasaría. Creemos que lo más probable es que el flujo neto de energía se dirige hacia el núcleo, lo cual favorecería al proceso de formación. Este es un punto a explorar en nuestros próximos traba jos. En el otro extremo, la condición de borde en el límite entre la nebulosa y el planeta debería ser evaluada con más cuidado. Funda- mentalmente, cuando se desata el runaway de gas, la velocidad a la cual el disco puede suministrar material al planeta seguramente regulará el estadío final de la formación. La condición de borde externa determinará la masa final del planeta. Queda pendiente de estudio si este proceso modifica sustancialmente la escala de tiempo de acreción de gas. Por último, no hay que perder de vista que los planetas no se formaron como entidades aisladas en el disco sino que el proceso tuvo lugar durante la formación simultánea con otros embriones de su vecindad. Las interacciones entre ellos, y entre ellos con el disco puede ser determinante, sobre todo si, como se espera, los planetas no se mantuvieron en órbitas fijas durante su formación sino que sufrieron variaciones en sus semiejes. 144 Bibliografía Adachi, I., Hayashi, C., Nakazawa, K. 1976. The gas drag effect on the el liptical motion of a solid body in the primordial solar nebula. Progress of Theoretical Physics 56, 1756-1771. Alexander, D. R., Ferguson, J. W. 1994. Low-temperature Rosseland opacities. Astrophy- sical Journal 437, 879-891. Alibert, Y., Mordasini, C., Benz, W., Winisdoerffer, C. 2005. Models of giant planet for- mation with migration and disc evolution. Astronomy and Astrophysics 434, 343-353. Alibert, Y., Mousis, O., Mordasini, C., Benz, W. 2005. New Jupiter and Saturn formation models meet observations. Astrophysical Journal 626, L57-L60. Armitage, P. J. 2007. Lecture notes on the formation and early evolution of planetary systems. ArXiv Astrophysics e-prints arXiv:astro-ph/0701485. Barge, P., Sommeria, J. 1995. Did planet formation begin inside persistent gaseous vortices? Astronomy and Astrophysics 295, L1-L4. Bate, M. R. 2006. The brown dwarf-planet relation. Planet Formation: Theory, Observa- tions and Experiments 1. Cambridge University Press. Ed. H. Klahr & W. Brandner. Beaulieu, J.-P., and 72 colleagues. 2006. Discovery of a cool planet of 5.5 Earth masses through gravitational microlensing. Nature 439, 437-440. Becklin, E. E., Zuckerman, B. 1988. A low-temperature companion to a white dwarf star. Nature 336, 656-658. Benedict, G. F., McArthur, B. E., Forveille, T., Delfosse, X., Nelan, E., Butler, R. P., Spiesman, W. Marcy, G., Goldman, B., Perrier, C., Jefferys, W. H., Mayor, M. 2002. A mass for the extrasolar planet Gliese 876b determined from Hubble Space Telescope Fine Guidance Sensor 3 astrometry and high-precision radial velocities. Astrophysical Journal Letters, 581, L115-L118. Benvenuto, O. G., Brunini, A. 2005. Methods for computing giant planet formation and evolution. Monthly Notices of the Royal Astronomical Society 356, 1383-1395. Benvenuto, O. G., Brunini, A., Fortier, A. 2007. Envelope instability in giant planet for- mation. Icarus 191, 394-396. Benvenuto, O. G., Brunini, A. 2008. The effects of ablation on the cross section of planetary envelopes at capturing planetesimals. Icarus 195, 882-894. Benvenuto, O. G., A., Fortier, Brunini, A. 2009. Forming Jupiter, Saturn, Uranus and Neptune in few mil lion years by core accretion. Icarus, 204, 752-755. Bodenheimer, P. 2006. Historical notes on planet formation. Planet Formation: Theory, Observations and Experiments 1. Cambridge University Press. Ed. H. Klahr & W. 145 Brandner. Bodenheimer, P., Pollack, J. B. 1986. Calculations of the accretion and evolution of giant planets. The effects of solid cores. Icarus 67, 391-408. Boss, A. P. 1997. Giant planet formation by gravitational instability. Science 276, 1836- 1839. Boss, A. P. 1998. Evolution of the Solar Nebula. IV. Giant gaseous protoplanet formation. Astrophysical Journal 503, 923. Brown, T. M. 2008. Characterizing extrasolar planets. Extrasolar Planets 65.Cambridge University Press. Ed. H. J. Deeg, J. A. Belmonte & A. Aparicio. Caballero, J. A., Martín, E. L., Dobbie, P. D., Barrado Y Navascués, D. 2006. Are isolated planetary-mass objects real ly isolated?. A brown dwarf-exoplanet system candidate in the σ Orionis cluster. Astronomy and Astrophysics 460, 635-640. Cameron, A. G. W. 1978. Physics of the primitive solar accretion disk. Moon and Planets 18, 5-40. Chambers, J. E. 2001. Making more terrestrial planets Icarus, Volume 152, Issue 2, 205- 224. Chambers, J. E. 2006. Planet formation with migration. Astrophysical Journal 652, L133- L136. Clayton, D. D. 1968. Principles of stel lar evolution and nucleosynthesis. New York: McGraw-Hill, 1968 . Delfosse, X., Forveille, T., Mayor, M., Perrier, C. Naef, D., Queloz, D. 1998. The clo- sest extrasolar planet. A giant planet around the M4 dwarf GL 876. Astronomy and Astrophysics 338, L67-L70. Desch, S. J. 2007. Mass distribution and planet formation in the solar nebula. Astrophysical Journal 671, 878-893. Doyle, L. R. 2008. Overview of extrasolar planet detection methods. Extrasolar Planets 1. Cambridge University Press. Ed. H. J. Deeg, J. A. Belmonte & A. Aparicio. Dodson-Robinson, S. E., Willacy, K., Bodenheimer, P., Turner, N. J., Beichman, C. A. 2009. Ice lines, planetesimal composition and solid surface density in the solar nebula. Icarus, 200, 672-693. Dodson-Robinson, S. E., Bodenheimer, P., Laughlin, G., Willacy, K., Turner, N. J., Bei- chman, C. A. 2008. Saturn forms by core accretion in 3.4 Myr. Astrophysical Journal 688, L99-L102. Durisen, R. H., Boss, A. P., Mayer, L., Nelson, A. F., Quinn, T., Rice, W. K. M. 2007. Gravitational instabilities in gaseous protoplanetary disks and implications for giant planet formation. Protostars and Planets V 607-622. Fortier, A., Benvenuto, O. G., Brunini, A. 2007. Oligarchic planetesimal accretion and giant planet formation. Astronomy and Astrophysics 473, 311-322. Fortier, A., Benvenuto, O. G., Brunini, A. 2009. Oligarchic planetesimal accretion and giant planet formation II. Astronomy and Astrophysics 500, 1249-1252. Greenberg, R., Hartmann, W. K., Chapman, C. R., Wacker, J. F. 1978. Planetesimals to planets - Numerical simulation of col lisional evolution. Icarus 35, 1-26. Greenzweig, Y., Lissauer, J. J. 1992. Accretion rates of protoplanets. II - Gaussian distri- 146 butions of planetesimal velocities. Icarus 100, 440-463. Guillot, T., Burrows, A., Hubbard, W. B., Lunine, J. I., Saumon, D. 1996. Giant planets at smal l orbital distances. Astrophysical Journal 459, L35. Guillot, T. 1999. Interiors of giant planets inside and outside the Solar System.. Science 296, 72-77. Guillot, T. 2001. Physics of subestel lar objects: interiors, atmospheres, evolution. 31st Saas- Fee Advanced Course on Brown Dwarfs and Planets. Guillot, T. 2005. The interiors of giant planets: models and outstanding questions. Annual Review of Earth and Planetary Sciences 33, 493-530. Haisch, K. E., Jr., Lada, E. A., Lada, C. J. 2001. Disk frequencies and lifetimes in young clusters. Astrophysical Journal 553, L153-L156. Hayashi, C. 1981. Structure of the solar nebula, growth and decay of magnetic fields and ef- fects of magnetic and turbulent viscosities on the nebula. Progress of Theoretical Physics Supplement 70, 35-53. Henyey, L. G., Wilets, L., Böhm, K. H., Lelevier, R., Levee, R. D. 1959. A Method for Atomic Computation of Stel lar Evolution. Astrophysical Journal 129, 628-636. Henyey, L. G., Forbes, J. E., Gould, N. L. 1964. A New Method of Automatic Computation of Stel lar Evolution. Astrophysical Journal 139, 306-317. Hillenbrand, L. A. 2005. Observational constraints on dust disk lifetimes: implications for planet formation. ArXiv Astrophysics e-prints arXiv:astro-ph/0511083. Hoyle, F. 1960. The origin of the solar nebula. Quarterly Journal of the Royal Astronomical Society 1, 28. Hubickyj, O., Bodenheimer, P., Lissauer, J. J. 2005. Accretion of the gaseous envelope of Jupiter around a 5 10 Earth-mass core. Icarus 179, 415-431. Ida, S. 1990. Stirring and dynamical friction rates of planetesimals in the solar gravitational field. Icarus 88, 129-145. Ida, S., Makino, J. 1993. Scattering of planetesimals by a protoplanet - Slowing down of runaway growth. Icarus 106, 210. Ida, S., Lin, D. N. C. 2008. Toward a deterministic model of planetary formation. IV. Effects of Type I Migration. Astrophysical Journal 673, 487-501. Ida, S., Lin, D. N. C. 2008. Toward a deterministic model of planetary formation. V. Accumulation near the ice line and super-Earths. Astrophysical Journal 685, 584-595. Jayawardhana, R., Ivanov, V. D. 2006. Spectroscopy of young planetary mass candidates with disks. Astrophysical Journal Letters 647, L167-L170. Jewitt, D., Luu, J. 1992. Discovery of the candidate Kuiper belt object 1992 QB1. Nature 362, 730-732. Johansen, A., Oishi, J. S., Low, M.-M. M., Klahr, H., Henning, T., Youdin, A. 2007. Rapid planetesimal formation in turbulent circumstel lar disks. Nature 448, 1022-1025. Kasting, J. F., Whitmire, D. P., Reynolds, R. T. 1993. Habitable zones around main se- quence stars. Icarus 101, 108-128. Klahr, H. H., Bodenheimer, P. 2003. Turbulence in accretion disks: vorticity generation and angular momentum transport via the global baroclinic instability. Astrophysical Journal 582, 869-892. 147 Kippenhahn, R. Weigert, A., Hofmeister, E. 1967. Methods in computational physics. In- terscience. New York, 129. Kippenhahn, R., Weigert, A. 1990. Stel lar structure and evolution. Springer-Verlag Berlin Heidelberg New York. Also Astronomy and Astrophysics Library. Kokubo, E. 2001. Planetary accretion: from planetesimals to protoplanets. Reviews in Mo- dern Astronomy 14, 117. Kokubo, E., Ida, S. 1996. On runaway growth of planetesimals. Icarus 123, 180-191. Kokubo, E., Ida, S. 1998. Oligarchic growth of protoplanets. Icarus 131, 171-178. Kokubo, E., Ida, S. 2000. Formation of protoplanets from planetesimals in the solar nebula. Icarus 143, 15-27. Kokubo, E., Ida, S. 2002. Formation of protoplanet systems and diversity of planetary systems. Astrophysical Journal 581, 666-680. Lissauer, J. J. 1987. Timescales for planetary accretion and the structure of the protopla- netary disk. Icarus 69, 249-265. Lissauer, J. J., Stevenson, D. J. 2007. Formation of giant planets. Protostars and Planets V, 591-606. Marcy, G. W., Butler, R. P., Vogt, S. S., Fischer, D., Lissauer, J. J. 1998. A planetary companion to a nearby M4 dwarf, Gliese 876. Astrophysical Journal Letters, 505, L147- L149. Masset, F. S., D’Angelo, G., Kley, W. 2006. On the migration of protogiant solid cores. Astrophysical Journal 652, 730-745. Mayor, M., Queloz, D. 1995. A jupiter-mass companion to a solar-type star. Nature 378, 355. Militzer, B., Hubbard, W. B., Vorberger, J., Tamblyn, I., Bonev, S. A. 2008. A massive core in Jupiter predicted from first-principles simulations. Astrophysical Journal 688, L45-L48. Militzer, B., Hubbard, W. B. 2008. Comparison of Jupiter interior models derived from first-principles simulations. ArXiv e-prints arXiv:0807.4266. Millan-Gabet, R., Malbet, F., Akeson, R., Leinert, C., Monnier, J., Waters, R. 2007. The circumstel lar environments of young stars at AU scales. Protostars and Planets V, 539-554. Mizuno, H. 1980. Formation of the giant planets. Progress of Theoretical Physics 64, 544- 557. Mizuno, H., Nakazawa, K., Hayashi, C. 1978. Instability of a gaseous envelope surrounding a planetary core and formation of giant planets. Progress of Theoretical Physics 60, 699-710. Nagasawa, M., Lin, D. N. C., Thommes, E. 2005. Dynamical shake-up of planetary sys- tems. I. Embryo trapping and induced col lisions by the sweeping secular resonance and embryo-disk tidal interaction. The Astrophysical Journal, Volume 635, Issue 1, 578-598. Nagasawa, M., Thommes, E. W., Kenyon, S. J., Bromley, B. C., Lin, D. N. C. 2007. The diverse origins of terrestrial-planet systems. Protostars and Planets V, 639-654. Nettelmann, N., Holst, B., Kietzmann, A., French, M., Redmer, R., Blaschke, D. 2008. Ab initio equation of state data for hydrogen, helium, and water and the internal structure 148 of Jupiter. Astrophysical Journal 683, 1217-1228. Oppenheimer, B. R., Kulkarni, S. R., Matthews, K., Naka jima, T. 1995. Infrared spectrum of the cool brown dwarf GL:229B. Science 270, 1478. Papaloizou, J. C. B., Nelson, R. P., Kley, W., Masset, F. S., Artymowicz, P. 2007. Disk- planet interactions during planet formation. Protostars and Planets V, 655-668. Pečnik, B., Wuchterl, G. 2007. Protoplanetary dynamics - I. Dynamical modes of isothermal protoplanets. Monthly Notices of the Royal Astronomical Society 381, 640-646. Perri, F., Cameron, A. G. W. 1974. Hydrodynamic instability of the solar nebula in the presence of a planetary core. Icarus 22, 416-425. Podolak, M. 2003. The contribution of smal l grains to the opacity of protoplanetary at- mospheres. Icarus 165, 428-437. Podolak, M., Podolak, J. I., Marley, M. S. 2000. Further investigations of random models of Uranus and Neptune. Planetary and Space Science 48, 143-151. Podolak, M., Pollack, J. B., Reynolds, R. T. 1988. Interactions of planetesimals with pro- toplanetary atmospheres. Icarus 73, 163-179. Pollack, J. B., McKay, C. P., Christofferson, B. M. 1985. A calculation of the Rosseland mean opacity of dust grains in primordial solar system nebulae. Icarus 64, 471-492. Pravdo, S. H., Shaklan, S. B. 2009. An ultracool star’s candidate planet. Astrophysical Journal 700, 623-632. Press, W. H., Flannery, B. P., Teukolsky, S. A., & Vetterling, W. T. 1992, Numerical Recipes: the art of scientific computing. Cambridge University Pess, Third Edition, Cambridge. Rafikov, R. R. 2005. Can giant planets form by direct gravitational instability?. Astrophy- sical Journal 621, L69-L72. Rivera, E. J., Lissauer, J. J., Butler, R. P., Marcy, G. W., Vogt, S. S., Fischer, D. A., Brown, T. M., Laughlin, G., Henry, G. W. 2005. A ∼7.5 Earth-mass planet orbiting the nearby star, GJ 876. Bulletin of the American Astronomical Society 37, 1487. Rogers, F. J., Iglesias, C. A. 1992. Rosseland mean opacities for variable compositions. Astrophysical Journal 401, 361-366. Safronov, V.S. 1969, in Evolution of of the Protoplanetary Cloud and Formation of the Earth and Planets. Nauka, Moscow [Engl. tansl. NASA TTF-677, 1972] Sato, B. and 20 colleagues. 2005. The N2K Consortium. II. A transiting hot Saturn around HD 149026 with a large dense core, Astrophysical Journal 633, 465-473. Santos, N. C., Israelian, G., Mayor, M., Bento, J. P., Almeida, P. C., Sousa, S. G., Ecuvi- llon, A. 2005. Spectroscopic metal licities for planet-host stars: Extending the samples. Astronomy and Astrophysics 437, 1127-1133. Saumon, D., Chabrier, G., van Horn, H. M. 1995. An equation of state for low-mass stars and giant planets. Astrophysical Journal Supplement Series 99, 713. Saumon, D., Guillot, T. 2004. Shock compression of deuterium and the interiors of Jupiter and Saturn. Astrophysical Journal 609, 1170-1180. Selsis, F., Kasting, J. F., Levrard, B., Paillet, J., Ribas, I., Delfosse, X. 2007. Habitable planets around the star Gliese 581?. Astronomy and Astrophysics 476, 1373-1387. Shiraishi, M., Ida, S. 2008. Infal l of planetesimals onto growing giant planets: onset of 149 runaway gas accretion and metal licity of their gas envelopes. Astrophysical Journal 684, 1416-1426. Stevenson, D. J. 1982. Formation of the giant planets. Planetary and Space Science 30, 755-764. Thommes, E. W., Duncan, M. J., Levison, H. F. 2003. Oligarchic growth of giant planets. Icarus 161, 431-455. Thommes, E. W., Duncan, M. J. 2006. The accretion of giant-planet cores. Planet Forma- tion, 129. Thommes, E. W., Matsumura, S., Rasio, F. A. 2008. Gas disks to gas giants: simulating the birth of planetary systems. Science 321, 814. Tsiganis, K., Gomes, R., Morbidelli, A., Levison, H. F. 2005. Origin of the orbital archi- tecture of the giant planets of the Solar System. Nature 435, 459-461. Udry, S., and 10 colleagues 2007. The HARPS search for southern extra-solar planets. XI. Super-Earths (5 and 8 M⊕ ) in a 3-planet system. Astronomy and Astrophysics 469, L43-L47. Udry, S. 2008. Statistical properties of exoplanets. Extrasolar Planets, 24. Cambridge Uni- versity Press. Ed. H. J. Deeg, J. A. Belmonte & A. Aparicio. Ward, W. R. 1997. Protoplanet migration by nebula tides. Icarus 126, 261-281. Weidenschilling, S. J. 1977. The distribution of mass in the planetary system and solar nebula. Astrophysics and Space Science 51, 153-158. Weidenschilling, S. J., Marzari, F. 1996. Gravitational scattering as a possible origin for giant planets at smal l stel lar distances. Nature 384, 619-621 Wolszczan, A., Frail, D. A. 1992. A planetary system around the mil lisecond pulsar PSR1257 + 12. Nature 355, 145-147. Wuchterl, G. 1990. Hydrodynamics of giant planet formation. I - Overviewing the kappa- mechanism. Astronomy and Astrophysics 238, 83-94. Wuchterl, G. 1991. Hydrodynamics of giant planet formation III: Jupiter’s nucleated ins- tability. Icarus 91, 53-64. Wuchterl, G. 1995. Giant planet formation: A comparative view of gas-accretion. Earth Moon and Planets 67, 51-65. Wuchterl, G. 1996. Formation of giant planets close to stars. Bulletin of the American Astronomical Society 28, 1108. Wuchterl, G. 2008. From clouds to planet systems: formation and evolution of stars and planets. Extrasolar Planets 89. Cambridge University Press. Ed. H. J. Deeg, J. A. Bel- monte & A. Aparicio. Youdin, A. 2008. From grains to planetesimals: Les Houches Lecture. arXiv:0807.1114. Zapatero Osorio, M. R., Béjar, V. J. S., Martín, E. L., Rebolo, R., Barrado y Navascués, D. Mundt, R., Eislöffel, J., Caballero, J. A. 2002. A methane, isolated, planetary-mass object in Orion. Astrophysical Journal 578, 536-542. 150 Símbolos y unidades a = radio de la órbita del planeta entorno a la estrella central c = velocidad de la luz cs = velocidad del sonido CD = coeficiente (adimensional) de amortiguamiento de la componente gaseosa del disco CH4 = metano CO = monóxido de carbono d = distancia entre el observador y el objeto observado ǫac = luminosidad por acreción de planetesimales e = excentricidad de la órbita em = excentricidad de los planetesimales E = energía Ek,m = energía cinética de un planetesimal de masa m EOS= ecuación de estado (de sus siglas en inglés Equation Of State) φ = potencial gravitatorio G = Constante de Gravitación Universal (G = 6,67259 × 10−8 dyn cm2g−1) h = altura de escala del disco de planetesimales H = altura de escala de la componente gaseosa del disco protoplanetario H = átomo de hidrógeno H+ = ion positivo de hidrógeno H2 = molécula de hidrógeno He = átomo de helio He+ = átomo de helio una vez ionizado He++ = átomo de helio dos veces ionizado i = inclinación de los planetesimales respecto del plano fundamental Ji = momentos gravitatorios κ = opacidad media de Rosseland L = luminosidad LS = L⊙ = luminosidad solar (L⊙ = 3, 839 × 1033) erg/s) m = masa de un planetesimal Mc = masa del núcleo sólido del planeta Mcross = masa de cruce (Mcross ≡ Mc = Mg ) M∗ = masa de la estrella central MJ = masa de Júpiter (MJ = 1, 9 × 1030 g) 151 Mg = masa de la envoltura gaseosa Mr = masa contenida en una esfera de radio r M⊙ = masa del Sol (M⊙ = 1,989 × 1033g) M⊕ = masa de la Tierra (M⊕ = 5,97 × 1027g) Mp = masa del planeta ∇ = gradiente adimensional de temperatura (∇ ≡ d ln T d ln P ) ∇ad = gradiente adiabático ∇rad = gradiente radiativo NH3 = amoníaco NSMM= Nebulosa Solar de Masa Mínima Ωk = velocidad angular kepleriana pc= parsec (1 pc= 3 × 108 cm = 206.265 UA = 3,262 años luz) P = presión Pc = presión del gas de la envoltura en la superficie del núcleo PPT= Plasma Phase Transition. Región de la EOS del hidrógeno donde se produce la ionización por presión, la cual ocurre a través de una transición de fase de primer orden. r = coordenada radial rm = radio de un planetesimal de masa m Ra = radio de acreción (cid:16)Ra = GMr s (cid:17) c2 Rc = radio del núcleo sólido del planeta Rec = radio ecuatorial del planeta Reff = radio efectivo de captura del planeta 3M∗ (cid:17)1/3(cid:19) RH = radio de Hill (cid:18)RH = a (cid:16) Mp R∗ = radio de la estrella central Rp = radio del planeta R⊕ = radio de la Tierra (R⊕ = 6, 38 × 108 cm) R⊙ = radio del Sol (R⊙ = 6, 9 × 1010 cm) ρ = densidad volumétrica de gas de la nebulosa protoplanetaria ρc = densidad del gas de la envoltura en la superficie del núcleo ρg = densidad volumétrica de gas de la envoltura del protoplaneta ρm = densidad en masa de un planetesimal Σ = densidad superficial de sólidos en el disco S = entropía por unidad de masa Smix = entropía de mezcla SCVH= ecuación de estado de Saumon, Chabrier & Van Horn (1995) t = coordenada temporal tcross = tiempo de cruce T = temperatura Tc = temperatura del gas de la envoltura en la superficie del núcleo U = energía interna por unidad de masa UA= Unidad Astronómica (1 UA = 1, 495 × 1013 cm) v = velocidad 152 vesc = velocidad de escape vk = velocidad kepleriana de un objeto entorno al Sol en una órbita circular vrel = velocidad relativa entre el protoplaneta y los planetesimales X = fracción en masa de hidrógeno Y = fracción en masa de helio Z = fracción en masa de “elementos pesados” 153
1207.2127
1
1207
2012-07-09T19:01:01
Effect of the stellar spin history on the tidal evolution of close-in planets
[ "astro-ph.EP", "astro-ph.SR" ]
We investigate how the evolution of the stellar spin rate affects, and is affected by, planets in close orbits, via star-planet tidal interactions. To do this, we used a standard equilibrium tidal model to compute the orbital evolution of single planets orbiting both Sun-like stars and 0.1 M\odot M-dwarfs. We tested two stellar spin evolution profiles, one with fast initial rotation (P=1.2 day) and one with slow initial rotation (P=8 day). We tested the effect of varying the stellar and planetary dissipation and the planet's mass and initial orbital radius. Conclusions: Tidal evolution allows to differentiate the early behaviors of extremely close-in planets orbiting either a rapidly rotating star or a slowly rotating star. The early spin-up of the star allows the close-in planets around fast rotators to survive the early evolution. For planets around M-dwarfs, surviving the early evolution means surviving on Gyr timescales whereas for Sun-like stars the spin-down brings about late mergers of Jupiter planets. In light of this study, we can say that differentiating between one spin evolution from another given the present position of planets can be very tricky. Unless we can observe some markers of former evolution it is nearly impossible to distinguish the two very different spin profiles, let alone intermediate spin profiles. Though some conclusions can still be drawn from statistical distributions of planets around fully convective M-dwarfs. However, if the tidal evolution brings about a merger late in its history it can also entail a noticeable acceleration of the star in late ages, so that it is possible to have old stars that spin rapidly. This raises the question of better constraining the age of stars.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. spintidesMsarxiv September 28, 2018 c(cid:13) ESO 2018 2 1 0 2 l u J 9 . ] P E h p - o r t s a [ 1 v 7 2 1 2 . 7 0 2 1 : v i X r a Effect of the stellar spin history on the tidal evolution of close-in planets Emeline Bolmont1,2, Sean N. Raymond1,2, Jeremy Leconte3, and Sean P. Matt4,5 1 Univ. Bordeaux, LAB, UMR 5804, F-33270, Floirac, France 2 CNRS, LAB, UMR 5804, F-33270, Floirac, France 3 Laboratoire de M´et´eorologie Dynamique, Institut Pierre Simon Laplace, Paris, France 4 Laboratoire AIM Paris-Saclay, CEA/Irfu Universit´e Paris-Diderot CNRS/INSU, 91191 Gif-sur-Yvette, France 5 NASA Ames Research Center, M.S. 245-6, Moffett Field, CA 94035-1000, USA Received xxx ; accepted xxx ABSTRACT Context. The spin rate of stars evolves substantially during their lifetime, due to the evolution of their internal structure and to external torques arising from the interaction of stars with their environments and stellar winds. Aims. We investigate how the evolution of the stellar spin rate affects, and is affected by, planets in close orbits, via star-planet tidal interactions. Methods. We used a standard equilibrium tidal model to compute the orbital evolution of single planets orbiting both Sun-like stars and very low-mass stars (0.1 M(cid:12)). We tested two stellar spin evolution profiles, one with fast initial rotation (1.2 day rotation period) and one with slow initial rotation (8 day period). We tested the effect of varying the stellar and planetary dissipation and the planet's mass and initial orbital radius. Results. For Sun-like stars the different tidal evolution between initially rapidly and slowly rotating stars is only evident for extremely close-in gas giants orbiting highly dissipative stars. However, for very low mass stars the effect of initial rotation of the star on the planet's evolution is apparent for less massive (1M⊕) planets and for typical dissipation values. We also find that planetary evolution can have significant effects on the stellar spin history. In particular, when a planet falls on the star it makes the star spin up. Conclusions. Tidal evolution allows to differentiate the early behaviors of extremely close-in planets orbiting either a rapidly rotating star or a slowly rotating star. The early spin-up of the star allows the close-in planets around fast rotators to survive the early evolution. For planets around M-dwarfs, surviving the early evolution means surviving on Gyr timescales whereas for Sun-like stars the spin- down brings about late mergers of Jupiter planets. In light of this study, we can say that differentiating between one spin evolution from another given the present position of planets can be very tricky. Unless we can observe some markers of former evolution it is nearly impossible to distinguish the two very different spin profiles, let alone intermediate spin profiles. Though some conclusions can still be drawn from statistical distributions of planets around fully convective M-dwarfs . However, if the tidal evolution brings about a merger late in its history it can also entail a noticeable acceleration of the star in late ages, so that it is possible to have old stars that spin rapidly. This raises the question of better constraining the age of stars. Key words. Stars: rotation -- Planets and satellites: dynamical evolution and stability -- Planet-star interactions -- 1. Introduction The spin rate is an important quantity for the evolution of a star and also for the evolution of any planets orbiting close-in. The parameter that governs the direction of tidal evolution for a planet orbiting a star (or a satellite orbiting a planet) is the initial semi-major axis with respect to the corotation radius, the orbital radius where the orbital period matches the central body's spin period. For a planet interior to the corotation radius the planet's mean motion is faster than the primary's rotation, so the tidal bulge raised by the planet on the primary lags behind the position of the planet. The planet feels a drag force that slows it down and causes its orbital radius to shrink, in some cases leading to an eventual merger with the primary. However, for a planet exterior to the corotation radius, the tidal bulge on the star is in advance with respect to the position of the planet and tidal forces push the planet outward. The rotational evolution of Sun-like stars can be described in 3 main stages: the pre-main sequence (PMS) stage, the zero-age main sequence (ZAMS) approach and the main sequence (MS) relaxation (Bouvier, 2008). During the PMS stage the young stars are observed to have a range of spin periods, typically from a few to ∼ 10 days, and there is evidence that a highly efficient braking mechanism is at work (Herbst et al., 2007). It is still not clear what mechanisms are responsible for the observed distri- bution and the angular momentum loss, but these may be due to the interaction between the star and surrounding accretion disk (e.g., Ghosh & Lamb, 1978; Shu et al., 1994; Matt & Pudritz, 2005b; Matt et al., 2010), powerful stellar winds (Hartmann & MacGregor, 1982; Hartmann & Stauffer, 1989; Tout & Pringle, 1992; Paatz & Camenzind, 1996; Matt & Pudritz, 2005a, 2008; Matt et al., 2012), or other processes. Toward the end of the PMS phase, the fastest stars in the observed distributions appear to spin up in a way consistent with angular momentum conser- vation, while the rotation rates of the slowest rotators does not appear to change significantly. Thus, near the ZAMS, the spin period distributions are the widest, typically ranging from a few hours to ∼ 10 days (e.g., Bouvier et al., 1997; Bouvier, 2008). Once on the main sequence, the stellar structure evolves slowly 1 Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets enough that the torque from ordinary stellar winds becomes im- portant. Thus, on gigayear timescales, the average spin rates de- crease (Skumanich, 1972), and the range of observed spin rates narrows. To bracket the range of observed stellar spin rates, we consider here two populations: initially fast rotators, whose evo- lution follows the upper envelope of the observed spin rate dis- tributions, and slow rotators that follow the lower envelope. During the PMS and approach to ZAMS, the observed spin period distributions of M-dwarfs is qualitatively similar to that of Sun-like stars. Observations of young clusters constrain the ro- tation period of low-mass stars younger than a few ×100 Myr (Stassun et al., 1999; Herbst et al., 2001; Irwin et al., 2008) but that approach fails for old clusters due to the faintness of old M-dwarfs. Nonetheless, old slowly-rotating M-dwarfs have been detected (Benedict et al., 1998; Kiraga & Stepien, 2007; Charbonneau et al., 2009). Contrary to Sun-like stars that are mostly radiative except for a small (in terms of mass) convec- tive region at the surface, very low mass stars (M∗ < 0.35M(cid:12)) are entirely convective (Chabrier & Baraffe, 1997). For Sun- like stars with a radiative core, the interface between the core and convective envelope is thought to be important for the mag- netic dynamo, whereas in fully convective low mass stars other mechanisms have to be invoked to explain their observed mag- netic activity (Reiners & Basri, 2007). For example, Chabrier & Kuker (2006) showed that mean field modeling can produce a α2 dynamo, which creates large-scale nonaxisymmetric fields and Browning (2008) showed that three-dimensional nonlinear magnetohydrodynamic simulations of the interiors of fully con- vective M-dwarfs can also produce a large-scale dynamo. The coupling between stellar spin history and tidal evolu- tion has been studied by Zahn (1994) for close binaries and by Dobbs-Dixon et al. (2004) for short period planets. Individual systems where tidal interactions are thought to have played a role have also been the subject of various studies. Lin et al. (1996) proposed that the planet orbiting 51 Peg stopped its disk- induced inward migration because of its presence outside coro- tation before the disk dispersal. Individual systems of the OGLE survey have been studied by Patzold et al. (2004). Some studies give constraints for stellar dissipation, such as Carone & Patzold (2007) for OGLE-TR-56b and by Lanza et al. (2011) for the CoRoT-11 system. In this study, we try to have a more general and system- atic approach of the effect of the stellar spin evolution on the tidal evolution of close-in planets. To this end, we couple stellar evolutionary models (Baraffe et al., 1998), wind parametrization (Bouvier, 2008; Irwin et al., 2011) and tidal evolution. We con- sider two limiting cases for the stellar spin evolution that corre- spond to a star whose initial rotation is either very fast or very slow. These different evolutionary paths can be seen in Figure 1 of Bouvier (2008) for Sun-like stars, or in Figures 13 to 15 of Irwin et al. (2011) for M-dwarfs . The slowly rotating stars be- gin with a rotation period of 8 days and the fast rotating stars with a period of 1.2 days. Both fast and slow rotators evolve as explained in Bouvier et al. (1997), where the loss of angular momentum due to the stellar wind is quantified given different star-dependent parameters among which is the rotation rate of the star. The higher the spin of a star, the more active the star, the stronger the winds and the stronger the braking. Here we use a standard equilibrium model to study the tidal evolution of planets orbiting stars. Our paper is structured as follows. The tidal and star evolutionary models are briefly dis- cussed in Section 2. Some preliminary analysis based on order of magnitude study are made in Section 3 before giving the results of the tidal evolution of planets around the two types of stars 2 considered here in Section 4. Finally, in Section 5 we discuss the difficulty of linking the results of tidal evolution and observa- tions and the important effect of late mergers on the rotation rate of the stars. 2. Model description We have developed a model to study the orbital evolution of planets around stars by solving the tidal equations for arbitrary eccentricity and also taking into account the observed spin evo- lution of stars. 2.1. Tidal model The tidal model that we used is a re-derivation of the equilib- rium tide model of Hut (1981) as in Eggleton et al. (1998). We consider both the tide raised by the star on the planet and by the planet on the star. We use the constant time lag model (Leconte et al., 2010) and the internal dissipation constant σ that was cali- brated for giant exoplanets and their host stars by Hansen (2010). Taking into account both stellar tide and planetary tide, the secular tidal evolution of the semi-major axis a is given by (Hansen, 2010): (cid:104) Na1(e) − Ωp (cid:104) n Na1(e) − Ω∗ n (cid:105) (cid:105) Na2(e) Na2(e) , 1 a da dt = − 1 Tp − 1 T∗ (1) (2) where the dissipation timescale T∗ is defined as T∗ = 1 9 M∗ Mp(Mp + M∗) a8 R10∗ 1 σ∗ and depends on the stellar mass M∗, its dissipation σ∗ and the planet mass Mp. Ωp is the planet's rotation frequency, and n is the mean orbital angular frequency. The planet parameters are obtained by switching the p and ∗ indices. Na1(e) and Na2(e) are eccentricity-dependent factors, which are valid even for very high eccentricity (Hut, 1981): Na1(e) = Na2(e) = 1 + 31/2e2 + 255/8e4 + 185/16e6 + 85/64e8 (1 − e2)15/2 1 + 15/2e2 + 45/8e4 + 5/16e6 (1 − e2)6 . , The secular tidal equations can be extended to arbitrary obliquity, which has been done in Leconte et al. (2010). The equations for the eccentricity and planetary rotation rate can be found in Bolmont et al. (2011), where tidal evolution was studied for a planet-brown dwarf system. 2.2. Planets model We consider planets with a wide mass range, from 1 M⊕ to 5 MJ (MJ = mass of Jupiter). For terrestrial planets we use the dissipation values based on Earth's dissipation value. Neron de Surgy & Laskar (1997) in- ferred the quantity k2,⊕∆T⊕ = 213 s from the DE245 data for Earth. k2,p is the Earth's potential Love number of degree 2, which is a parameter depending on the moment of inertia of the body. It tells us how the body responds to compression (k2 = 3/2 means that the body is an incompressible ideal fluid planet). ∆Tp Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets is the constant time-lag. Hansen's σp and the quantity k2,p∆Tp are related through: R5 pσp G , 3 2 k2,p∆Tp = (3) where Rp is the planetary radius and G is the gravitational The terrestrial planet's compositions are assumed to be an Earth-like mixture of rock and iron, following the mass-radius relation of Fortney et al. (2007). constant. For 1 MJ, 2 MJ and 5 MJ planets, we use respectively the values of radius and mass of Jupiter, WASP-33b (Christian et al., 2006; Collier Cameron et al., 2010) and OGLE2-TR-L9 (Snellen et al., 2009). For the dissipation factor, we use Hansen (2010)'s estimates for gas giants : σp = 2.006 × 10−60 g−1cm−2s−1. Because the planetary spin synchronization timescale is short compared to the other timescales considered here (Leconte et al., 2010; Heller et al., 2011), the planet rotation period is fixed to the pseudo-synchronization value at every calculation timestep. In this study, we chose not to treat the evolution of the obliquity of the bodies for simplicity. 2.3. Stellar evolution Sun-like stars and M-dwarfs are self-gravitating objects born from a collapsing dense molecular cloud. In the beginning of their evolution they contract, and when the temperature at their core reaches the value of ∼ 3 × 106 K, the PPI nuclear reaction starts and the stars enter the main sequence. The Sun is thought to have entered the MS around a few 10 Myr after its birth. After 8 Gyr, the Sun will leave the MS to evolve towards a red giant. To compute the influence of dissipation into a star, one needs to know the internal structure, mainly the evolution of the ra- dius with time, the moment of inertia (through the gyration ra- dius, which is here considered constant) and the tidal dissipation factor σ∗. The first two quantities are provided by stars evolu- tionary models (Chabrier & Baraffe, 1997; Baraffe et al., 1998). However evolution models use unconstrained values for the ra- dius of the star at early ages (t (cid:46) 106yr), so some quantitative un- certainties can arise early in the stellar evolution (Baraffe et al., 2002). However, here we consider evolution after t0 = 5 Myr for Sun-like stars and after t0 = 8 Myr for M-dwarfs , so these uncertainties should remain negligible. We assume that both Sun-like stars rotate as solid bodies, as in Bouvier et al. (1997), although more recent work has included the effect of internal differential rotation between the radiative core and the convective envelope (Bouvier, 2008). Given that low mass stars are believed to undergo the same kind of stel- lar wind braking as Sun-like stars (Irwin et al., 2011), we also consider solid rotation (Morin et al., 2008). In this work, M∗ is held constant, and the effect of mass loss (through processes like stellar winds) on the internal structure of the star is considered negligible. use a dissipation factor of σ∗,dM = 2.006× 10−60 g−1cm−2s−1 for the 0.1M(cid:12) stars in our calculations (Hansen, 2010). No scaling of the M-dwarfs dissipation factor has been performed compared to the brown dwarfs value. In this work, we varied the stellar dissi- pation factor by a few orders of magnitude so that the real value for M-dwarfs is in the considered range. This dissipation is much larger than for a Sun-mass star, and as we will see in subsection 3.2, this has implications for tidal evolution. Stars more massive than 0.35M(cid:12) that have radiative zones may be less dissipative and more similar to a Sun-like star. 2.3.2. Rotational angular velocity Our calculations begin during the stellar PMS. The evolution of the observed spin distributions of PMS stars have often been pa- rameterized in terms of a "disk locking" scenario (e.g., Bouvier et al., 1997; Rebull et al., 2004, 2006; Edwards et al., 1993; Choi & Herbst, 1996), in which the spin period of the star is assumed to remain constant at a specified "initial" rate, for some specified amount of time (hypothesized to be associated with the dissipa- tion of the disk). For simplicity, we adopt this basic picture and start our calculations at the point of "disk dispersal," after which the stellar angular momentum evolution is computed according to physical equations. A primary goal of the present work is to determine how dif- ferent stellar spin histories influence the star-planet tidal interac- tion. To this end, we consider two different spin evolution tracks that approximately follow the fast and slow envelopes of the ob- served stellar spin distributions, and we do this for two different stellar masses, 0.1 and 1M(cid:12). In order to describe these tracks in a physically self-consistent way, we adopt a simplified model for the stellar spin, adapted from Bouvier et al. (1997) for solar mass stars and from Irwin et al. (2011) for 0.1M(cid:12) stars. For both 0.1 and 1M(cid:12) stars we add the effect of tides to the formula of Bouvier et al. (1997) for the loss of angular momen- tum due to the stellar winds (based on the formulae of Kawaler (1988) and MacGregor & Brenner (1991)). The expression for the angular momentum loss rate is: (cid:32) R∗ (cid:33)1/2(cid:32) M∗ R(cid:12) M(cid:12) (cid:33)−1/2 (cid:35) (4) (5) 1 J dJ dt = + sat −1 J 1 J (cid:34) KΩα∗ ω3−α h 2T∗ No1(e) − Ω∗ n No2(e) , where h is the orbital angular momentum, n is the mean orbital angular frequency, T∗ is the stellar dissipation timescale, and the functions No1 and No2 are defined as: No1(e) = No2(e) = 1 + 15/2e2 + 45/8e4 + 5/16e6 (1 − e2)13/2 , 1 + 3e2 + 3/8e4 (1 − e2)5 . 2.3.1. Stellar dissipation The stellar dissipation factor is poorly constrained but Hansen (2010) gives estimates for Sun-like stars : σ∗ = 6.4 × 10−59 g−1cm−2s−1σ∗ where σ∗ = 7.8 × 10−8 . Thus for Sun-like stars, the dissipation factor is σ∗ = 4.992 × 10−66 g−1cm−2s−1. Like a brown dwarf, a 0.1M(cid:12) star is fully convective so we expect that the dissipation mechanisms within the 0.1M(cid:12) star should be closer to a brown dwarf than a Sun-like star. Thus, we Here K, and ωsat are parameters of the model from Bouvier et al. (1997). To reproduce the present rotation of the Sun, we use the value of K = 1.6 × 1047 cgs and ωsat = 14 Ω(cid:12). Bouvier et al. (1997) showed that for fast rotators (Ω∗ > ωsat), α = 1 and for slow rotators (Ω∗ < ωsat), α = 3. Irwin et al. (2011) proposed various parameters to reproduce the observational data for stars with masses 0.1 < M/M(cid:12) ≤ 0.35, and the spin evolution they simulated was calculated for a star of mass 0.25M(cid:12). They also noted that the M-dwarf fast and slow rotators could not be fit with a single value of the parameter 3 Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets K. We consider here that the spin evolution of 0.1M(cid:12) stars can be described with the same parameters as 0.25M(cid:12) stars, and we note that this assumption can explain the differences between the curves we show below and the curves of Irwin et al. (2011). In this work, we used the values: ωsat = 0.65 Ω(cid:12), K f ast = 2.03 × 1045 cgs, Kslow = 8.0 × 1045 cgs, for initially fast rotators for initially slow rotators These values allows us to reproduce the values of the spin of the star at t = 5 Gyr in the two extreme trends seen in Fig. 14 of Irwin et al. (2011). Fig. 1. Radius (top panel) and spin (bottom panel) evolution of a 0.1M(cid:12) and a 1M(cid:12) star. In the bottom panel, the full and dashed dotted blue lines represent respectively the evolution of the rotation period of an initially fast rotating star and an initially slow rotating star with no planet. As in the top panel, the blue curves correspond to the 0.1M(cid:12) star and the black curves to the Solar-mass star. The blue diamonds in the bottom panel correspond to the values of the spin at 5 Gyr for the two extreme trends of Irwin et al. (2011). The vertical dashed lines represent t0 for the two stellar masses. Fig. 1 shows 0.1 and 1M(cid:12) star radius evolution track as well as the different spin period profiles. For both stellar masses, the slow rotators have an initial period of P∗0 = 8 days and fast rotators have an initial period of P∗0 = 1.2 days, as in Bouvier (2008). Finally, the expression for the stellar rotation is : (cid:33) Ω∗(t) = Ω∗(t0) × rg2∗(t0) rg2∗(t) (cid:32)(cid:90) t × exp ftidesdt t0 (cid:32) R∗(t0) (cid:33)2 (cid:32)(cid:90) t R∗(t) × exp t0 (cid:33) , fwinddt (6) (7) 4 where t0 corresponds to the time of disk dispersal, and ftides is given by : (cid:12)(cid:12)(cid:12)(cid:12)R∗=cst,rg2∗=cst dΩ∗ dt No1(e) − Ω∗ n (cid:104) 1 Ω∗ γ∗ 2T∗ ftides = = (cid:105) . (8) No2(e) Here, γ∗ = h I∗Ω∗ is the ratio of orbital angular momentum h to spin angular momentum and rg2∗ is the square of the parameter rg∗ (which is the radius of gyration of Hut (1981)) which is de- fined as : I = M∗(rg∗R∗)2, where I is the moment of inertia of the star. fwind is given by : − 1 I∗ KΩ2∗ − 1 I∗ Kω2 sat (cid:17)−1/2 (cid:17)1/2(cid:16) M∗ (cid:16) R∗ (cid:17)1/2(cid:16) M∗ (cid:16) R∗ (cid:17)−1/2 M(cid:12) R(cid:12) R(cid:12) M(cid:12) fwind = , if Ω∗ < ωsat if Ω∗ > ωsat , (9) The integration of the equations of Section 2.1 was per- formed using a fourth order Runge-Kutta integrator with an adaptive timestep routine (Press et al., 1992). The precision of the calculations was chosen such that the final semi-major axis of each integrated system was robust to numerical error at a level of at most one part in 103. Fig. 1 shows that after the time of disk dispersal τdisk (long dashed vertical lines in Fig. 1), the star spins up due to contrac- tion. After a few ×108 yrs, stellar winds start to efficiently spin down the star. 3. Order of magnitude analysis 3.1. Parameters space We present results for 0.1 and 1M(cid:12) stars1. We investigate the evolution of planets with masses of 1M⊕, 10M⊕, 1MUranus(= 14.5M⊕), 1MJ, 2MJ and 5MJ. For the moment we assume zero initial eccentricity so that only the stellar tide governs the evolu- tion. In Subsection 4.3.2 we present the outcome of an extreme case with an initial eccentricity of 0.01. We consider that planets begin their tidal evolution at the time of disk dispersal τdisk. The planet formation timescale is proportional to the orbital frequency (Safronov, 1969; Raymond et al., 2007) and is thus far shorter than the disk lifetime for close-in orbits, but planet-disk interactions probably overwhelm planet-star tidal interactions during this time. For Sun-like stars, this time is taken to be : τdisk = 5 × 106 yrs, which then cor- responds to our initial time t0, and for M-dwarfs , t0 = τdisk = 8 × 106 yrs. We tested values for the stellar dissipation σ∗ between 1 and 1000 times the mean values given by Hansen (2010). 3.2. Order of magnitude and timescales The stellar dissipation timescale is given by equation 2. This timescale depends on the stellar mass M∗, radius R∗, and tidal dissipation factor σ∗, and the semi-major axis a of the planet. As R∗ shrinks with time, it is clear that for fixed a the stellar dissipation timescale increases with time, so the stellar tide be- comes weaker. Assuming Mp (cid:28) M∗, the ratio of the dissipation timescale of Sun-like stars (T(cid:12)) to M-dwarfs (TdM) is: 1 Some simulations were performed with a stellar mass of 0.8M(cid:12), but the results were very similar to what we obtained with the 1M(cid:12) star, so we do not discuss those results. Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets R(cid:12) which gives the following values: σdM σ(cid:12) , (10) (cid:33)10 T(cid:12) TdM (cid:32) RdM (cid:40)T(cid:12) ≈ 10 TdM, = T(cid:12) ≈ 1 1000 TdM, at 8 Myr at 5 Gyr. We can see that comparing the magnitude of the tidal effect around a M-dwarf and a Sun-like star is equivalent to compare their radii and their dissipation factors. At all times, Sun-like stars have a bigger radius than M-dwarfs but have a smaller dis- sipation factor. At 8 Myr, the dissipation timescale of M-dwarfs is shorter than of Sun-like stars so the early tidal evolution will be more important around M-dwarfs than around Sun-like stars. However, at 5 Gyr, the dissipation timescale of M-dwarfs is much larger than of Sun-like stars so the late tidal evolution will be more important around Sun-like stars than around M-dwarfs. Fig. 2. Stellar dissipation timescale T∗ and SMA-evolution timescale τa versus semi-major axis for stars of mass 0.1M(cid:12) and 1M(cid:12) and a planet of Jupiter mass. The stellar dissipation timescale was calculated for mean dissipation factors and for two different star ages: 8 Myr and 5 Gyr. The difference between these two times is a difference in the radius of the star, which is smaller at 5 Gyr than at 8 Myr. Another important timescale is the semi-major axis evolution timescale (SMA-evolution timescale). Assuming zero orbital ec- centricity, it is given by: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) a a (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:32) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T∗ 1 − Ω∗ n (cid:33)−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . τa = (11) In contrast with the stellar dissipation timescale T∗, τa is af- fected by the spin of the star. T∗ is the limit of τa when Ω∗ → 0. Figure 2 shows the dependance of the stellar dissipation timescale T∗ and the semi-major axis evolution timescale τa on the semi-major axis a for a system with a Jupiter mass planet, for M∗ = 0.1M(cid:12) and 1M(cid:12). As expected, the farther a planet the weaker the stellar tide. A Jupiter at a > 0.05 AU around a 8 Myr star of mass 0.1M(cid:12) does not experience any noticeable semi- major axis evolution over 10 Gyr. Fig. 2 also shows that between 8 Myr and 5 Gyr the stellar dissipation timescale increases. At 8 Myr, T∗ for a 0.1M(cid:12) star is shorter than for a 1M(cid:12) star so the stellar tide will have a stronger effect on the planet orbiting an M-dwarf than a Sun-like star. However, at 5 Gyr the stellar dissi- pation timescale of the M-dwarf is longer than 10 Gyr for planets at a > 0.007 AU. This means that for M-dwarfs the system will "freeze" after some time such that the interesting tidal evolution will only occur early in the evolution. For Sun-like stars, the in- crease of the stellar dissipation timescale is less pronounced so tides still matter for planets closer than ∼ 0.018 AU at 5 Gyr. Concerning the evolution of τa the trends are similar to those for T∗ but a few differences can be seen. In particular, the curves show a peak feature at a semi-major axis corresponding to the corotation radius. This is because τa diverges when the semi- major axis is close to the corotation radius: if a planet forms precisely at the corotation radius of a non-evolving star, it will experience no tidal migration. The system is thus perfectly syn- chronized as the planet - which is already in synchronization - and the star always show each other the same sides. However, the corotation radius is an unstable equilibrium distance because inside the corotation radius - to the left of the peak - the planets migrate inwards and outside the corotation radius - to the right of the peak- the planets migrate outwards. For a system with an evolving star, the stellar tide always causes tidal migration. Fig. 2 also shows that outward migration (taking place further away from the star) always occurs on longer timescales than inward migration (taking place closer in). Fig. 2 also shows that Jupiter-mass planets closer than 0.02 AU around Sun-like stars with the mean dissipation value tidally migrate inward. Planets farther than 0.02 AU experience no noticeable semi-major axis change because τa is so long. In section 4, we will show that Sun-like stars need a dissipation factor of 1000 × σ∗ in order to have noticeable tidally-induced changes within 10 Gyr. In Fig 2, this is equivalent to lowering the curves corresponding to 1M(cid:12) stars, effectively decreasing the different timescales so that interesting behavior can be observed within the 10 Gyr timescale we consider here. We can also consider the timescale of stellar spin evolution τΩ∗ (again for zero eccentricity): (cid:32) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Ω∗ Ω∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2T∗ γ∗ 1 − Ω∗ n (cid:33)−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) , τΩ∗ = (12) where γ∗ = h I∗Ω∗ is the ratio of orbital angular momentum h to the spin angular momentum and I∗ is the moment of inertia of the star. The ratio of the stellar spin synchronization timescale of Sun-like stars to M-dwarfs assuming Mp (cid:28) M∗ is: 1 − ΩdM/n 1 − Ω(cid:12)/n , (cid:32) RdM rg2(cid:12) rg2dM σdM σ(cid:12) τΩ(cid:12) τΩdM (cid:33)8 (13) R(cid:12) = where rg2i is the square of the parameter rgi (which is the radius of gyration of Hut (1981)). This timescale depends on the moment of inertia of the star. If the star has a high inertia, the tidal forces need more time to bring the star into synchronization with the orbital frequency. Fig. 3 shows that the stellar spin evolution timescale is shorter for M-dwarfs than for Sun-like stars at 8 Myr, and it is longer for M-dwarfs than for Sun-like stars at 5 Gyr for close-in planets. Due to the strong dependence of tidal effect with respect to the orbital distance, the acceleration of the spin due to a planet inside the corotation radius is faster than its deceleration due to a planet outside the corotation radius. The spin-evolution timescale is short for close-in planets, when the planets fall towards the star 5 Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets before its first crossing with the corotation radius and an outward migration after. For initial semi-major axes larger than 0.02 AU, the difference between the two spin profiles is negligible. One super Earth, GJ 1214 b, has been detected around a 0.16M(cid:12) star (Charbonneau et al., 2009), so this planet could be experiencing interesting tidal evolution. Fig. 3. Spin-evolution timescale τΩ versus SMA for stars of mass 0.1M(cid:12) and 1M(cid:12) and a Jupiter-mass planet. The spin-evolution timescale was calculated for mean dissipation factors and for stellar ages of 8 Myr and 5 Gyr. there will be an angular momentum transfer between the planets orbit and the spin of the star leading to a noticeable spin-up of the star. Stars generally have enough inertia that tidal interactions with a planet require longer than 10 Gyr to bring the star in synchronization. However, for a very dissipative M-dwarf and a Jupiter mass planet the stellar synchronization timescale is short enough to lead to a perfect synchronized system (as we will dis- cuss in Section 4.3.2). The timescale of planet spin evolution is usually much shorter than the stellar spin evolution timescale. That is why in this work we consider the planet is always in pseudo- synchronization -if the eccentricity is non zero- or in synchro- nization - if the eccentricity is zero. 4. Results 4.1. Orbital evolution of planets For 0.1M(cid:12) stars, the effect of different spin profiles on the or- bital evolution of planets is apparent for Earth mass planets and mean dissipations values. For a 1M⊕ planet beginning at an ini- tial semi-major axis of 9 × 10−3 AU, slow rotators systems will tend to make the planet fall on the star whereas fast rotators sys- tems will allow the planet to survive for 10 Gyr. Interesting effects start to occur for planets of mass higher than 10M⊕. Equation 2 shows that the stellar dissipation timescale also depends on the mass of the planet. The more mas- sive the planet, the shorter the dissipation timescale. For 1M⊕ planets, evolution timescales are too big to lead to significant changes in less than a few Gyrs. However for planets of mass higher than 10M⊕, the evolution timescales are compatible with visible changes over a few 107 yrs. These planets begin to react as they cross the shrinking or expanding corotation radius. Fig. 4 shows the results of simulations for 10M⊕ planets orbiting a 0.1M(cid:12) star, using mean dissipation values. A planet beginning at 9 × 10−3 AU falls on the star in a few hundreds of thousands years if the star is a slow rotator, but survives if the star is a fast rotator. This latter planet experiences a small inward migration 6 Fig. 4. Tidal evolution of 10M⊕ mass planets starting at different ini- tial semi-major axes around either a fast rotating or a slow rotating 0.1M(cid:12) star with mean dissipation factor. Top panel: evolution of the semi-major axis. The full colored lines correspond to fast rotating stars and the dash-dotted lines correspond to slowly rotating stars. The solid and dash-dotted red lines represent the evolution of the corotation ra- dius in both cases if there is no planet. The long black dashes represent the Roche limit. Middle panel: The corresponding stellar rotation evo- lution (the same line code is used). The black long dashes represent Tsat = 2π/ωsat. Bottom panel : Equatorial velocity of the star vs time (the same line code is used). This effect is more dramatic if the stellar dissipation in- creased to 1000×σ∗. The stellar tide contributes in pushing away the planets orbiting a fast rotating star very efficiently and we can find the same behavior as for massive brown dwarfs (Bolmont et al., 2011). For high dissipations and fast rotators, the planets are pushed farther away and for the inner most planets converge towards a given distance. If we considered an equally spaced dis- tribution of planets, the distribution would be more packed at the end of the evolution. As we saw in Subsection 3.2, the tidally-induced early semi- major axis evolution occurs on shorter timescale around a 0.1M(cid:12) star than around a 1M(cid:12) star. So to be able to observe any discrep- ancies between spin profiles for Sun-like stars the stellar dissi- pation factor must be larger than the mean value. Hereafter, for Sun-like stars we use a dissipation factor of a thousand times Hansen (2010)'s mean value. Fig. 5 shows the tidal evolution of Jupiter-mass planets or- biting 1M(cid:12) stars. Differences can only be seen between fast and slow rotators for very close-in planets. For slow rotators, a planet Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets Fig. 5. Tidal evolution of Jupiter mass planets of different initial orbital distance around either a fast rotating or a slow rotating 1M(cid:12) star with a large dissipation factor. Top panel: the colored lines represent the semi- major axis of the planets. Solid lines correspond to an initially fast ro- tating star and dashed dotted lines to an initially slow-rotating star. The red curves represent the corotation radius assuming no planet. Middle panel: evolution of the stellar rotation period. Bottom panel: Evolution of the equatorial velocity of the star. In middle and bottom panel, a red full circle represents the Sun's present rotation, and the red curves cor- respond to stars with no planets. beginning at 0.03 AU falls on the star in ∼ 106 yrs, whereas for the fast rotators it falls in (cid:46) 109 yrs. The evolution of planets around fast rotators is interesting because planets beginning very close to the corotation radius experience an outward migration for a few ×107 yrs. Around t = 50 Myrs, the star has spun down sufficiently due to stellar winds that these planets cross the outward-drifting corotation ra- dius. After the crossing, these planets migrate back inward and crash onto the star at about1 Gyr. For planets beginning their evolution past 0.05 AU, the difference between the two spin pro- files is negligible. Planets beginning their evolution at a semi- major axis bigger than 0.08 AU survive the evolution on 10 Gyr timescales. 4.2. Orbital evolution of planets compared with the age of the star The time on the x-axis in Figs. 4 and 5 starts at t0, thus after 5-8 Myr of stellar evolution, but the stellar evolution timescale may offer a more appropriate measure. Figure 6 shows the orbital evolution of a 10M⊕ planet around a 0.1M(cid:12) star. We can see that the planets which do not survive crash on the star on a timescale much shorter than the age of the star. Planets orbiting 0.1M(cid:12) stars undergo similar evolution to planets orbiting brown dwarfs (see Bolmont et al., 2011). Most of the tidal evolution occurs at early times, when the radius of the stars/brown dwarfs is still large enough that the stellar/brown dwarf tide is strong enough to drive changes in the planets' semi- Fig. 6. Tidal evolution of a 10M⊕ mass planet starting at different initial semi-major axis around either a fast rotating or a slow rotating 0.1M(cid:12) star. The full colored lines correspond to semi-major axis evolution. The full red line and the dashed red line respectively correspond to the corotation radius of a fast rotating M-dwarf and of a slow rotating M- dwarf with no planet. The black long dashes represent the Roche limit. major axes. However, the radii of 0.1M(cid:12) stars and brown dwarfs decrease substantially in time such that after a certain interval the stellar/brown dwarf tide becomes weak and the system freezes (see explanation in Subsection 3.2). Thus, we expect that similar conclusions can be drawn for planets around 0.1M(cid:12) stars and brown dwarfs. A statistical dis- tribution of planets around fully convective M-dwarfs can pro- vide information about their dissipation factors. Indeed, as in Bolmont et al. (2011), we can make inferences from how densely packed the orbital distribution of close-in planets is; a higher stellar dissipation leading to a more packed distribution of final semi-major axis, and increasing the shortest possible final semi- major axis (see Fig. 19 and 20 of Bolmont et al., 2011). For Jupiter mass planets around 1M(cid:12) stars the situation is different. Planets which survived the early evolution due to the crossing of the shrinking corotation radius still fall on the star on Gyr timescale due to the spin-down of the star. Figure 7 shows the evolution of Jupiter mass planets around a Sun-like star. For fast rotators, either the planets fall very quickly com- pared to the stellar evolution or they fall in more than a few 108 yrs. Contrary to M-dwarfs , the tidal forces at late ages are still important for Sun-like stars (see Subsection 3.2). So the dif- ferences of evolution at early ages due to the difference of initial stellar spin period disappear in the end due to the long term evo- lution. No conclusions can be drawn as was the case for fully convective M-dwarfs . 4.3. Planetary influence on spin evolution The spin evolution of a star with no planets is determined by its initial spin, its contraction rate and the efficiency of its stellar wind. However, if a planet is orbiting the star there are angular momentum exchanges between the orbit of the planet and the spin of the star. In particular, if a planet is spiraling inward to- wards the star, it will have the effect of making the star spin up. 7 Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets Fig. 7. Tidal evolution of a 1MJ mass planet starting at different ini- tial semi-major axis around either a fast rotating (top panel) or a slow rotating (bottom panel) 1M(cid:12) star. The full colored lines correspond to semi-major axis evolution. The colored dashed lines correspond to the corotation radius and the red dashed lines correspond to the corotation radius of a star with no planet. The black long dashes represent the Roche limit. Fig. 8. Tidal evolution of a Jupiter mass planet starting at different initial semi-major axis around either fast rotating or a slow rotating 0.1M(cid:12) star. Top panel: evolution of semi-major axis of a a planet around an initially fast rotating M-dwarf (bold solid lines) and around an initially slow rotating M-dwarf (dashed dotted lines). Bottom panel: Rotation period of initially fast rotators (solid lines) and slow rotators (dashed dotted lines). The black long dashes represent Tsat = 2π/ωsat. The red curves represent the evolution of the spin of the star if there is no planet. 4.3.1. Spinning-up and mergers Figure 8 shows the evolution of Jupiter-mass planets around ei- ther initially fast rotating or an initially slow rotating 0.1M(cid:12) stars, using mean dissipation values. In Irwin et al. (2011) the time of the dispersal of the gas disk, τdisk, varies with the na- ture of the star - fast rotator or slow rotator. Here we considered that both fast and slow rotators decoupled from the disk at the same time t0 = 8 Myrs. This way we can compare the effects of tides on the evolution of planets beginning in the same initial conditions - except of course the initial stellar spin. As above, we see in Fig. 8 that planets can survive in a wider range of initial semi-major axes around initially fast ro- tating stars. Compared with Fig. 4, the stellar spin evolution is more strongly modified by planetary migration. When the planet begins spiraling towards the star due to stellar tide, angular mo- mentum is transferred from the planet's orbit to the stellar spin. Thus, a falling planet can turn a slowly rotating star into a fast rotating star. In this work we followed the prescription on Irwin et al. (2011) such that initially fast rotators always have K f ast as wind parameter and initially slow rotators always have Kslow. However, in the simulations from Fig. 8 some slow rotators actually become fast rotators early in their history. Given the prescription from Irwin et al. (2011), a slow-turned-fast rotator evolves back into a slow rotator due to the difference in wind parameterization. One can argue that a slow-turned-fast rotator should rather switch and spin down like a fast rotator instead. 8 The late spin evolution of such a star would have little effect on any surviving planets because they must lie at larger orbital distances than the planet that perished and also because after a few ×107 yrs, the radius of the star is small enough for the tidal effects to be negligible. The planet starting at 0.014 AU around a slowly rotating star (middle panel of Fig. 8) has an interesting evolution. The planet starts inside the corotation radius, so the stellar tide pulls it in- wards. Given the planet's large mass, it transfers a significant amount of angular momentum to the star and spins it up. The corotation radius thus shrinks until it catches up with the planet and inverts the tidal forces on the planet, which then experi- ences a slow outward migration. At 8× 107 yrs, the star has spun down due to the stellar winds sufficiently that the planet crosses back inside the corotation radius. For the rest of its evolution the planet migrated slowly inward. Stellar spin-up when a planet migrates in and falls on the star is also apparent for Sun-like stars. Figure 5 shows delayed peeks in the spin rate when the star experiences a merger2 with a Jupiter-mass planet. When a merger occurs, the star has an 2 Metzger et al. (2012) identified three different outcomes for the merger of a planet into a star, depending on at what orbital distance the Roche lobe of the planet becomes smaller than the actual size of the planet. They also discuss the observational signatures of such events. Some configurations lead to a luminosity peak which could last days or years. However, here we keep a simple description of merger events, and assume that a merger occurs when the planet reaches the Roche limit. Then, all the angular momentum is transferred from the planet's orbit into the spin of the star. Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets excess rotation which disappears on Gyr timescales due to stellar winds. For a system with a planet beginning at 0.06 AU, the merger occurs after 8 Gyrs, but just after the merger this old star spins five times faster than the Sun today, corresponding to a equatorial velocity of about 10 km/s. This effect is stronger for more massive planets. A 5 MJ planet beginning its evolution at 0.06 AU falls on the star in about 8 Gyr. This makes the parent star spin up to more than 20 times faster than present Sun (see Fig. 9), with an equatorial velocity of about 60 km/s. This corresponds approximatively to the fast rotators' initial rotation period. Fig. 10. Tidal evolution of a Jupiter mass planet around a fast rotating 0.1M(cid:12) star with a high dissipation factor. Top panel : Evolution of semi- major axis of Jupiter mass planets (full line) beginning at different initial semi-major axis and of the corresponding corotation distance (dashed line). The red dashed line corresponds to the corotation radius of a star with no planet. Middle panel : Evolution of the eccentricity of the plan- ets. Bottom panel : Evolution of the rotation period of the planets (full line) and of the corresponding star (dashed line). In both cases, the stel- lar spin initially increases to reach the spin period of the planet (we consider planets in pseudo-synchronization). Later in the evolution ei- ther the spin-up due to contraction or the spin-down due to stellar wind leads the spin of the star to differentiate from the spin of the planet. Tidal theory states that an equilibrium state is obtained for the two body problem when the orbit is circular, and when both orbit and spins are aligned and synchronized (Hut, 1980). In this equilibrium state, both planet and star thus have a spin which is equal to the mean orbital frequency, Ω∗ = Ωp = n. The planet having a low inertia compared to the star, it will reach synchro- nization early in the evolution, typically in a few thousand years. However, usually the star because of its higher inertia never reaches synchronization. In most cases, it would require much more than 10 Gyr for the star to reach synchronization. However, in the example shown in Fig. 10, where the stellar dissipation has been significantly increased, the stellar tide is efficient enough to bring about synchronization of the stellar rotation with the spin of the planet in a few 104 years to a few 105 years. Fig. 10 shows a case. Let us focus on the case of the pur- ple curve which corresponds to a planet beginning at 0.014 AU. Initially the stellar rotation period is of 1.2 days, which corre- sponds to an corotation distance of about 0.01 AU. In 104 years, the corotation radius moves outward and reached the planet's orbital distance (as can be seen in the bottom panel, the stellar spin period increases to match the planet's orbital period; re- call that we assume pseudo-synchronization for planets). After 105 years the stellar tide has made the eccentricity decrease to 9 Fig. 9. As Figure 5, but with 5MJ planets. We can see that the planet being more massive than in Fig. 5 the acceleration of the star is more visible. Some slow rotators even become fast rotators for a few 10 Myr because of the merger of the planet. When a merger occurs the star is spun up to a rotation rate which is of the order of 20 Ω(cid:12). Unlike close-in planets around 0.1M(cid:12) M-dwarfs , which sur- vived for Gyr timescales if they did not fall in first few Myrs, close-in planets orbiting Sun-like stars that survived the early evolution are still doomed to fall on the star. This difference in behavior can be explained by the fact that M-dwarfs have much smaller radii than Sun-like stars at late ages and thus much weaker stellar tides. For a planet at 0.03 AU, the semi-major axis evolution timescale τa for a 5 Gyr old M-dwarf is longer than 10 Gyr (Fig. 2), but just 1 Gyr for a 5 Gyr old Sun-like star with a dissipation of 1000 × σ∗. M dwarf stellar tides are simply not strong enough to make the planets fall at late times. 4.3.2. Stellar synchronization We now study a system that displays relatively spectacular tidal evolution. The system is made up of a Jupiter-mass planet or- biting an initially fast rotating, highly dissipative (1000 × σ∗) 0.1M(cid:12) star. The planetary dissipation is set to zero to isolate the influence the stellar tide. Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets Ne1(e) zero. Indeed, because Ω∗/n < 18 Ne2(e) and because of the pseudo 11 synchronization of the planet, this system would be in the region where e < 0 of the phase diagram ω/n vs e presented in Leconte et al. (2010). Between 105 years and 107 years, the star and the planet remain in perfect synchronization with a circular orbit and spin synchronization. This is a equilibrium state for the system, or it would be if the star was not evolving in time. Indeed, the stellar radius shrinks which acts to spin-up the star while stellar winds act to spin-down the star, and these effects drive the sys- tem away from this equilibrium state. After a few ×107 years, contraction spins the star up, and then stellar winds make the star spin down until 1 Gyr. The planet has crossed the corotation radius and the stellar tide causes it to fall slowly towards the star. The orbital angular momentum of the planet is transferred to the star and spins it slowly back up. The other (blue) planet in Fig. 10 shows a similar evolution but with two differences. First, the effect of the stellar tide is weaker because the planet is initially farther out (0.022 AU), so synchronization occurs later and for a shorter time. Second, the other physical phenomena influencing the spin evolution of the star kick in earlier. The acceleration phase is much more pro- nounced for this case than for the previous one. The second dif- Ne1(e) ference is that initially, Ω∗/n > 18 Ne2(e), the star is in the phase 11 space region where e > 0, so the stellar tide contributes in in- creasing the eccentricity. However, the spin of the star decreases faster than the mean orbital angular frequency n so around a few 104 years, the system is back in the region where e < 0 and the stellar tide contributes in decreasing the eccentricity. Between ∼ 6 Myr and ∼ 2 Gyr, the planet is outside the corotation radius Ne1(e) and Ω∗/n > 18 Ne2(e) so the stellar tide contributes in increasing 11 the eccentricity once again. However, in the meantime the radius has shrunk to values such that the stellar tide is negligible and a small eccentricity is kept in the system. When the star starts to spin down due to the stellar wind, the both planet and star are once more in the phase space region where e < 0. This case represents an interesting case study although it is not likely that the stellar dissipation is strong enough to allow such evolution. 4.3.3. Implication on stellar age determination As we have seen, tidal interactions between a star and a hot Jupiter can bring about significant acceleration of the spin of the star in some cases. This is the same main conclusion as Pont (2009), who explains that the observed excess rotation of stars with a transiting system can be due to the tidal interaction be- tween the planet and the star. We note that stars without planets may still bear the traces of their violent pasts due to planet-star mergers. We can see in Fig. 5 and 9 that while approaching the star, the spin of the star changes continuously to reach a maxi- mum at the merger. To determine the age of a star, one technique is "stellar gy- rochronology" (Barnes, 2003; Barnes et al., 2010; Meibom et al., 2011; Epstein & Pinsonneault, 2012). It consists of measuring the spin of a star, and inferring the age of the star by assuming a spin history. Generally speaking, fast rotating stars are classified as young stars and slow rotating stars are classified as old stars that have been spun down due to stellar winds. In light of this study, we can see that this method might, in some cases, imply the wrong ages of rapidly rotating stars. A rapidly rotating star could be a young star but it could also be an old star which experienced a merger with a planet. For example, the merger of 5 MJ planet on an initially fast or slow rotating 10 makes the star spin up almost to the initial rotation period of the group of fast rotators. Thus, old stars that are spun up would be mistaken for young objects. An old star that underwent a merger with a Jupiter size planet would reproduce several characteristics of young stars: they will be fast rotators, they will have infrared excess due to a hot dust disk and accretion signatures. A merger will also produce ex- treme UV signatures, soft X-rays and a peak of luminosity de- pending on the nature of the merger (Metzger et al., 2012). These stars would most likely be mistaken for young stars. Gyrochronological ages of fast rotators should be checked against other techniques, in order to verify their youth. Another technique consists of estimating the age of the star by comparing its location on the Herzprung-Russel diagram with theoretical stellar evolution tracks. 5. Conclusions We have shown that the stellar spin history affects the tidal evo- lution of close-in planets, albeit in a confined region of param- eter space. In cases where the stellar spin history does matter, the difference between two close-in planets -- one orbiting an initially fast-rotating star and the other orbiting an initially slow- rotating star -- comes from an early phase of outward tidal mi- gration for the planet around the fast rotator (caused by the star's closer-in corotation distance). This phase of outward migration does not occur for the planet orbiting the slow rotator. This out- ward migration and the decrease of the radius of the star weaken the later tidal evolution and effectively delay or sometimes even prevent the later in-spiralling of planets onto their stars. At later times both slow and fast rotators spin down due to stellar winds (Skumanich, 1972; Bouvier et al., 1997; Bouvier, 2008; Irwin et al., 2011), so the orbital history of close-in planets orbiting old stars depends on something that is not directly observable: the stellar spin evolution. For Sun-like stars the stellar spin history affects tidal evo- lution only in relatively extreme circumstances. In particular, the spin history has a strong effect if stars are very dissipative, with dissipation rates σ∗ of about 1000 times the fiducial value (Hansen, 2010). For strongly dissipative stars, there are differ- ences in tidal evolution for planets orbiting initially slowly- vs. rapidly rotating stars for very close-in (a (cid:46) 0.05 AU), massive planets (Mp (cid:38) MJ). In contrast, the stellar spin history plays a role in a much wider region of parameter space for 0.1M(cid:12) stars, mainly be- cause these stars are fully convective and so we think they are much more dissipative, with dissipation factors assumed to match those of brown dwarfs. For these stars the differences in tidal evolution for planets orbiting initially slowly- vs. fast- rotating stars are apparent for mean dissipation values, for plan- ets out to ∼ 0.02 AU, and for planets with masses as small as 1M⊕. Low mass stars and Sun-like stars have different dissipation factors and different radii, so the evolution timescales are differ- ent and evolve differently (see Figs. 2 and 3). The early tidal in- teraction is stronger for planets around very low mass stars, and the difference between fast rotator profile and slow rotator pro- file is apparent for planets of one Earth mass with mean dissipa- tion values. After some time, the system freezes in a given state because the radius of the star has shrunk too much for the tidal evolution to occur on less than 10 Gyr timescale. For Sun-like stars, the tidal evolution for mean dissipation factor occurs on very long timescales, which is why much more massive planets Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets and higher stellar dissipation rates are needed to produce stellar spin-driven differences in tidal evolution. The inward migration of a Jupiter orbiting inside the coro- tation radius of an initially slow rotating 0.1M(cid:12) star can lead to a significant spinning-up of the star. By the transfer of angular momentum from the planet's orbit, an initially slow rotating star can become a fast rotating star. This effect is more dramatic if the planet actually falls on the star. Irwin et al. (2011) used dif- ferent wind parametrization for fast or slow rotators so they can infer from the present spin rate if the star was initially fast ro- tating or not. However we show here that it might not be that straight forward. If the star experienced a merger with a planet it can modify the rotation rate of the star and change the slow rotator into a fast rotator. Massive planets orbiting very low-mass stars with high dis- sipation rates (σ∗ × 1000) can create systems in perfect synchro- nization where the spin of the star is equal to the spin of the planet (Fig. 10). However, the equilibrium is not stable and the system departs from it as the star spins up due to contraction or spins down due to the stellar winds. This strongly alters the stel- lar spin profile because the star can be efficiently spun down by a planet initially located outside the corotation radius or spun up by a planet interior to corotation. Unfortunately, hot Jupiters around M-dwarfs are extremely rare due to the inefficiency of the planets formation processes around low mass stars (Laughlin et al., 2004; Ida & Lin, 2005; Kennedy & Kenyon, 2008). Only three hot Jupiters are known to exist around stars with masses less than 0.7M(cid:12) (Pepe et al., 2004; Hellier et al., 2011; Borucki et al., 2011; Johnson et al., 2012) and none around stars with masses as small as 0.1M(cid:12). Hot Jupiters around very low mass stars remain to be detected. A statistical distribution of planets around fully convec- tive M-dwarfs could constrain the tidal dissipation factor σ∗. Specifically, from the location of the inner edge of the plane- tary semi-major distribution one can infer a inferior limit for the dissipation factor. The more distant the inner edge, the more dis- sipative the star. However, in order to draw these conclusions, one needs a good estimate of the stellar ages. For Sun-like stars such conclusions cannot be made because, if the dissipation rate is high enough to affect the orbital evolu- tion at early times, significant tidal evolution still takes place at late times, as well. Slow rotators and fast rotators have similar evolutions after a few 108 yrs so the observation of a hot Jupiter orbiting a star of known age and known dissipation would not allow us to infer if the stellar spin history. Indeed, in both cases, different initial semi-major axis can lead to the same observed semi-major axis. One can imagine trying to infer a planet's or- bital evolution from the composition of its atmosphere to know if the planet came from a "cold" region (0.04 AU) or a "hot" region (0.03 AU). Unfortunately, this exercise is fraught with uncertain- ties in both the expected atmospheric composition of planets that form at different orbital distances and the tidal parameters (Ω∗,0, σ∗, the age of observed stars). Nonetheless, we emphasize the planets crashing on the star at late ages can entail a significant spin-up of the star and cre- ate a population of old fast rotating stars. The spin-up of the star due to a merger has been pointed out in Levrard et al. (2009), where they found that planets falling on their host star due to tides never reach a tidal equilibrium. Jackson et al. (2009) also addressed the problem of tidally induced mergers and the effect of these mergers on the parent star. They also found that a con- siderable spin-up is to be expected and also a change in stellar composition. Planet-star mergers thus may confuse stellar age determinations. In general, fast rotators are thought to be young, although we have shown that a merger can lead to old, fast rotat- ing stars that would mimic many of the characteristics of young stars. An independent determination of the age of observed stars is therefore very important, especially for fast rotators. Acknowledgements. We are grateful to Andrew West for suggesting that we study low-mass stars rather than simply concentrating on the case of Sun-like stars. We thank Franck Selsis and Franck Hersant for their ideas and support. We also thank the CNRS's PNP program for funding, and the Conseil Regional d'Aquitaine for help purchasing the computers on which these calculations were performed. SPM acknowledges support by the ERC through grant 207430 STARS2 (http://www.stars2.eu).. References Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 1998, A&A, 337, 403 Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 2002, A&A, 382, 563 Barnes, R., Raymond, S. N., Greenberg, R., Jackson, B., & Kaib, N. A. 2010, ApJ, 709, L95 Barnes, S. A. 2003, ApJ, 586, 464 Benedict, G. F., McArthur, B., Nelan, E., et al. 1998, AJ, 116, 429 Bolmont, E., Raymond, S. N., & Leconte, J. 2011, A&A, 535, A94 Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 736, 19 Bouvier, J. 2008, A&A, 489, L53 Bouvier, J., Forestini, M., & Allain, S. 1997, A&A, 326, 1023 Browning, M. K. 2008, ApJ, 676, 1262 Carone, L. & Patzold, M. 2007, Planet. Space Sci., 55, 643 Chabrier, G. & Baraffe, I. 1997, A&A, 327, 1039 Chabrier, G. & Kuker, M. 2006, A&A, 446, 1027 Charbonneau, D., Berta, Z. K., Irwin, J., et al. 2009, Nature, 462, 891 Choi, P. I. & Herbst, W. 1996, AJ, 111, 283 Christian, D. J., Pollacco, D. L., Skillen, I., et al. 2006, MNRAS, 372, 1117 Collier Cameron, A., Guenther, E., Smalley, B., et al. 2010, MNRAS, 407, 507 Dobbs-Dixon, I., Lin, D. N. C., & Mardling, R. A. 2004, ApJ, 610, 464 Edwards, S., Strom, S. E., Hartigan, P., et al. 1993, AJ, 106, 372 Eggleton, P. P., Kiseleva, L. G., & Hut, P. 1998, ApJ, 499, 853 Epstein, C. R. & Pinsonneault, M. H. 2012, ArXiv e-prints Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661 Ghosh, P. & Lamb, F. K. 1978, ApJ, 223, L83 Hansen, B. M. S. 2010, ApJ, 723, 285 Hartmann, L. & MacGregor, K. B. 1982, ApJ, 259, 180 Hartmann, L. & Stauffer, J. R. 1989, AJ, 97, 873 Heller, R., Leconte, J., & Barnes, R. 2011, A&A, 528, A27+ Hellier, C., Anderson, D. R., Collier Cameron, A., et al. 2011, A&A, 535, L7 Herbst, W., Bailer-Jones, C. A. L., & Mundt, R. 2001, ApJ, 554, L197 Herbst, W., Eisloffel, J., Mundt, R., & Scholz, A. 2007, Protostars and Planets V, 297 Hut, P. 1980, A&A, 92, 167 Hut, P. 1981, A&A, 99, 126 Ida, S. & Lin, D. N. C. 2005, ApJ, 626, 1045 Irwin, J., Berta, Z. K., Burke, C. J., et al. 2011, ApJ, 727, 56 Irwin, J., Hodgkin, S., Aigrain, S., et al. 2008, MNRAS, 383, 1588 Jackson, B., Barnes, R., & Greenberg, R. 2009, ApJ, 698, 1357 Johnson, J. A., Gazak, J. Z., Apps, K., et al. 2012, AJ, 143, 111 Kawaler, S. D. 1988, ApJ, 333, 236 Kennedy, G. M. & Kenyon, S. J. 2008, ApJ, 673, 502 Kiraga, M. & Stepien, K. 2007, Acta Astron., 57, 149 Lanza, A. F., Damiani, C., & Gandolfi, D. 2011, A&A, 529, A50 Laughlin, G., Bodenheimer, P., & Adams, F. C. 2004, ApJ, 612, L73 Leconte, J., Chabrier, G., Baraffe, I., & Levrard, B. 2010, A&A, 516, A64+ Levrard, B., Winisdoerffer, C., & Chabrier, G. 2009, ApJ, 692, L9 Lin, D. N. C., Bodenheimer, P., & Richardson, D. C. 1996, Nature, 380, 606 MacGregor, K. B. & Brenner, M. 1991, ApJ, 376, 204 Matt, S. & Pudritz, R. E. 2005a, ApJ, 632, L135 Matt, S. & Pudritz, R. E. 2005b, MNRAS, 356, 167 Matt, S. & Pudritz, R. E. 2008, ApJ, 681, 391 Matt, S. P., Pinz´on, G., de la Reza, R., & Greene, T. P. 2010, ApJ, 714, 989 Matt, S. P., Pinz´on, G., Greene, T. P., & Pudritz, R. E. 2012, ApJ, 745, 101 Meibom, S., Mathieu, R. D., Stassun, K. G., Liebesny, P., & Saar, S. H. 2011, ApJ, 733, 115 Metzger, B. D., Giannios, D., & Spiegel, D. S. 2012, ArXiv e-prints Morin, J., Donati, J.-F., Petit, P., et al. 2008, MNRAS, 390, 567 Neron de Surgy, O. & Laskar, J. 1997, A&A, 318, 975 Paatz, G. & Camenzind, M. 1996, A&A, 308, 77 Patzold, M., Carone, L., & Rauer, H. 2004, A&A, 427, 1075 Pepe, F., Mayor, M., Queloz, D., et al. 2004, A&A, 423, 385 11 Emeline Bolmont et al.: Effect of the stellar spin history on the tidal evolution of close-in planets Pont, F. 2009, MNRAS, 396, 1789 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, Numerical recipes in FORTRAN. The art of scientific computing (Cambridge University Press) Raymond, S. N., Scalo, J., & Meadows, V. S. 2007, ApJ, 669, 606 Rebull, L. M., Stauffer, J. R., Megeath, S. T., Hora, J. L., & Hartmann, L. 2006, ApJ, 646, 297 Rebull, L. M., Wolff, S. C., & Strom, S. E. 2004, AJ, 127, 1029 Reiners, A. & Basri, G. 2007, ApJ, 656, 1121 Safronov, V. S. 1969, Evoliutsiia doplanetnogo oblaka. (English transl.: Evolution of the Protoplanetary Cloud and Formation of Earth and the Planets, NASA Tech. Transl. F-677, Jerusalem: Israel Sci. Transl. 1972) Shu, F., Najita, J., Ostriker, E., et al. 1994, ApJ, 429, 781 Skumanich, A. 1972, ApJ, 171, 565 Snellen, I. A. G., Koppenhoefer, J., van der Burg, R. F. J., et al. 2009, A&A, 497, 545 Stassun, K. G., Mathieu, R. D., Mazeh, T., & Vrba, F. J. 1999, AJ, 117, 2941 Tout, C. A. & Pringle, J. E. 1992, MNRAS, 256, 269 Zahn, J.-P. 1994, A&A, 288, 829 12
1811.05860
1
1811
2018-11-14T15:43:00
Heliospheric modulation of the interstellar dust flow on to Earth
[ "astro-ph.EP" ]
Aims. Based on measurements by the Ulysses spacecraft and high-resolution modelling of the motion of interstellar dust (ISD) through the heliosphere we predict the ISD flow in the inner planetary system and on to the Earth. This is the third paper in a series of three about the flow and filtering of the ISD. Methods. Micrometer- and sub-micrometer-sized dust particles are subject to solar gravity and radiation pressure as well as to interactions with the interplanetary magnetic field that result in a complex size-dependent flow pattern of ISD in the planetary system. With high-resolution dynamical modelling we study the time-resolved flux and mass distribution of ISD and the requirements for detection of ISD near the Earth. Results. Along the Earth orbit the density, speed, and flow direction of ISD depend strongly on the Earth's position and the size of the interstellar grains. A broad maximum of the ISD flux (2x10^{-4}/m^2/s of particles with radii >~0.3\mu m) occurs in March when the Earth moves against the ISD flow. During this time period the relative speed with respect to the Earth is highest (~60 km/s), whereas in September when the Earth moves with the ISD flow, both the flux and the speed are lowest (<~10 km/s). The mean ISD mass flow on to the Earth is ~100 kg/year with the highest flux of ~3.5kg/day occurring for about 2 weeks close to the end of the year when the Earth passes near the narrow gravitational focus region downstream from the Sun. The phase of the 22-year solar wind cycle has a strong effect on the number density and flow of sub-micrometer-sized ISD particles. During the years of maximum electromagnetic focussing (year 2031 +/- 3) there is a chance that ISD particles with sizes even below 0.1\mu m can reach the Earth. Conclusions. We demonstrate that ISD can be effectively detected, analysed, and collected by space probes at 1 AU distance from the Sun.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. aa November 15, 2018 c(cid:13)ESO 2018 8 1 0 2 v o N 4 1 . ] P E h p - o r t s a [ 1 v 0 6 8 5 0 . 1 1 8 1 : v i X r a Heliospheric modulation of the interstellar dust flow on to Earth Peter Strub(cid:63)1, Veerle J. Sterken2, Rachel Soja3, Harald Krüger1, Eberhard Grün4, 5, and Ralf Srama3, 1 Max-Planck-Institut für Sonnensystemforschung, Justus-von-Liebig-Weg 3, 37077 Göttingen, Germany 2 Astronomical Institute University of Bern, Sidlerstrasse 5, 3012 Bern, Switzerland 3 Institut für Raumfahrtsysteme, Universität Stuttgart, Pfaffenwaldring 29, 70569 Stuttgart, Germany 4 Max-Planck-Institut für Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany 5 Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, CO 80303, USA ABSTRACT Aims. Based on measurements by the Ulysses spacecraft and high-resolution modelling of the motion of interstellar dust (ISD) through the heliosphere we predict the ISD flow in the inner planetary system and on to the Earth. This is the third paper in a series of three about the flow and filtering of the ISD. Methods. Micrometer- and sub-micrometer-sized dust particles are subject to solar gravity and radiation pressure as well as to in- teractions with the interplanetary magnetic field that result in a complex size-dependent flow pattern of ISD in the planetary system. With high-resolution dynamical modelling we study the time-resolved flux and mass distribution of ISD and assess the necessary requirements for detection of ISD near the Earth. Results. Along the Earth orbit the density, speed, and flow direction of ISD depend strongly on the Earth's position and the size of the interstellar grains. A broad maximum of the ISD flux (∼ 2 × 10−4m−2s−1 of particles with radii (cid:38) 0.3 µm) occurs in March when the Earth moves against the ISD flow. During this time period the relative speed with respect to the Earth is highest (∼ 60 km s−1), whereas in September when the Earth moves with the ISD flow, both the flux and the speed are lowest ((cid:46) 10 km s−1). The mean ISD mass flow on to the Earth is approximately 100 kg year−1 with the highest flux of ∼ 3.5 kg day−1 occurring for about 2 weeks close to the end of the year when the Earth passes near the narrow gravitational focus region of the incoming ISD flow, downstream from the Sun. The phase of the 22-year solar wind cycle has a strong effect on the number density and flow of sub-micrometer-sized ISD particles. During the years of maximum electromagnetic focussing (year 2031 +/- 3) there is a chance that ISD particles with sizes even below 0.1 µm can reach the Earth. Conclusions. We demonstrate that ISD can be effectively detected, analysed, and even collected by space probes at 1 AU distance from the Sun. Key words. interstellar dust -- dust -- heliosphere 1. Introduction Interstellar dust (ISD) particles are messengers from the remote sites where they formed and from the environment that they tra- versed during their journey through space and time. They are born as stardust and take their initial elemental and isotopic sig- natures from the cool atmospheres of giant stars or from stel- lar explosions. Ultraviolet (UV) irradiation, interstellar shock waves, and mutual collisions modify and deplete particles in the interstellar medium (ISM). In dense molecular clouds particles grow by agglomeration and accretion. The Sun and the heliosphere are surrounded by a local dense warm cloud of gas and dust, the Local Interstellar Cloud (LIC). About 1% of the mass of this cloud is ISD. The motion of the heliosphere with respect to this cloud causes an inflow of the ISD into the heliosphere from a direction of 259◦ ecliptic lon- gitude, and +8◦ ecliptic latitude (Landgraf 1998; Frisch et al. 1999a; Strub et al. 2015). The relative velocity of the ISD in the heliosphere is 26 km s−1 (Grün et al. 1994; Strub et al. 2015). Most of our knowledge of ISD comes from astronomical ob- servations: modelling combined with observations of the depen- dence of starlight extinction on wavelength reveals some mate- rial properties and allows an ISD size distribution to be derived (cid:63) Contact: [email protected], [email protected] (Mathis et al. 1977; Draine & Lee 1984; Weingartner & Draine 2001; Zubko et al. 2004). Interstellar dust was first positively identified inside the solar system more than two decades ago. After its fly-by of Jupiter, the dust detector onboard the Ulysses spacecraft detected impacts of micron and submicron-sized particles (10−17 to 10−14 kg) pre- dominantly from a direction opposite to the expected impact di- rection of interplanetary dust particles (Grün et al. 1993). On average, the measured impact velocities exceeded the local solar system escape velocity (Grün et al. 1994). Subsequent analy- sis showed that the motion of the ISD particles through the so- lar system was parallel to the flow of neutral interstellar hydro- gen and helium gas (Frisch et al. 1999a). While the ISD flow persisted at higher latitudes above the ecliptic plane, even over the poles of the Sun, the interplanetary dust is strongly depleted away from the ecliptic plane. The flow patterns of ISD through the heliosphere were dis- cussed in detail by Sterken et al. (2012). Sterken et al. (2013) described the modulation of the ISD size distribution at Sat- urn, Jupiter, and asteroid distances from the Sun (3 AU). The observed variation of the ISD flow is due to the modulation of the dust stream by the Lorentz force, the radiation pres- sure force, and gravity (Landgraf 2000a). Particles with a ra- dius ad (cid:38) 100 nm pass the heliospheric bow shock and enter the heliosphere (Linde & Gombosi 2000). As a consequence, inter- Article number, page 1 of 17 A&A proofs: manuscript no. aa stellar particles constitute the dominant known particulate com- ponent in the outer solar system (in number flux, not in mass flux). Although a large portion of the ISD measurements by Ulysses were obtained far from the ecliptic plane and outside the inner solar system, it is also known that ISD particles of various sizes can reach the Earth's orbit (Altobelli et al. 2003, 2005, 2006). This opens the possibility for Earth-orbiting space- craft to detect and study ISD. Previous models, while accurately describing the ISD environment at larger heliocentric distances, did not have the resolution to enable a good time-resolved under- standing of the dust environment at the Earth (Grün et al. 1994; Landgraf 2000b; Sterken et al. 2012). In this paper we describe a model of sufficiently high spatial resolution to allow the study of the characteristics of the ISD flow at the Earth. We study the time-resolved flux and mass distribution and assess the neces- sary requirements for detection of ISD near the Earth. Section 2 describes the physical modelling of the ISD dy- namics in the solar system, including initial conditions, the three modelled forces, and the material assumptions for the modelling. Section 3 presents the Monte Carlo computer simulation results of the ISD flow at Earth orbit. Section 4 discusses in detail the temporal evolution of ISD observing conditions (density, speed, impact direction and size distribution) relative to Earth and the total mass influx on to Earth. Section 5 concludes this study with an outlook for ISD missions at 1 AU. 2. A time-resolved model of the ISD flow through the solar system The ISD model is a three-dimensional (3D), time-resolved repre- sentation of the flow properties of ISD particles in the inner solar system created using a Monte Carlo simulation. It is generated by integrating the equation of motion of ISD particles as they move through the solar system in order to obtain particle tra- jectories. These trajectories are used to produce data cubes that contain the average density, velocity components, and velocity dispersion of particles in each region of space (within 10 AU of the Sun) and time (within the solar cycle) for a given particle size. Sterken et al. (2015, Fig. 3-4) described the effect of differ- ent interplanetary magnetic field (IMF) models on the simula- tion results and used a constant rate of change of the solar mag- netic dipole for describing the ISD flux near Saturn, Jupiter, and the asteroid belt. The model is based on the work of Landgraf (2000a) and Sterken et al. (2012, 2013). The main goal of our model is to improve both the spatial and temporal resolution by a factor of six over the model of Sterken et al. (2012, 2013). This leads to a spatial mesh cell size of 0.25 AU, with a temporal res- olution of approximately 12.2 days, as opposed to the resolution of 1.5 AU and 73 days in the earlier model. Also, the variable rotation rate of the solar magnetic dipole from WSO Observa- tions is used for discussing the ISD flow and flux near the Earth, approximated by a piecewise constant rotation rate. The initial conditions, relevant forces, and general flow prop- erties are briefly introduced in Sections 2.1 and 2.2. A more de- tailed description of the latter is given in Sterken et al. (2012, model and flow pattern) and Sterken et al. (2013, filtering). ter the solar system at a uniform direction and velocity, with their positions randomly distributed on a plane 50 AU upstream of the Sun and perpendicular to the velocity vector. We use an initial velocity of v∞ = 26 kms−1 and an inflow direction from an eclip- tic longitude1 lecl = 259◦ and ecliptic latitude becl = 8◦. This is the same as the direction of the neutral gas flow inside the so- lar system (Witte et al. 1993; Lallement & Bertaux 2014; Wood et al. 2015), and it is compatible with the Ulysses measurements of the ISD flow (Frisch et al. 1999b; Strub et al. 2015). This is equivalent to the ISD particles being at rest with respect to the LIC. Our model does not simulate the effect of the heliospheric boundary on the ISD trajectories, which results in a diminu- tion of the small particle contribution (Linde & Gombosi 2000). However, as the normalisation of the simulated dust densi- ties is based on Ulysses dust fluxes measured inside the helio- sphere, the time-averaged effects of the filtering at the helio- spheric transition region have been accounted for. A possible time-dependence of the filtering as suggested by Sterken et al. (2015) is not reflected in the present model. 2.2. Relevant forces The trajectories of charged ISD particles were integrated under solar gravity, the solar radiation pressure force, and the Lorentz force from the charged particles moving through the IMF. The solar radiation pressure force and gravity are not considered vari- able in time and therefore produce a stationary flow pattern. The Lorentz force depends on the solar cycle, and therefore leads to a temporal variability in the flow of ISD (Morfill & Grün 1979; Landgraf 2000b; Sterken et al. 2012). The solar radiation pressure force and gravity both decline with the square distance to the Sun. We can therefore use the solar radiation pressure force divided by the gravity as a dimen- sionless constant parameter (β) that depends on the particle ma- terial properties such as particle size, morphology, and the re- flectivity and absorptivity of the particle integrated over the solar spectral energy distribution (SED) (Burns et al. 1979). We use the same β-curve as Sterken et al. (2013), which was adapted from the curve for astronomical silicates given in Gustafson (1994) by scaling it to match the maximum value βmax (cid:39) 1.4 − 1.8 (average 1.6) measured by Ulysses (Landgraf et al. 1999). Sterken et al. (2013) discuss the filtering of ISD at different places in the solar system (as close as the asteroid belt) for this β-curve, as well as independently of any β-curve. Landgraf (2000b) used the astronomical silicates β-curve with a maximum of 1.4 (Gustafson 1994; Draine & Lee 1984). The Lorentz force is important for the dynamics of ISD par- ticles with radius ad (cid:46) 0.4 µm. The particles become charged as they move through the heliosphere, as a result of (1) predomi- nantly the loss of electrons through photoionisation by solar UV light, (2) the collection of ions and electrons from the ambient solar wind plasma, and (3) secondary electron emission. Similar to previous studies, we assume a constant equilibrium potential of +5 V, which is compatible with Cassini CDA measurements of charged particles (Kempf et al. 2004) and theoretical charg- ing models (Mukai 1981; Horányi 1996). A constant charge of +5 V is a good approximation for particles moving through the solar wind plasma between 50 and 1 AU from the Sun (Slavin et al. 2012; Kimura & Mann 1998). The resulting Lorentz force 2.1. Initial conditions The model uses the same initial conditions as Landgraf (2000a) and Sterken et al. (2012, 2013): the simulated ISD particles en- 1 We follow the convention to use l, b to denote ecliptic coordinates given in the heliocentric frame of reference, and λ, β in the geocentric frame of reference. Article number, page 2 of 17 Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth depends on the particle's charge to mass ratio (Q/m), its veloc- ity relative to the solar wind velocity (rp,sw), and the strength of the IMF at the particle's location, Bsw. Q/m is defined using the +5 V potential, along with an assumption of compact spherical particles with a constant bulk density of 2 g cm−3. The interplanetary magnetic field is the continuation of the solar magnetic field in interplanetary space, frozen into the plasma of the solar wind. Due to the Sun's rotation, its outward motion leads to a spiral pattern of alternating magnetic field po- larities called the Parker spiral (Parker 1958). The Parker model determines the direction and strength of the IMF, but the polarity of the field depends on the position with respect to the heliospheric current sheet (HCS) and on the phase of the solar cycle. We approximate the HCS as a plane that sepa- rates the regions of positive polarity from the regions of negative polarity. It is aligned with the solar equatorial plane during solar minimum, while at solar maximum it is tilted by 90◦, parallel to the solar rotation axis. As a result, the modelled HCS turns around a reference axis in the solar equatorial plane at a rate of 360◦ over 22 years. Additionally, the HCS follows the solar ro- tation. This leads to focussing and defocussing configurations of the IMF for ISD particles; see Table 1. In contrast to the approximations used in models 1 and 2 of Sterken et al. (2012), the rotation rate of the HCS is not con- stant over time. Instead, the rate is piecewise constant with a faster rate on the two intervals from solar minimum to maxi- mum, and a slower rate from solar maximum to minimum. This improves the representation of the observed rotation rates of the HCS dipole (Fig. 1). It is somewhat similar to model 3 of Sterken et al. (2012), which utilises the measured inclination of the solar magnetic dipole, but here we approximate the inclination with a monotonous function instead of the noisy data from the obser- vations to speed up the calculations, and to extrapolate into the future. We assume one sidereal rotation every 25.38 days for both the poles and the equator of the Sun. This is the Carrington ro- tation rate of the Sun which corresponds to the sidereal rotation rate at 26◦ latitude. As we average the magnetic field over the solar rotation period, the only effect of the solar rotation rate on the simulation is as a scaling factor for the azimuthal component of the solar magnetic field (Sterken et al. 2012, equation 28). We adopt the IMF description given in Parker (1958), use a mag- netic field strength of B0 = 2300 nT at 10 solar radii from the Sun (Cravens 1997), and assume a constant solar wind speed of 400 kms−1. As proposed by Landgraf (1998), the magnetic field strength is averaged over one solar rotation. This facilitates the use of larger integration time steps for faster computation. Ac- cording to our test simulations with a magnetic field following the solar rotation (similar to model 2 of Sterken et al. 2012), the error of this approximation is below a cell size for all parti- cles except those that pass within 0.25 AU of the Sun. As these particles constitute only a very small fraction in the inner solar system, we consider this a good trade-off. 2.3. Resulting simulation output The simulations produce a set of data cubes, one for each simu- lated particle size with specific β and Q/m. The cubes contain the spatial densities, velocity components, and the velocity disper- sion inside a box within 10 AU of the Sun (between −10 AU and +10 AU in each coordinate), at a spatial resolution of 0.25 AU, and a time resolution of 12.2 days. Thirteen particle sizes between 0.05 and 4.9 µm were used for calculations. (Table 2). However, for quantitative predic- Fig. 1. Tilt of the heliospheric current sheet, computed from measure- ments at the Wilcox Solar Observatory (solid line, Hoeksema 2018), and the piecewise-linear tilt used in the simulation (dashed line). Table 1. An overview of the modelled solar cycle. Year (WSO) 1976 1987 2000 2009 2013 Year (ISD model) 1974.5 1978 1985.5 1989 1996.5 2000 2007.5 2011 2018.5 2022 2029.5 2033 2040.5 Min / Max Cycle Min Max Min Max Min Max Min Max Min Max Min Max Min defocus defocus → focus focus focus → defocus defocus defocus → focus focus focus → defocus defocus defocus → focus focus focus → defocus defocus tions in this study, only seven sizes between 0.11 and 2.29 µm are used: particle sizes below and above this interval were not present with a sufficient signal-to-noise ratio (S/N) in the Ulysses dataset used for normalisation. Furthermore, there is a gap be- tween 0.1 and 0.34 µm because those particles do not reach the Earth's orbit for the β-curve we assumed. The two smallest sim- ulated particle sizes are used in Section 4 to discuss the possibil- ities of detecting such particles at Earth orbit. The normalisation of the simulated dust flux was achieved using the ISD mass distribution measured by Ulysses. Over sev- eral extended periods comprising a total of 13 years, the Ulysses dust detector identified particles of interstellar origin and mea- sured their mass distribution (Krüger et al. 2015). The simulated dust flux was extracted for the same periods and normalised such that the integrated number of impacts over the whole measure- ment period matches the Ulysses impact rates for all simulated sizes. The general ISD flow pattern and resulting flux filtering and enhancements are discussed in Sterken et al. (2012, 2013). In Appendix A we give an overview of the results of our simula- tions showing the spatial distribution of ISD grains, which varies Article number, page 3 of 17 197019801990200020102020020406080197019801990200020102020Time [yr]020406080Tilt angle [deg] A&A proofs: manuscript no. aa markedly with the solar cycle. Nevertheless, there is some sys- tematic structure that can be predicted. This will be looked at in the following sections. In Section 3 we show the variations of four parameters along the Earth's orbit over the whole solar cycle of 22 years: ISD density and velocity, the angular deflection for different parti- cle sizes, and the widening of the flow direction due to dynamic effects. All these quantities are given in the heliocentric eclip- tic reference frame (i.e. not accounting for the relative velocity of the Earth, in order to show the underlying modulation of the ISD flow). In Section 4 we show the ISD mass inflow rate on to Earth as well as the ISD count rate and impact velocity for a dust detector on an Earth-like orbit (e.g. at one of the Lagrangian points), taking into account the relative motion of the Earth. The variations of the dust flow registered at Earth are a superposition of the spatial and time variabilities of the ISD dust flow through the heliosphere and the sampling of these inhomogeneities by the Earth as it moves on its orbit. We note that the fluxes are cal- culated assuming a detector with a 4π field of view. However, as the ISD flow is collimated, the results are virtually the same for a more realistic detector geometry with an opening angle of (cid:46) 2π pointing towards the dust apex direction. 3. The ISD flow at 1 AU from the Sun Here we give an overview of the heliocentric ISD flow sam- pled at the positions of the Earth along its orbit. We discuss the time variations of ISD density, heliocentric speed, the heliocen- tric flow direction, and the spread of the flow direction for three different dust particle sizes (Table 2) representative for three dif- ferent size regimes, each dominated by a different force: Electro- magnetic force (ad = 0.072 µm), radiation pressure (0.34 µm), and gravity (0.49 µm). ISD particles with β > 1.4 cannot be ob- served at Earth orbit because the Sun's radiation pressure is too high. For the β-curve used in our simulations, this applies to par- ticle sizes of 0.15 µm < ad < 0.3 µm (Appendix. A). We compare the heliocentric flow properties of these par- ticles in Figures 2 to 5 at the Earth's positions along its orbit. For medium-sized and large particles (0.34 µm and 0.49 µm, re- spectively), sampling of different regions along the Earth's orbit is the dominant cause for the observed variations, but for small particles (ad (cid:46) 0.1 µm) the ISD flow direction is intrinsically time-dependent due to the strong modulation by the solar wind magnetic field. For particles with β < 1 the gravitational focussing leads to high-density regions downstream of the Sun, where Earth passes on 13 December every year. The point closest to the upstream direction is crossed on 12 June. Figure 2 shows the spatial density and its fluctuations over the solar cycle. Besides strong annual variations due to the de- flection by radiation pressure, the focussing due to the mag- netic field is most pronounced for the smallest particles, and is strongest around the year 2031, in agreement with other studies of ISD in the solar system (further away from the Sun). This figure also zooms in on the period of moderate elec- tromagnetic focussing between 2024 and 2027, where seasonal variations become obvious. The flux of small particles (0.07 µm) displays a strong enhancement during times of maximum fo- cussing around 2031, but no annual variation. Conversely, the flux of medium-sized particles, 0.34 µm, shows a repeated an- nual pattern. It becomes zero for several months around Decem- ber because the Earth is in the β-cone for these particles, while for a period of several months centred around June the flux is appreciable (on average ∼ 30% of the incoming flux). The dust Article number, page 4 of 17 Table 2. Particle radii ad and radiation pressure factors β used in the simulations. The force dominating the dynamics of a given size can be the radiation pressure (RP), solar gravity, and/or the electromagnetic (EM) force. Particles around 0.2 µm are not observable in the Earth's orbit because the repulsive force of the radiation pressure leads to a deflection distance rh > 1 AU from the Sun for β (cid:38) 1.4. ad [µm] 0.049 0.072 0.11 0.16 0.23 0.34 0.49 0.72 1.1 1.6 2.3 3.4 4.9 mass [kg] 1.00 × 10−18 3.16 × 10−18 1.00 × 10−17 3.16 × 10−17 1.00 × 10−16 3.16 × 10−16 1.00 × 10−15 3.16 × 10−15 1.00 × 10−14 3.16 × 10−14 1.00 × 10−13 3.16 × 10−13 1.00 × 10−12 (not observable at Earth) (not observable at Earth) dominant force β 0.55 EM 0.95 EM 1.38 RP and EM 1.59 1.49 1.17 RP and EM 0.81 0.52 0.33 0.21 0.14 0.09 0.06 gravity gravity gravity gravity gravity gravity gravity flow lines are compressed near the edge of the β-cone, leading to an increased density there. Additionally, the flux of large par- ticles is largely constant for most of the year. There is a strong maximum downstream of the Sun around December 13 as a re- sult of the gravitational focussing by the Sun, even though the Earth does not pass centrally through the focussing spot (its or- bital plane is inclined by 8◦ with respect to the ISD flow). The density in the focussing region shows a typical enhancement by a factor of ∼5, and lasts 1-2 months (Fig. 2, lower right panel, zoomed particle density at 0.49 µm). Small particles with ad (cid:46) 0.1 µm exhibit an irregular vari- ability on timescales much less than one year (Fig. 2, lower pan- els). This can be explained by the strong interaction with the magnetic field, which leads to strong concentrations on small spatial and temporal scales. However, the precise location of these density enhancements is very sensitive to Q/m, and there- fore to the fact that in our simulations we use only a single par- ticle size per bin, rather than a size distribution. With more real- istic size distributions some of these density variations are likely to average out. Figure 3 demonstrates the variation in the heliocentric speed of ISD particles at the Earth position. The heliocentric speed of the smallest particles displays strong variations around the speed of the incoming ISD flow (26 km s−1). The mid-range particles are decelerated when they reach 1 AU because of the dominant radiation pressure. The dynamics of large particles is dominated by solar gravity; at 1AU, particles with sizes of ad = 0.49 µm (ad (cid:38) 0.72 µm) experience a net acceleration to 30 km s−1 (40 km s−1, respectively). In Fig. 4 we show the inflow direction in terms of ecliptic longitude/latitude. The ISD flow of specific sizes is not highly collimated. Within a volume ele- ment of 0.25 AU on the side, the trajectories of ISD particles vary along the Earth's orbit by up to 30◦. Due to solar gravity and ra- diation pressure, the largest directional variation inside a single mesh cell can be seen downstream of the Sun where particles become focussed and reach the Earth's orbit from a wide range of directions after experiencing a strong gravitational focussing (Fig. 5). Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Figure 6 shows the deflection angle relative to the inflow direction of ISD particles at the Earth's position (in the eclip- tic reference frame) over time. The flow of the small particles is mostly irregular, whereas the larger particles (0.34 µm and 0.49 µm) generally follow a yearly modulation. Close to the downstream focussing spot, the deflection is largest because here the effects of gravitational focussing are strongest (Sterken et al. 2013). However, for particles of ad = 0.49 µm, the deflection drops steeply inside the focussing spot. This is due to the fact that trajectories from different directions hit the same spot and the de- flections cancel out on average. This is reflected in the increase in width of the ISD flow towards the focussing spot (Fig. 5). 4. The ISD flow as seen by an observer at Earth We discuss the simulated flow patterns of the ISD near the Earth or near a spacecraft in an orbit around the Sun at 1 AU. The flow characteristics are modulated by the velocity of the Earth or the spacecraft around the Sun (∼ 30 km s−1) in addition to the ef- fects discussed earlier in this paper. In particular, we concentrate on the impact speed distribution, the inflow direction, the mass density distributions, and the total mass flux on to Earth. We ig- nore the effects of the Earth's gravity and of the Earth's magnetic field. 4.1. Speed distribution and relative flow directions The impact velocity distribution is required to predict the parti- cle flux and to determine the speed range in which the particles can be observed at the Earth/spacecraft. Figure 7 shows the av- erage velocity of ISD particles with respect to the Earth for in- dividual particle sizes. The impact velocity is modulated by two different effects: the exact one-year cycle due to the Earth's or- bit (including Earth's velocity); and the time-dependent effects of the IMF. The speed in the ecliptic reference frame (Fig. 3) only varies strongly for particles of size ad < 0.1 µm. Therefore, the amplitude of the Earth's relative velocity dominates the an- nual variations in the impact speed of the larger particles. When the velocity vectors of ISD particles and the Earth are parallel, the value of the relative velocity vector becomes small (a few km s−1) in contrast to the situation half a year later when both velocity vectors are anti-parallel and the relative speed becomes 50 to 60 km s−1. For particles of 0.49 µm in size, the highest im- pact speed is approximately 60 km s−1 in December each year. Figure 8 shows the inflow direction, in the Earth's reference frame, taking into account the velocity owing to the Earth's or- bital motion. This direction can vary over a full circle (360◦) (cf. Fig. 4). This can be used to identify ISD particles and distinguish them from other dust populations. 4.2. Mass flux on to Earth and mass distribution The size distribution of ISD particles is also useful for the char- acterisation of ISD at the Earth and the resulting mass flux into the Earth's atmosphere. The mass distribution at the Earth varies widely during the course of a year. Figure 9 shows the positions of the Earth that are used to demonstrate the seasonal changes in the mass distribution. The seasonal mass distributions of the ISD flow at Earth orbit are shown in Fig. 10. Variations of the mass distribution at Earth are shown for a period of 1 year, in a defo- cussing configuration (2018-2019) and in a focussing configura- tion of the IMF (2029-2030). As the mean IMF is largely con- stant over a period of this length, the variations reflect changes in the spatial distribution of the ISD particles due to radiation pres- sure and gravity, as well as the relative velocity observed in the Earth's frame of reference. During the upstream portion of the Earth's motion, small particles do not reach the Earth (β-cone effect), while during the downstream portion large particles are focussed and their flux is enhanced. In addition to this modula- tion there is the focussing/defocussing effect for small particles (< 0.3 µm) due to the Lorentz force. Figure 11 shows the annual variations over the 22-year so- lar cycle at four different positions along the Earth's orbit. This illustrates the varying effect of the changing IMF on the mass distributions. The overall patterns of the dust flux are similar to those observable at the orbits of Jupiter and Saturn and in the asteroid belt (Sterken et al. 2013), except for the β-cone that af- fects more particles at Earth orbit than further away from the Sun. In addition, due to the orbit of the Earth, the frequency and the amplitude of the modulations are different. The cumulative mass flux (summed over all masses) on to the Earth reflects this orbital variation (Fig. 12). The contributions from different particle sizes change during the course of the year. The strong peak at the end of the year is caused by the large particles that are focussed downstream while at other times of the year smaller particles dominate the mass flux. 4.3. Observing conditions for spacecraft at Earth's orbit How can we verify the simulation results of this new model for ISD? Recent in-situ measurements of ISD at 1 AU were per- formed primarily by Cassini (Altobelli et al. 2003), but observa- tion time was limited to a few weeks. New mission concepts are in development, which allow dust observations for many years. The Destiny+ mission is currently in development with a fore- seen launch date of 2022 (Arai et al. 2018). It will carry a mod- ern dust telescope to investigate ISD, interplanetary dust, and dust of the active asteroid Phaethon (Kobayashi et al. 2018). The Destiny+ spacecraft will have an Earth-like interplanetary tra- jectory (Sarli et al. 2018). Predictions of the dust flow during the Destiny+ mission are shown in Krüger et al. (2018b). Here we discuss two distinct populations of particles that can be observed at Earth orbit (Fig. 13): the larger particles (ad (cid:38) 0.3 µm, corresponding to a particle mass of m (cid:38) 3 × 10−16 kg), whose dynamics are dominated by gravity and radiation pres- sure; and the small particles (ad (cid:46) 0.2 µm), whose dynamics are dominated by the Lorentz force of the IMF. While there is a tem- poral variation caused by the solar cycle for the smaller particles, the density of the larger particles with β < 1 never drops to zero (Fig. 2). However, the flux of dust particles on to a spacecraft at Earth orbit can drop to small values when their velocity vec- tors become similar (Fig. 13). For most of the orbit the flux is relatively constant, with a sharp increase when the Earth crosses the gravitational focussing cone downstream of the Sun. At 1 AU the mean interplanetary dust flux of ad (cid:46) 0.3 µm particles is 5 day−1m−2 (Grün et al. 1985). The mean ISD flux is there- fore 10% to 80% of the interplanetary flux, depending on the focussing conditions of the IMF. Conversely, the small particles that are dominated by the Lorentz-force can only be observed during the strong focussing configuration of the solar magnetic field in the years after so- lar minimum. During other time periods, the ISD particles are strongly depleted by the defocussing magnetic field, and their flux drops to zero. They were strongly depleted in the Ulysses data for a number of reasons: The interval of maximum fo- cussing was just outside the 17 years of the Ulysses mission, and the intervals closest to the maximum focussing were strongly Article number, page 5 of 17 A&A proofs: manuscript no. aa Fig. 2. Simulated relative ISD density (relative to the density of the undisturbed ISD flow) at Earth's position (top row), and the same figures zoomed in to a range of 3 years (bottom row). Fig. 3. Simulated heliocentric dust speed at Earth's position in the ecliptic reference frame. We note the different y-axis scalings. contaminated by stream particles ejected from the Jovian sys- tem, severely hindering the identification of small particles from a potential ISD population (Strub et al. 2015). 4.4. Limitations of the model Our model of the ISD flow is subject to a number of limita- tions as a result of (1) uncertainty in our understanding of the physical properties of ISD inside the solar system (i.e. compo- sition, porosity, and mass; and therefore different radiation pres- sure effects), (2) the assumption of a constant electric charge, (3) limited information concerning the actual three-dimensional structure of the IMF on smaller scales (e.g. coronal mass ejec- tions, CMEs), (4) practical considerations of the computation time needed, such as the averaging of the magnetic field av- eraged over the solar rotation period, and (5) uncertainties in the absolute normalisation of the simulated fluxes based on the Ulysses measurements. The uncertainty of physical properties (1) leads to a number of consequences that have been discussed in detail by Sterken et al. (2013): a change in material properties would lead to a dif- ferent β-curve, which in turn modifies the dynamics of particles of a given size. This can be addressed in the future by running further simulations in order to cover a larger area in the parame- ter space of β and mass (i.e. Q/m), as has been done in Sterken et al. (2012) and Sterken et al. (2013) at lower resolution. The Article number, page 6 of 17 absolute flux in our ISD model was normalised by the mass cal- ibration of the Ulysses dust detector. Effects of particle structure and particle density on the mass calibration are not considered here. The approximation (2) of a constant charge of ISD particles is discussed and justified in Horányi (1996) and in Slavin et al. (2012), Fig. 2, for distances between one and several tens of as- tronomical units from the Sun. In the case of small variations in the magnetic field, the in- fluence of (3) on the simulation result is small, as the coupling scale of ISD particles with sizes ad (cid:38) 0.1 µm is of the order of (cid:38)1-10 AU, and can be neglected for our simulation. Due to the stochastical nature of (3), the effects of smaller-scale structures most likely cancel out on average. Larger deviations from the Parker IMF, however, as can be found inside CMEs, can lead to severe deflections from the simulated flux. Assuming a mag- netic field of 30 nT, as observed inside individual CMEs (Wang et al. 2005; O'Brien et al. 2018), this is six times higher than the modelled average solar magnetic field at Earth orbit of 5 nT, and leads to a typical gyro radius of 1 AU for 0.1 µm particles. The simulation results for small particles (cid:46) 0.1 µm should therefore be treated with caution and are valid for the nominal Parker IMF only. Due to their stochastical nature, the occurrence of CMEs cannot be predicted, and are beyond the scope of this model. We have verified the validity of (4) by numerically inte- grating the trajectories of test particles in a rotating magnetic 201520202025203020352040024681012 radius 0.072mu201520202025203020352040Year024681012Rel. dust density [1]2015202020252030203520400246810 radius 0.335mu201520202025203020352040Year0246810Rel. dust density [1]201520202025203020352040051015 radius 0.492mu201520202025203020352040Year051015Rel. dust density [1]2024.02024.52025.02025.52026.02026.52027.00.00.10.20.30.40.5 radius 0.072mu2024.02024.52025.02025.52026.02026.52027.0Year0.00.10.20.30.40.5Rel. dust density [1]2024.02024.52025.02025.52026.02026.52027.00.00.51.01.52.0 radius 0.335mu2024.02024.52025.02025.52026.02026.52027.0Year0.00.51.01.52.0Rel. dust density [1]2024.02024.52025.02025.52026.02026.52027.00246810 radius 0.492mu2024.02024.52025.02025.52026.02026.52027.0Year0246810Rel. dust density [1]201520202025203020352040010203040radius 0.072mu201520202025203020352040Year010203040abs. velocity [km/s]2015202020252030203520400510152025radius 0.335mu201520202025203020352040Year0510152025abs. velocity [km/s]201520202025203020352040010203040radius 0.492mu201520202025203020352040Year010203040abs. velocity [km/s] Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Fig. 4. Flow direction in ecliptic coordinates (l, b) of ISD particles at Earth's position in the heliocentric ecliptic reference frame for particle sizes 0.07 µm, 0.34 µm, 0.49 µm, and 0.72 µm. The solid horizontal lines correspond to the incoming flow direction. field. The resulting differences compared to an averaged mag- netic field are below 0.1 AU, which is smaller than the mesh resolution of the density cube, for all but the particles passing within less than 0.25 AU of the Sun, which constitutes only a negligible fraction of the overall ISD flux. Concerning the normalisation of the simulated fluxes (5), we note that the temporal variability of the flux and the direction in the Ulysses ISD dataset are not entirely reproduced by the model, and only the overall flux for each particle size bin is taken into account for the normalisation. Article number, page 7 of 17 20202025203020352040150200250300350radius 0.072mu20202025203020352040Year150200250300350ecl. longitude [deg]20202025203020352040−60−40−200204060radius 0.072mu20202025203020352040Year−60−40−200204060ecl. latitude [deg]20202025203020352040150200250300350radius 0.335mu20202025203020352040Year150200250300350ecl. longitude [deg]20202025203020352040−60−40−200204060radius 0.335mu20202025203020352040Year−60−40−200204060ecl. latitude [deg]20202025203020352040150200250300350radius 0.492mu20202025203020352040Year150200250300350ecl. longitude [deg]20202025203020352040−60−40−200204060radius 0.492mu20202025203020352040Year−60−40−200204060ecl. latitude [deg]20202025203020352040150200250300350radius 0.723mu20202025203020352040Year150200250300350ecl. longitude [deg]20202025203020352040−60−40−200204060radius 0.723mu20202025203020352040Year−60−40−200204060ecl. latitude [deg] A&A proofs: manuscript no. aa Fig. 5. 1σ stream width at Earth position in the heliocentric reference frame. Fig. 6. Deflection angle of ISD particles at Earth orbit with respect to the ISD inflow direction into the solar system. Fig. 7. Average velocity of particles in the Earth's reference frame for different selections of particle sizes (from left to right: 0.07 µm, 0.34 µm, 0.49 µm). The Ulysses ISD data were chosen as the calibration dataset because this dataset contains the most comprehensive homoge- neous observation by a single instrument over a period of 16 years, and covers a large portion of the 22 year solar cycle (Strub et al. 2015). With a total of 987 ISD particles used for determin- ing the mass distribution (Krüger et al. 2015), it also has the best statistical errors of all observations. However, Krüger et al. (2018a) compared the model predictions to flux measurements from other missions such as Helios, Cassini, and Galileo, and they agree within a factor of typically (cid:46) 2 − 3, despite the differ- ences in solar distance covered by these missions. We conclude that this marks the limits of the current understanding of the ISD flow through the heliosphere. 5. Scenarios for interstellar dust missions The investigation of ISD particles requires a statistical signifi- cant sample in order to study the dynamics, mass distribution and composition over time and space. What is the composition of the ISD particles and how did they form? What are their dy- namical parameters, and what can be learned from a test and re- finement of the model based on improved measurements? How Article number, page 8 of 17 do ISD particles interact with the heliosphere, and how does their dynamics and overall flux modulation depend on the location in our solar system? The ISD dynamics and flux described above allow for a detailed study of the ISD as long as observational campaigns are carefully planned. This is a lesson learned from the Cassini mission, where a special observation campaign in 1999 led to the discovery of interstellar particles as close as 1 AU to the Sun (Altobelli 2004). Later, only the long integration time during ISD campaigns in Saturn's orbit allowed for an analysis of the composition of 36 ISD particles (Altobelli et al. 2016). In order to plan observational strategies, the modelling and prediction of ISD particle properties and their variations over time and space are essential. What is the best observational strategy and what are the requirements for a mission close to Earth's orbit? What is the ideal pointing scenario and what are the related observational times? Our model indicates that observations at 1 AU are well suited to study the ISD in the solar system including its varying dy- namics during different phases of the solar cycle. This can be achieved on a high Earth orbit, but a lunar orbit or an orbit about 201520202025203020352040020406080100radius 0.072mu201520202025203020352040Year020406080100width 1sigma angle [deg]2015202020252030203520400204060radius 0.335mu201520202025203020352040Year0204060width 1sigma angle [deg]201520202025203020352040010203040radius 0.492mu201520202025203020352040Year010203040width 1sigma angle [deg]201520202025203020352040020406080100120radius 0.072mu201520202025203020352040Year020406080100120deflection angle [deg]201520202025203020352040020406080100radius 0.335mu201520202025203020352040Year020406080100deflection angle [deg]201520202025203020352040010203040radius 0.492mu201520202025203020352040Year010203040deflection angle [deg]20282030203220342036203820400204060radius 0.072mu2028203020322034203620382040Year0204060rel. velocity [km/s]20282030203220342036203820400204060radius 0.335mu2028203020322034203620382040Year0204060rel. velocity [km/s]20282030203220342036203820400204060radius 0.492mu2028203020322034203620382040Year0204060rel. velocity [km/s] Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Fig. 8. Apparent apex direction in ecliptic coordinates (λ, β) of ISD particles in the Earth reference frame (taking into account the orbital velocity of the Earth) for particle sizes 0.07 µm, 0.34 µm, 0.49 µm, and 0.72 µm. the libration point L2 of the Earth-Sun system would be viable options as well. As discussed in Sect. 4, the Earth speed of about 30 km s−1 leads to a strong modulation of the ISD flow which has a typical velocity of 26 km s−1. This strongly affects both the registered dust flux and the impact velocity along the Earth's orbit. These high-amplitude variations in flux and velocity can be used to dis- tinguish the interplanetary from the ISD population (Grün et al. 2009). Figure 14 schematically shows the Earth's orbit with the ISD observatory at the libration point L2. Highest fluxes and im- Article number, page 9 of 17 202020252030203520400100200300radius 0.072mu20202025203020352040Year0100200300ecl. longitude [deg]20202025203020352040−50050radius 0.072mu20202025203020352040Year−50050ecl. latitude [deg]202020252030203520400100200300radius 0.335mu20202025203020352040Year0100200300ecl. longitude [deg]20202025203020352040−50050radius 0.335mu20202025203020352040Year−50050ecl. latitude [deg]202020252030203520400100200300radius 0.492mu20202025203020352040Year0100200300ecl. longitude [deg]20202025203020352040−50050radius 0.492mu20202025203020352040Year−50050ecl. latitude [deg]202020252030203520400100200300radius 0.723mu20202025203020352040Year0100200300ecl. longitude [deg]20202025203020352040−50050radius 0.723mu20202025203020352040Year−50050ecl. latitude [deg] A&A proofs: manuscript no. aa particles of size ad > 0.3 µm at 1 AU ranges between approx- imately 0.01 m−2 day−1 (1.3 × 10−7 m−2 s−1) and 20 m−2 day−1 (2.3 × 10−4 m−2 s−1), which is only a fraction of the flux of in- terplanetary particles. Therefore, accurate trajectory information is necessary in order to reliably distinguish both types of dust particles. The following considerations assume a sensitive detector area of 0.1 m2, which is the same as in the dust detectors on board Ulysses and Galileo. The boresight however is assumed to point constantly towards the ISD flow. This is equivalent to a 4π field of view with a constant sensitive area. For a given dust flow, this leads to approximately three times higher dust fluxes than the rotating scanning pattern used by Ulysses. The year-integrated flux of large particles with ad (cid:38) 0.3 µm (i.e. above the β-cone cutoff size) exhibits marked variations by a factor of approximately seven with the (de-)focussing cycle of the solar magnetic field, and reaches a maximum of 116 particles in 2030, and a minimum of 16 particles in 2041. We note that due to the steep mass distribution these particle counts are dominated by the number in the smallest size bin. Small dust particles with ad (cid:46) 0.11 µm are primarily observ- able at Earth's orbit close to the phase of strongest focussing. A number of 2610 particles is expected in 2030 which is maximum focussing. In-situ observations allow us to link optical particle properties with dynamical properties. For a limited range of 1 < β < 1.4, a direct measurement of the β-curve would be possible through β-spectroscopy (Altobelli et al. 2005). For a given particle mass, the size of the β-cone (the zone devoid of particles) largely depends on β. Therefore, by measuring the size of this exclusion zone for different masses, the β-mass-relation, a determining factor for the ISD dynamics, can be established. With new instrumentation, the mass distribution of the ISD can be measured with a better accuracy. At the moment, this is a major uncertainty in our model: The normalisation of the incom- ing flux is based on the mass distribution measured by Ulysses, which is limited by relatively low number counts per bin. A multi-year mission or an increased sensitive area (> 0.1m2) is therefore required in order to analyse a sufficient number of ISD particles. A new mission may also provide much higher sensi- tivities for small particles. During a period of electromagnetic focussing, a dust instrument could possibly measure ISD parti- cles (cid:46) 0.1µm if it is able to distinguish them from sub-micron sized interplanetary dust particles. If successful, such extra mea- surements may strongly support understanding the filtering of ISD - especially in the boundary regions of the heliosphere (he- liosheath, etc.). One significant point along the orbit is the gravitational fo- cussing region downstream of the Sun. There the density and flux of large ISD particles are about one order of magnitude higher than the typical values along the rest of the orbit. This would provide an important opportunity to detect larger ISD particles with β < 1. The simulation results allow for detailed investiga- tions of the required ISD ram direction, which deviates from the direct Sun direction. A nanodust analyser (NDA) has been developed for the de- tection and compositional analysis of dust particles originating in the inner heliosphere (O'Brien et al. 2014). The NDA is the first instrument optimised for the detection and compositional analysis of nanodust particles that are arriving from close to the Sun's direction. The NDA is a linear time-of-flight impact mass spectrometer derived from the successful Cassini CDA instru- ment. The operation and performance of the NDA instrument, Fig. 9. Geometry of the Earth's orbit and points used for extracting the simulated mass distributions (see Fig. 10). The arrows indicate the undisturbed flow direction of ISD particles for reference, i.e. the flow direction of the particles outside the solar system. The colours of the dates along the orbit match the colours used in Fig. 10. pact velocities are measured when the Earth moves anti-parallel to the ISD flow (left), and lowest fluxes and impact speeds occur when the spacecraft moves parallel to the ISD direction (right). As the in-situ detectors typically have large but limited opening angles, the simulated flow directions can be used to optimise ob- servation campaigns and to align the instrument boresight to the ISD ram direction (cf. Fig. 8). Generally, two different methods are available for ISD obser- vations: in-situ measurements, and sample return with a subse- quent in-depth laboratory analysis of the samples here on Earth. Both options require different operational scenarios. In the past several in-situ missions serendipitously detected and analysed ISD: Ulysses, Galileo, Cassini and Helios. The latter two pro- vided the first compositional analysis of ISD grains (Altobelli et al. 2006, 2016). Sample return and subsequent analysis can make use of the latest laboratory techniques that are not avail- able for space instrumentation, promising very accurate results. The Stardust mission is an example of such a strategy, and suc- cessfully brought back ISD samples to the Earth for laboratory analysis (Westphal et al. 2014). However, the laboratory analy- sis took many years to complete because the process to identify a few submicron-sized particles on a 0.1 m2 collector is very chal- lenging. We briefly discuss both mission scenarios. 5.1. In-situ observations Reliable in-situ dust instrumentation is available for the analy- sis of dust fluxes, masses, electrical charges and, in particular, composition (Auer et al. 2002; Auer 2001; Srama et al. 2004; Sternovsky et al. 2007; Krüger et al. 2017). The instruments are typically scalable in size to allow for the large sensitive ar- eas that are needed to detect a statistically meaningful number of particles. As can be seen in Fig. 13, the flux of interstellar Article number, page 10 of 17 −2−1012−2−1012−2−1012ecl. X [AU]−2−1012ecl. Y [AU] 2018.90 2019.00 2019.10 2019.20 2019.30 2019.40 2018.50 2018.60 2018.70 2018.80 Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Fig. 10. Mass distribution in the ecliptic frame of reference over the course of one year, for a defocussing configuration of the ISM (left panel, 2018-2019) and for a focussing configuration of the ISM (right panel, 2029-2030). Since the 3-dimensional flow configuration is virtually constant over these periods of 1 year, the observed changes reflect the spatial variations of the mass distribution along the Earth orbit, in a similar way to the mass distributions for Saturn, Jupiter and asteroid distances in Sterken et al. (2013). Fig. 11. Mass distribution over the whole 22-year solar cycle at four given points along the Earth's orbit (compare to Sterken et al. (2013)). Upper left panel: upstream of the Sun; Upper right panel: at an angle of 90 degrees between upstream- and downstream position; Lower left panel: downstream of the Sun, inside the spot of maximum focussing; Lower right panel: 18◦ away from the downstream downstream direction, outside the strongly localised spot of maximum focussing. The variations seen in each panel reflect the changes due to the focussing/defocussing effects of the IMF throughout the solar cycle. including the efficiency of the solar wind rejection grids, have been verified in the laboratory using a dust accelerator and a so- lar wind simulator. This instrument is well suited to analysing the down-stream focused ISD flux from the solar direction. 5.2. Dust sample return The second method to analyse ISD particles is dust collection with sample return. While Stardust has collected and brought back to Earth some ISD, the process to identify the few submi- cron sized dust particles on a 0.1 m2 collector was very challeng- Article number, page 11 of 17 10−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2018.502018.602018.702018.802018.902019.002019.102019.202019.302019.4010−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2029.502029.602029.702029.802029.902030.002030.102030.202030.302030.4010−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2018.452020.452022.452024.452026.452028.452030.452032.452034.452036.452038.4510−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2018.702020.702022.702024.702026.702028.702030.702032.702034.702036.702038.7010−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2018.952020.952022.952024.952026.952028.952030.952032.952034.952036.952038.9510−1810−1610−1410−1210−100.010.101.0010.0010−1810−1610−1410−1210−10Mass [kg]0.010.101.0010.00Dust density [1/m^3]2018.902020.902022.902024.902026.902028.902030.902032.902034.902036.902038.90 A&A proofs: manuscript no. aa Fig. 12. Mass flow on to Earth for all particle sizes. The mass flow takes into account the Earth's velocity relative to the ISD flow and is therefore given in the reference frame of the Earth. Fig. 13. Number flow of the ISD particles over a full solar cycle (2018-2040). Left panel: Small particles (ad < 0.2 µm). Right panel: Large particles, (ad > 0.3 µm). Fluxes are given in particles day−1 m−2 assuming the detection area is always orthogonal to the ISD flow. The smooth line shows the moving average of the flux over a period of 1 year. The number flux can reach values close to zero when the Earth and ISD velocity vectors become similar. ing (Westphal et al. 2014). An active cosmic dust collector (Grün et al. 2012) would tremendously improve future dust collections in interplanetary space by determining the impact position and time together with the velocity vector of the impacting particle. A sample return of interstellar matter mission (SARIM) was described by Srama et al. (2009). In this scenario the spacecraft would be placed at the L2 libration point of the Sun/Earth sys- tem, outside the Earth's debris belts and inside the solar-wind charging environment. SARIM is a three-axes stabilised space- craft and collects interstellar particles between July and October when the relative encounter speeds with ISD particles are low- est (4 to 20 km s−1, Fig. 14). Active dust collectors with a to- tal sensitive area of 1 m2 determine the trajectory, speed, mass, and impact location of individual dust impacts. This allows for a discrimination between interstellar and interplanetary dust parti- cles. During a three-year dust collection period, several hundred interstellar and several thousand interplanetary particles can be collected by such a detector. At the end of the collection phase, collector modules would be stored and sealed in a sample return capsule. The probe with the capsule would then return to Earth and an in-depth dust analysis with advanced laboratory methods would then be possible. flow is small compared to the influx of ∼ 30 000 t yr−1 of inter- planetary dust on to the Earth (Love & Brownlee 1993). How- ever, there is significant interest in this exotic material since in- terstellar particles are the raw material from which a protoplan- etary disk and subsequently planets, asteroids and comets form. The flow of ISD particles near the Earth is strongly modu- lated by the Earth's orbit around the Sun and by the solar cy- cle variation of the interplanetary magnetic field. While the flow pattern generated by gravity and the radiation pressure are con- stant in time, the solar magnetic field leads to significant varia- tions in flux, direction, velocity, and size distribution with a pe- riodicity of 22 years. This must be considered when planning observations. The identification of ISD and its distinction from interplanetary dust can be accomplished by using modern dust instruments that characterise both the dynamical state as well as the composition of the analysed particles. Acknowledgements. This work was ESA contract funded 4000106316/12/NL/MV -- IMEX. Wilcox Solar Observatory data used in this study was obtained via the web site http://wso.stanford.edu at 2017:10:04_05:08:21 PDT courtesy of J.T. Hoeksema. We would like to thank our referee, Mihály Horányi, for his valuable comments helping to improve the presentation of our results. under 6. Conclusions The Earth is currently collecting an average of about 100 kg per year of interstellar material in the form of ISD particles. This References Altobelli, N. 2004, PhD thesis, Ruprecht-Karls-Universität Heidelberg Altobelli, N., Grün, E., & Landgraf, M. 2006, A&A, 448, 243 Article number, page 12 of 17 202020252030203520400.00.20.40.60.8traj.: Earth, size = all20202025203020352040Year0.00.20.40.60.8Mass flow onto Earth [kg/day]2025.02025.52026.02026.52027.00.00.10.20.30.40.5traj.: Earth, size = all2025.02025.52026.02026.52027.0Year0.00.10.20.30.40.5Mass flow onto Earth [kg/day]202020252030203520400100200300400 20202025203020352040Year0100200300400number flux [#/day/m2]2020202520302035204005101520 20202025203020352040Year05101520number flux [#/day/m2] Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Landgraf, M. 2000a, J. Geophys. Res., 105, 10303 Landgraf, M. 2000b, J. Geophys. Res., 105, no. A5, 10,303 Landgraf, M., Augustsson, K., Grün, E., & Gustafson, B. A. S. 1999, Science, 286, 2319 85, 035113 156, 7 Linde, T. J. & Gombosi, T. I. 2000, J. Geophys. Res., 105, 10411 Love, S. G. & Brownlee, D. E. 1993, Science, 262, 550 Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425 Morfill, G. E. & Grün, E. 1979, Planet. Space Sci., 27, 1269 Mukai, T. 1981, A&A, 99, 1 O'Brien, L., Auer, S., Gemer, A., et al. 2014, Review of Scientific Instruments, O'Brien, L., Juhász, A., Sternovsky, Z., & Horányi, M. 2018, Planet. Space Sci., Parker, E. N. 1958, ApJ, 128, 664 Sarli, B. V., Horikawa, M., Yam, C. H., Kawakatsu, Y., & Yamamoto, T. 2018, Journal of the Astronautical Sciences, 65, 82 Slavin, J. D., Frisch, P. C., Müller, H.-R., et al. 2012, ApJ, 760, 46 Srama, R., Ahrens, T. J. Altobelli, N., Auer, S., et al. 2004, Space Sci. Rev., 114, 465 141 Srama, R., Stephan, T., Grün, E., et al. 2009, Experimental Astronomy, 23, 303 Sterken, V. J., Altobelli, N., Kempf, S., et al. 2013, A&A, 552, A130 Sterken, V. J., Altobelli, N., Kempf, S., et al. 2012, A&A, 538, A102 Sterken, V. J., Strub, P., Krüger, H., von Steiger, R., & Frisch, P. 2015, ApJ, 812, Sternovsky, Z., Amyx, K., & Bano, G. e. a. 2007, Rev. Sci. Instrum., 78 Strub, P., Krüger, H., & Sterken, V. J. 2015, ApJ, 812, 140 Wang, C., Du, D., & Richardson, J. D. 2005, Journal of Geophysical Research (Space Physics), 110, A10107 Weingartner, J. C. & Draine, B. T. 2001, ApJ, 548, 296 Westphal, A. J., Stroud, R. M., Bechtel, H. A., et al. 2014, in Lunar and Plan- etary Inst. Technical Report, Vol. 45, Lunar and Planetary Institute Science Conference Abstracts, 2269 Witte, M., Rosenbauer, H., Banaszkiewicz, M., & Fahr, H. 1993, Advances in Space Research, 13, 121 Wood, B. E., Müller, H.-R., & Witte, M. 2015, ApJ, 801, 62 Zubko, V., Dwek, E., & Arendt, R. G. 2004, ApJS, 152, 211 Fig. 14. Mission concept for the characterisation of galactic and inter- planetary dust. The spacecraft is located in the Lagrange point L2 of the Earth-Sun system. Two positions of the Earth and the satellite are shown (spring and autumn). In spring and summer the relative impact speed and the related dust fluxes of ISD are high. Altobelli, N., Kempf, S., Krüger, H., et al. 2005, J. Geophys. Res., 110, 7102 Altobelli, N., Kempf, S., Landgraf, M., et al. 2003, J. Geophys. Res., 108, A10, Altobelli, N., Postberg, F., Fiege, K., et al. 2016, Science, 352, 312 Arai, T., Kobayashi, M., Ishibashi, K., et al. 2018, in Lunar and Planetary In- stitute Science Conference Abstracts, Vol. 49, Lunar and Planetary Institute Science Conference Abstracts, 2570 Auer, S. 2001, in Interplanetary Dust, ed. E. Grün, B. A. S. Gustafson, S. F. Dermott, & H. Fechtig (Springer Verlag, Berlin Heidelberg New York), 385 -- 444 Auer, S., Grün, E., Srama, R., Kempf, S., & Auer, R. 2002, Planet. Space Sci., 7 50, 773 553 Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1 Cravens, T. E. 1997, Physics of Solar System Plasmas, Cambridge Atmospheric and Space Science Series (Cambridge University Press) Draine, B. T. & Lee, H. M. 1984, ApJ, 285, 89 Frisch, P. C., Dorschner, J., Geiss, J., et al. 1999a, ApJ, 525, 492 Frisch, P. C., Dorschner, J. M., Geiss, J., et al. 1999b, Astrophys. J., 525, 492 Grün, E., Gustafson, B. E., Mann, I., et al. 1994, A&A, 286, 915 Grün, E., Srama, R., Altobelli, N., et al. 2009, Experimental Astronomy, 23, 981 Grün, E., Sternovsky, Z., Horanyi, M., et al. 2012, Planet. Space Sci., 60, 261 Grün, E., Zook, H., Fechtig, H., & Giese, R. 1985, Icarus (ISSN 0019-1035), vol.62, May 1985, p.244-272., 62, 244 Grün, E., Zook, H. A., Baguhl, M., et al. 1993, Nature, 362, 428 Gustafson, B. A. S. 1994, Annual Review of Earth and Planetary Sciences, 22, Gustafson, B. S. 1994, Ann. Rev. Earth Planet. Sci., 22, 553 Hoeksema, J. 2018, Wilcox Solar Observatory, http://wso.stanford.edu Horányi, M. 1996, ARA&A, 34, 383 Kempf, S., Srama, R., Altobelli, N., et al. 2004, Icarus, 171, 317 Kimura, H. & Mann, I. 1998, ApJ, 499, 454 Kobayashi, M., Srama, R., Krüger, H., Arai, T., & Kimura, H. 2018, in Lunar and Planetary Institute Science Conference Abstracts, Vol. 49, Lunar and Plane- tary Institute Science Conference Abstracts, 2050 Krüger, H., Altobelli, N., Strub, P., et al. 2018a, A&A, submitted Krüger, H., Kobayashi, M., Arai, T., et al. 2017, European Planetary Science Congress, 11, EPSC2017 Krüger, H., Strub, P., Grün, E., & Sterken, V. J. 2015, ApJ, 812, 139 Krüger, H., Strub, P., Srama, R., et al. 2018b, Planet. Space Sci., submitted Lallement, R. & Bertaux, J. L. 2014, A&A, 565, A41 Landgraf, M. 1998, PhD thesis, Ruprecht-Karls-Universität Heidelberg Article number, page 13 of 17 30 km/s26 km/sGALACTIC DUSTsummerwinterγ Appendix A: The ISD flow through the inner planetary system A&A proofs: manuscript no. aa Our simulations give a detailed insight into the flow properties of the ISD in the inner solar system within 4 AU of the Sun. The resulting flow pattern is a combination of the effects of so- lar gravity and radiation pressure (β) and the Lorentz force from the solar magnetic field. The relative importance of these forces depends on a particle's β and Q/m (Table 2). We characterise the results using a selection of plots for three different sized particles (0.07 µm, 0.34 µm, and 0.49 µm, assuming the adapted astronomical silicates curve) representing the Lorentz-force- dominated, radiation pressure, and gravity-dominated regimes. We show cuts through the three-dimensional density cubes along the ecliptic Cartesian coordinate planes (y-z, x-z, and x-y) at different epochs. The two-dimensional cuts through the three- dimensional ISD density (Fig. A.1 to A.3) demonstrate the rich and clumpy structure of the distribution of interstellar material throughout the inner solar system, which varies markedly with the solar cycle and particle size.2 Figure A1 shows the densities of 0.07 µm ISD particles. Dur- ing electromagnetically defocussing conditions in 2020 no such small particles reach the region within 4 AU from the Sun; whereas strong concentrations of ISD particles are found dur- ing focussing conditions in 2030. During the transitional phase (e.g. 2036) ISD concentrations are found at higher (or lower) latitudes. The density distributions of 0.34 µm ISD particles are shown in Figure A2. For these particles the repulsion by solar radiation pressure is the dominant force (β = 1.17); hence, an exclusion zone around the Sun (β cone) is formed. However, the electromagnetic forces are still strong enough to prevent this type of particle from reaching the inner solar system during defo- cussing conditions (2020) and to strongly focus the flux towards the ecliptic plane where interstellar particles flow around a flat- tened β-cone (x-y plane in 2030, cf. Fig. 45-46 in Sterken et al. (2012)). During the transitional phases (2036), ISD particles are concentrated at the edge of the β-cone at higher and lower lati- tudes. Solar gravity is the strongest force for 0.49 µm particles, hence a pronounced focussing region is observed downstream from the Sun. Nevertheless, some focussing (2030) and defo- cussing effects (2020) are observed. The Earth moves around the Sun on a roughly circular orbit at 1AU radius, thereby probing different regions of the ISD flow. This causes an annual variation of apparent flow direction and its deviation (deflection angle) from the upstream inflow direction (ecliptic longitude/ latitude lecl = 259◦, becl = 8◦) which is most pronounced for medium-sized and large particles (0.34 µm and 0.49 µm, respectively, Fig. A3). The solar magnetic field changes periodically with a 22-year cycle, with a polarity flip every 11 years; its effects are therefore intrinsically time-dependent. In the years 2014-2025, and 2036 onwards, the result is an overall defocussing of the ISD flow with a depletion of ISD in the inner solar system. In contrast, a focussing towards the solar equatorial plane occurs in the years 2025-2036, leading to an increased dust density. As the force depends on Q/m, the overall effect is mostly seen in the small particles ad (cid:46) 0.1 µm. Particles in the range 0.2 µm (cid:46) ad (cid:46) 0.4 µm are moderately affected, whereas particles with radii ad (cid:38) 0.4 µm do not respond significantly to the Lorentz force. 2 Simulations using smaller cell sizes and/or a randomisation of Q/m in a small interval might reduce some clumpy patterns. Article number, page 14 of 17 Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Fig. A.1. Cuts along the ecliptic coordinate planes through the simulated density cubes for particles of size 0.072 µm at different epochs. The years have been chosen to select a representative case in the defocussing phase (2020), the focussing phase (2030), and a transitional phase (2036). The Sun is in the centre. The ISD density is colour coded: dark blue: no ISD particles reach this region of space; green, yellow, and red colours represent density enhancements with respect to the initial density at 50 AU. The projection of the original flow direction (at 50 AU) is shown in the upper left corner of each plot. Article number, page 15 of 17 −4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2020.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2030.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2036.00, radius 0.072mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density A&A proofs: manuscript no. aa Fig. A.2. Cuts along the ecliptic coordinate planes through the simulated density cubes for particles of size 0.335 µm at different epochs. The years have been picked to select a representative case in the defocussing phase (2020), the focussing phase (2030), and a transitional phase (2036). The Sun is in the centre. The ISD density is colour coded: dark blue: no ISD particles reach this region of space; green, yellow, and red colours represent density enhancements with respect to the initial density at 50 AU. The projection of the original flow direction (at 50 AU) is shown in the upper left corner of each plot. Article number, page 16 of 17 −4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2020.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2030.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2036.00, radius 0.335mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density Peter Strub et al.: Heliospheric modulation of the interstellar dust flow on to Earth Fig. A.3. Cuts along the ecliptic coordinate planes through the simulated density cubes for particles of size 0.492 µm at different epochs. The years have been picked to select a representative case in the defocussing phase (2020), the focussing phase (2030), and a transitional phase (2036). The Sun is in the centre. The ISD density is colour coded: dark blue: no ISD particles reach this region of space; green, yellow, and red colours represent density enhancements with respect to the initial density at 50 AU. The projection of the original flow direction (at 50 AU) is shown in the upper left corner of each plot. Article number, page 17 of 17 −4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2020.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2020.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2030.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2030.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024Y ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Z ecl. [AU] Year 2036.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density−4−2024−4−2024−4−2024X ecl. [AU]−4−2024Y ecl. [AU] Year 2036.00, radius 0.492mu−4−2024−4−20240.000.330.671.001.331.672.002.332.673.003.333.674.00rel. density
1611.02285
2
1611
2016-11-25T17:24:14
Initial mass function of planetesimals formed by the streaming instability
[ "astro-ph.EP" ]
The streaming instability is a mechanism to concentrate solid particles into overdense filaments that undergo gravitational collapse and form planetesimals. However, it remains unclear how the initial mass function of these planetesimals depends on the box dimensions of numerical simulations. To resolve this, we perform simulations of planetesimal formation with the largest box dimensions to date, allowing planetesimals to form simultaneously in multiple filaments that can only emerge within such large simulation boxes. In our simulations, planetesimals with sizes between 80 km and several hundred kilometers form. We find that a power law with a rather shallow exponential cutoff at the high-mass end represents the cumulative birth mass function better than an integrated power law. The steepness of the exponential cutoff is largely independent of box dimensions and resolution, while the exponent of the power law is not constrained at the resolutions we employ. Moreover, we find that the characteristic mass scale of the exponential cutoff correlates with the mass budget in each filament. Together with previous studies of high-resolution simulations with small box domains, our results therefore imply that the cumulative birth mass function of planetesimals is consistent with an exponentially tapered power law with a power-law exponent of approximately -1.6 and a steepness of the exponential cutoff in the range of 0.3-0.4.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability c(cid:13)ESO 2018 September 11, 2018 Initial mass function of planetesimals formed by the streaming instability Urs Schäfer1,2, Chao-Chin Yang2, and Anders Johansen2 1 Hamburg Observatory, University of Hamburg, Gojenbergsweg 112, 21029 Hamburg, Germany, e-mail: [email protected] 2 Lund Observatory, Department of Astronomy and Theoretical Physics, Lund University, Box 43, 22100 Lund, Sweden 6 1 0 2 v o N 5 2 . ] P E h p - o r t s a [ 2 v 5 8 2 2 0 . 1 1 6 1 : v i X r a ABSTRACT The streaming instability is a mechanism to concentrate solid particles into overdense filaments that undergo gravitational collapse and form planetesimals. However, it remains unclear how the initial mass function of these planetesimals depends on the box dimensions of numerical simulations. To resolve this, we perform simulations of planetesimal formation with the largest box dimensions to date, allowing planetesimals to form simultaneously in multiple filaments that can only emerge within such large simulation boxes. In our simulations, planetesimals with sizes between 80 km and several hundred kilometers form. We find that a power law with a rather shallow exponential cutoff at the high-mass end represents the cumulative birth mass function better than an integrated power law. The steepness of the exponential cutoff is largely independent of box dimensions and resolution, while the exponent of the power law is not constrained at the resolutions we employ. Moreover, we find that the characteristic mass scale of the exponential cutoff correlates with the mass budget in each filament. Together with previous studies of high-resolution simulations with small box domains, our results therefore imply that the cumulative birth mass function of planetesimals is consistent with an exponentially tapered power law with a power-law exponent of approximately −1.6 and a steepness of the exponential cutoff in the range of 0.3 -- 0.4. Key words. hydrodynamics -- instabilities -- methods: numerical -- planets and satellites: formation -- protoplanetary disks 1. Introduction One of the greatest problems in the theory of planet formation is to explain how millimeter- or centimeter-sized solid particles -- in the following referred to as pebbles -- grow to kilometer-sized planetesimals. Micron-sized dust grains can grow to pebble sizes by coagulation, but larger particles bounce or fragment under mutual collisions (Güttler et al. 2010; Zsom et al. 2010; Birn- stiel et al. 2011). Growth might continue despite this so-called bouncing barrier for very porous ice particles (Wada et al. 2008, 2009) or by mass transfer in high-speed collisions (Wurm et al. 2005; Windmark et al. 2012). A fundamental problem of planetesimal formation is the time constraint inflicted by the radial drift of solid particles, a prob- lem that persists even under the assumption of perfect sticking. The orbital velocity of the gas in a protoplanetary disk is sub- Keplerian because the gas is supported against the gravity of the central star by a radial pressure gradient. The gas exerts a drag force on the particles in the disk, whose orbital speed would be equal to the Keplerian speed if the gas was not present, causing them to lose angular momentum and drift radially towards the star. The drift velocity depends on the size of the particles, but is in general highest for meter-sized particles in the inner regions of the disk. A particle with a size of 1 m, initially orbiting at a distance of 1 au from the star, drifts towards the star and subli- mates in less than 100 years (Adachi et al. 1976; Weidenschilling 1977; Brauer et al. 2007). Hence, there needs to be some mech- anism assisting the growth of pebbles into planetesimals, which are sufficiently large for the effect of the drag force exerted on them by the gas to be negligible, on a timescale shorter than the radial drift timescale. The streaming instability provides a mechanism to concen- trate solid materials and form planetesimals despite the poor sticking efficiency of the particles and their radial drift. It was discovered analytically by Youdin & Goodman (2005) and con- firmed numerically by Youdin & Johansen (2007), Johansen & Youdin (2007), and Bai & Stone (2010a). The radial drift speed of solid particles decreases with increasing solid-to-gas density ratio because of the drag force exerted by the particles on the gas. A locally enhanced solid-to-gas ratio causes the local or- bital velocity of the gas to be closer to Keplerian, and thus a reduction of the local drift speed of the particles. Hence, clusters of particles drift more slowly than isolated particles, and down- stream clusters can accumulate upstream isolated particles, fur- ther reducing the drift speed of the clusters. Owing to this pos- itive feedback loop, particles can be concentrated into filaments reaching maximum densities of up to several thousand times the local gas density (Bai & Stone 2010a; Johansen et al. 2012; Yang & Johansen 2014; Johansen et al. 2015), sufficient to undergo gravitational collapse and form planetesimals (Johansen et al. 2007; Simon et al. 2016). For this strong clustering of particles to occur, the solid-to-gas column density ratio needs to exceed a critical value (Johansen et al. 2009b; Bai & Stone 2010b), which depends on the radial pressure gradient supporting the gas (Bai & Stone 2010c) and the particle size (Carrera et al. 2015; Yang et al. 2016). Although several studies have shown that the streaming in- stability can lead to the formation of planetesimals, their birth mass distribution has not been comprehensively investigated. However, the initial mass function of planetesimals is essential for the study of the formation of planetary systems because it determines the initial conditions for the evolution of the bodies Article number, page 1 of 11 A&A proofs: manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability that planetesimals evolve into, including planets, asteroids, and Kuiper belt objects. The asteroids in the asteroid belt provide a natural sample distribution that can be fitted with a broken power law. Bottke et al. (2005) argue that the current size distribution of asteroids larger than 120 km in diameter represents the birth size distribution of the planetesimals that formed in the asteroid belt (but have been strongly depleted by resonances with Jupiter, independent of their sizes), while smaller asteroids are largely fragments of collisions between the larger ones. Both Johansen et al. (2015) and Simon et al. (2016) per- formed numerical simulations of planetesimal formation by the streaming instability and find that the differential distribution of the planetesimal birth masses is well-fitted with a power law with an exponent of about −1.6, albeit with the difference that, while the former observe an exponential tapering of the power-law dis- tribution that constitutes the physical upper mass cutoff, the lat- ter do not include such a tapering in their fits. In this paper, we compare power-law fits with and without exponential cutoff to evaluate how well the high-mass end of the initial mass function is described by an exponential cutoff. Johansen et al. (2015) find the shape of the initial mass func- tion to be relatively independent of the resolution of the simu- lations and the solid particle column density. They show that a higher resolution leads to the formation of planetesimals with a wider range of sizes, between 30 km and 120 km in radius in their simulation with the highest resolution because the size of the smallest planetesimal declines with increasing resolution. On the other hand, they observe the size of the largest planetesimal to mainly depend on the particle column density, with smaller column densities yielding smaller sizes. Simon et al. (2016) also studied the dependence of the shape of the birth mass distribution on the resolution of the simulations and obtain the same result. They further find the shape of the distribution to be largely in- dependent of the strength of the self-gravity and the simulation time at which it is initiated, although the masses of the planetes- imals are shifted to higher values with the increasing strength of self-gravity. The planetesimals that formed in their simulations typically range in radius from 50 km to a few hundred kilome- ters. It remains unclear if the planetesimal initial mass function depends on the dimensions of the simulation box. Both Johansen et al. (2015) and Simon et al. (2016) employed only one box size of 0.2 gas scale heights in the radial, azimuthal, and vertical di- rections. However, Yang & Johansen (2014) find that, while in the simulations with this box size, the solid particles are concen- trated by the streaming instability into only one axisymmetric filament, multiple of these filaments form in simulations with larger box dimensions. This raises the question of whether the mass budget of planetesimal formation, and thus the shape of the initial mass function, is different when not only one filament is observed. In this paper, we study simulations with three differ- ent box sizes, the smallest of which is equal to that employed by Johansen et al. (2015) and Simon et al. (2016), while the others are two and four times larger, respectively, in the radial and azimuthal directions, which permits investigating planetesi- mal formation in several filaments. Furthermore, in simulations with larger box sizes, more planetesimals emerge, yielding bet- ter statistics in particular for the determination of the initial mass function. The paper is structured as follows: In Sect. 2, the simulation setup, i.e. the initial conditions and the parameters that govern the evolution of the simulations are described. In Sect. 3, we present our results regarding the formation of planetesimals by the streaming instability and their radial migration. Further, we Article number, page 2 of 11 Table 1. Simulation specifications Name run_0.2_320 run_0.4_320 run_0.8_320 run_0.2_640 run_0.4_640 Lx [Hg]× Ly [Hg]× Lz [Hg]a 0.2× 0.2× 0.2 0.4× 0.4× 0.2 0.8× 0.8× 0.2 0.2× 0.2× 0.2 0.4× 0.4× 0.2 Resolution [H−1 g ] 320 320 320 640 640 Nx × Ny × Nz b 64× 64× 64 128× 128× 64 256× 256× 64 128× 128× 128 256× 256× 128 Notes. (a) Box dimensions in the x-,y-, and z-directions, where Hg is the gas scale height. (b) Number of grid cells in the x-,y-, and z-directions. comment on the issue of permitting the mutual accretion of sink particles, which we use to model planetesimals. In Sect. 4, we discuss whether the planetesimal birth mass distribution is expo- nentially tapered and how its shape depends on the dimensions of the simulation box as well as the resolution. We conclude in Sect. 5. 2. Simulation setup We conduct three-dimensional computer simulations with the Pencil Code1, a hybrid code for gas, for which the magneto- hydrodynamic equations are solved on a fixed grid, with La- grangian particles representing solid bodies. The code employs sixth-order finite differences in space and third-order Runge- Kutta steps in time. We use the shearing box approximation (Goldreich & Lynden-Bell 1965), i.e. we assume that the size of the simula- tion box is small compared to the distance to the central star of the protoplanetary disk. Hence, the curvature of the disk is neglected and the stellar gravity is linearized. The rectangular simulation box is aligned such that the x-, y-, and z-directions correspond to the radial, azimuthal, and vertical directions, re- spectively, and co-rotates with the Keplerian velocity at its ori- gin. For both gas and particles, sheared periodic boundary con- ditions are employed at the radial and azimuthal boundaries and periodic boundary conditions at the vertical boundaries (Hawley et al. 1995; Brandenburg et al. 1995; Youdin & Johansen 2007; Johansen et al. 2009a). In total, we perform five simulations with three different sim- ulation box dimensions and two different resolutions, as listed in Table 1. The two smaller boxes have a size of 0.2 and 0.4 gas scale heights, respectively, in the radial and azimuthal directions with a resolution of either 320 or 640 grid cells per scale height, while the largest box has a radial and azimuthal size of 0.8 scale heights with a resolution of 320 grid cells per scale height. All simulation boxes have a vertical size of 0.2 scale heights. The names of the simulations are composed of the radial and az- imuthal dimension as the first number and the resolution as the second number. 2.1. Gas The simulation box is filled with an isothermal, non-magnetized gas with an isothermal equation of state pg = c2 s ρg, where pg and ρg are the pressure and density, respectively, and cs is the (con- stant) sound speed. While the gas density is initially constant in the radial and azimuthal direction, it is stratified in the vertical direction because we take into consideration the vertical grav- ity of the central star, which causes both gas and solid particles to sediment to the mid-plane at z = 0. This background density 1 http://pencil-code.nordita.org/ Schäfer et al.: Initial mass function of planetesimals formed by the streaming instability (cid:33) (cid:32) − z2 2H2 g stratification is determined by the equilibrium between vertical gravity and vertical pressure gradient, and is given by ρg(z) = ρg,0 exp , (1) where Hg = cs/ΩK is the gas scale height, ΩK = 2π/PK the Ke- plerian orbital frequency, and PK the Keplerian orbital period. Here and in the following, the subscript zero refers to the mid- plane. As formulated in Yang & Johansen (2014), we subtract the background density stratification from the equations of the motion for the gas to numerically balance this equilibrium state down to machine precision. Since the gas density is initially radially constant, there is no radial pressure gradient to support the gas and cause it to orbit with sub-Keplerian speed. Hence, a background pressure gradient set by the dimensionless parameter Π = −1 2 ∂ ln(pg,0) ∂ ln(R) = 0.05, Hg R (2) where R is the orbital distance, is imposed. We refer to this pa- rameter as Π adopting the notation by Bai & Stone (2010b). The resulting sub-Keplerian orbital velocity of the gas is given by vg = vK − ∆v, where vK = ΩKR is the Keplerian orbital velocity and ∆v = Πcs = 0.05cs, which is a representative value at orbital distances of the order of 1 au in a typical protoplanetary disk (Hayashi 1981; Bai & Stone 2010b; Bitsch et al. 2015). 2.2. Particles Two types of Lagrangian particles are employed in the simu- lations: super-particles and sink particles to model pebbles and planetesimals, respectively. To achieve a good load balancing among the processors, we use the particle block domain decom- position algorithm implemented by Johansen et al. (2011). For the calculation of the mutual drag forces between the gas and the super-particles and the mutual gravitational forces between the super-particles and the sink particles, we apply the triangular shaped cloud scheme to map the particle masses and velocities onto the grid, and similarly interpolate the back-reaction drag forces and the self-gravitational forces onto the particles (Hock- ney & Eastwood 1981; Youdin & Johansen 2007; Johansen et al. 2007). The gravitational potential of the super-particles as well as the sink particles is computed by solving Poisson's equation us- ing the fast Fourier transform algorithm (Gammie 2001), which entails gravitational softening. The softening length is of the or- der of the grid cell edge length. Even though the gravitational potential of the particles is vertically periodic, the particles are concentrated in a thin layer around the mid-plane such that their dynamics are not affected by the periodic potential away from the mid-plane. We neglect the contribution of the gas to the gravitational potential under the assumption that the gas component of the protoplanetary disk is not in the gravitationally unstable regime. Indeed, in our simulations the gas density deviates by at most 2% from the mid-plane density ρg,0, the gas density perturbations are thus very small compared to the super-particle densities in the filaments when planetesimals form, which are of the order of the Roche density (see below). The sedimentation of the super-particles to the mid-plane that is due to the vertical stellar gravity induces turbulence as a result of either the streaming instability or the Kelvin-Helmholtz instability, which are both caused by the mutual drag forces be- tween the gas and the super-particles (Bai & Stone 2010b). This turbulence stirs up the super-particles, and hence counteracts the sedimentation. To give sedimentation and turbulence time to at- tain an equilibrium, in our simulations self-gravity is not intro- duced until t = 25 PK. Its strength is then gradually increased from zero over 10 PK until it reaches its final value at t = 35 PK since initiating self-gravity instantaneously with full strength could cause significant impulses on the particles. While we initi- ate self-gravity at a simulation time at which the super-particles have already formed filaments, Johansen et al. (2015) introduced self-gravity with full strength at the start of their simulations and observe qualitatively the same planetesimal birth mass distribu- tion as we do. We achieve this gradual initiation of the self-gravity by sub- (cid:16) 0 stituting (cid:104) π(t−25 PK) (cid:105)(cid:17) Γ = 1− cos t ≤ 25 PK, 25 PK < t < 35 PK, t ≥ 35 PK, where the dimensionless self-gravity parameter 1 2γ γ 10 PK (3) (4) γ = 4πGρg,0 Ω2 K and G is the gravitational constant, into the right-hand side of Poisson's equation. We choose G = P−2 g,0 , and thus γ = 1/π = 0.318. We note that Simon et al. (2016) find the shape of the initial mass function of the planetesimals formed in their simulations of the streaming instability to be relatively indepen- dent of both the simulation time at which self-gravity is intro- duced and the strength of the self-gravity. The Roche density depends on the self-gravity parameter γ and is given by K ρ−1 ρR = 9Ω2 K 4πG = 9ρg,0 γ = 28.3ρg,0. (5) Each super-particle represents a large number of equally sized pebbles because it is computationally infeasible to simu- late the pebbles individually. While the mass of a super-particle is equal to the total mass of the pebbles it models, its friction time is the same as that of an individual constituent pebble. The mass of the super-particles is determined by the initial solid-to-gas column density ratio and their initial number. We set the solid-to-gas ratio Z = Σp,init Σg,init , (6) where Σp,init and √ 2πHgρg,0 (7) Σg,init = are the initial column densities of the super-particles and the gas, respectively, to Z = 0.02. This value corresponds to the critical solid-to-gas ratio necessary for strong clustering of pebbles due to the streaming instability to occur (Johansen et al. 2009b; Bai & Stone 2010b,c; Carrera et al. 2015), and is slightly higher than the solar metallicity. The initial number of super-particles is set equal to the total number of grid cells. The super-particles are randomly distributed among the entire simulation box to seed the streaming instability. The Stokes number of the super-particles τf = ΩKtf, where tf is the friction time, is set to τf = π/10 = 0.314, which, at an Article number, page 3 of 11 A&A proofs: manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability orbital distance of 2.5 au in the Minimum Mass Solar Nebula (MMSN), corresponds to a size of approximately 25 cm (Bai & Stone 2010b; Johansen et al. 2015). While we employ only one fixed particle size, Bai & Stone (2010b) performed simulations with particles of a range of sizes. They find that the particle-gas dynamics are dominated by the most massive particles, and that the critical solid-to-gas ratio required for strong particle cluster- ing owing to the streaming instability is determined by the total mass of all particles. After self-gravity has attained its full strength at t = 35 PK, every super-particle comprised in a cluster whose super-particle density ρp exceeds a threshold value ρp,thres is replaced by a sink particle. This sink particle creation threshold is set to ρp,thres = 200 ρg,0, i.e. about seven times the Roche density (see Eq. 5). We note that Johansen et al. (2015) compared simulations similar to ours with three different threshold values and find the masses of the sink particles emerging in their simulations to be largely independent of the threshold above a value of five times the Roche density. The simulation time at which the formation of sink particles is introduced is arbitrary since both the gravitationally bound super-particle clusters that exist beforehand and the sink parti- cles that emerge from them afterwards represent planetesimals. Owing to the limited resolution of the gravitational forces the behavior of many super-particles inside one grid cell is compa- rable to that of a few sink particles. Nevertheless, the computa- tional expense of the simulations is lowered substantially by the introduction of sink particles. Super-particles within the accretion radius of a sink particle, which is set equal to one grid cell edge length, are accreted by it, i.e. the super-particle mass and momentum are added to the sink particle mass and momentum, respectively, and the super- particle is removed. This accretion, however, might in parts be artificial because the physical accretion radius, i.e. the Bondi ra- dius for pebble accretion (Ormel & Klahr 2010; Lambrechts & Johansen 2012), could be smaller than the simulated accretion radius, especially in the case of less massive sink particles. On the other hand, since the mutual gravitational forces between the super-particles and the sink particles within one grid cell can only be computed inaccurately, the chosen accretion radius cor- responds to the highest accuracy our numerical simulations can offer. We further permit sink particles to accrete one another. This accretion is handled analogously to the super-particle accretion, and might as well be partially artificial. Nevertheless, it is re- quired because in a super-particle cluster exceeding the sink par- ticle creation threshold, all super-particles are replaced by sink particles. That is, the gravitational collapse of a super-particle cluster results in the creation of a cluster of sink particles, of which only one should persist to represent one new-born plan- etesimal. See Sect. 3.2 for further discussion of this topic. 2.3. Units and scaling relations We report our results using the Keplerian orbital period at the origin of the simulation box PK, the gas scale height Hg, and the mid-plane gas density ρg,0 as the units of time, length, and den- sity, respectively. We note that a shearing box freely scales with these units until self-gravity is initiated at t = 25 PK. Afterwards, ρg,0 = γΩ2 K 4πG = πγ GP2 K Article number, page 4 of 11 (8) Eq. (see and g ρg,0 = πγ G−1 H3 [M] = H3 (4)), hence g P−2 K . the unit of mass In the following, relations for the scaling of relevant quanti- ties and units with the orbital distance R, the temperature T , and the mass of the central star MS are given. A mean atomic weight of µ = 2.33 is used and the scaling relations for ρg,0 and [M] are calculated applying those for PK and Hg: (cid:19)−1/2 (cid:19)1/2 yr yr−1 PK = 4.0 ΩK = 1.6 cS = 0.80 2.5 au (cid:18) R (cid:18) R (cid:18) T (cid:18) R 2.5 au 180 K 2.5 au 1 M(cid:12) 1 M(cid:12) (cid:19)3/2 (cid:18) MS (cid:19)−3/2 (cid:18) MS (cid:19)1/2 (cid:19)3/2 (cid:18) T (cid:17)(cid:18) R km s−1 180 K Hg = 0.11 ρg,0 = 9.4× 10−10 (cid:16) γ [M] = 4.2× 1027 (cid:16) γ π−1 π−1 (9) (10) (11) au (12) (cid:19)1/2 (cid:18) MS (cid:19)−3 (cid:18) MS 1 M(cid:12) (cid:19)−1/2 (cid:19) g cm−3 2.5 au 1 M(cid:12) (cid:17)(cid:18) R 2.5 au (cid:19)3/2 (cid:18) T (cid:18) MS (cid:19)3/2 (cid:19)−1/2 180 K × g 1 M(cid:12) (13) (14) Hereafter, we use the above scaling relations, the properties of the asteroid belt, i.e. an orbital distance of R = 2.5 au, a tem- perature of T = 180 K, and a stellar mass of MS = 1 M(cid:12), and the chosen strength of the self-gravity γ = 1/π to convert simu- lation units into physical units. For example, the mid-plane gas density ρg,0 = 9.4× 10−10 g cm−3, which is almost one order of magnitude greater than the corresponding value in the MMSN, ρg,0 = 1.1× 10−10 g cm−3 (Hayashi 1981; Bai & Stone 2010b). The streaming instability has been shown to form planetesimals for pebble column densities similar to that in the MMSN (Jo- hansen et al. 2015). Nevertheless, we choose this comparably high gas density and thus high solid density to promote planetes- imal formation, enabling us to better constrain the initial mass function of planetesimals. 3. Evolution of the simulations Apart from the use of self-gravity and sink particles to model planetesimals, our simulation setup is identical with that applied by Yang & Johansen (2014). Hence, we find our simulations to be consistent with their simulations until t = 25 PK, when self- gravity is introduced (compare the evolution of the pebble den- sity in the simulation time span 20 PK ≤ t ≤ 25 PK shown in Fig. 1 with their Fig. 3). We thus only report on the evolution of our simulations between t = 25 PK and the end of the simula- tions, t = 40 PK. Schäfer et al.: Initial mass function of planetesimals formed by the streaming instability ments, but that they migrate in the radial direction, some of them only marginally, others the entire distance to one of the adjacent filaments. As a result, the sink particles emerge almost evenly distributed among the whole radial dimension of the box. From Fig. 3, it can be seen that the sink particles continue this radial migration, they on average pass through over half of, a few of them even through the whole radial extent of the box. The mean standard deviation of the radial displacement of the sink parti- cles from the locations at which they emerge, averaged over and weighted by the lifetime of every sink particle and then averaged over all sink particles in a simulation, amounts to between 26% and 36% of the radial box size in each of the five simulations. We note that the extent of the migration in the radial direction in- creases with the box size without converging for the box sizes we consider. The radial motions result from the mutual gravitational scattering of sink particles that closely pass by each other. It re- mains to be investigated whether planetesimals are composed of not only pebbles from the filament they form in, but also of an appreciable amount of pebbles from filaments they migrate to. Fig. 1. Pebble column density Σp, integrated over the vertical dimension and averaged over the azimuthal dimension of the simulation box, as a function of radial location x and simulation time t for the simulation with the largest box size, run_0.8_320. Locations at which sink parti- cles emerge are indicated with white dots. Though every pebble cluster forms in one of the filaments, they migrate up to the distance to one of the adjacent filaments. All sink particles emerge as soon as their forma- tion is initiated at t = 35 PK, and they emerge nearly evenly distributed among the entire radial extent of the box owing to the radial migration of the pebble clusters. 3.1. Planetesimal formation and migration If self-gravity is not taken into account the streaming instabil- ity concentrates pebbles into axisymmetric filaments that are elongated in the azimuthal direction (Johansen & Youdin 2007; Johansen et al. 2009b; Bai & Stone 2010b; Yang & Johansen 2014). In Figs. 1 and 2, we show how self-gravity causes these filaments to fragment into pebble clusters that undergo gravi- tational collapse and form planetesimals. In the first 5 PK after self-gravity has been initiated, the filaments begin to disperse be- cause the pebbles accumulate into clusters owing to their mutual gravitational attraction (upper panels of Fig. 2). At t = 35 PK, when the self-gravity reaches its full strength, these clusters con- tain most pebbles available in the simulation, and the filaments are no longer observable (lower left panel). At this point, we commence the formation of sink parti- cles. Almost all pebble clusters exceed the sink particle creation threshold, consequently the pebbles in each of these clusters are replaced by sink particles which merge into one massive sink particle that represents the gravitationally collapsed cluster. At t = 35.1 PK, this sink particle merging process is for the most part completed (lower right panel). However, a few low-mass pebble clusters do not turn into sink particles. All sink parti- cles emerge instantly at t = 35 PK, as can be seen from Fig. 1, but a couple of clusters remain at t = 35.1 PK (three such clus- ters can be spotted in the lower right panel of Fig. 2). Although they are not sufficiently dense to exceed the sink particle cre- ation threshold, these clusters probably represent gravitationally bound planetesimals with low masses. We observe that the planetesimals on average move through more than half of the radial extent of the simulation boxes. Fig- ure 1 shows that each pebble cluster forms in one of the fila- 3.2. Mutual sink particle accretion For the sink particles to realistically represent new-born plan- etesimals, only one of them should be allowed to form from ev- ery pebble cluster that undergoes gravitational collapse. How- ever, this requires a precise determination of the extent of each cluster, which entails several issues, for instance the treatment of overlapping clusters. Therefore, we replace every super-particle that is part of a cluster that exceeds the sink particle creation threshold by a sink particle and allow the sink particles to accrete one another until only one of them remains. We observe that in our five simulations, on average 81% of all accreted sink par- ticles are accreted within 0.1 PK after their formation, i.e. until t = 35.1 PK, because in all simulations all sink particles emerge togather at t = 35 PK. Therefore, the merging process of sink particles that emerged from the same pebble cluster is proba- bly largely completed at this point (see the lower right panel of Fig. 2), and most of the mutual sink particle accretions are a part of this process. On the other hand, we note that the merging of sink particles afterwards may be artificial. The accuracy of the calculation of the gravitational forces between sink particles is limited by the resolution, hence we cannot determine whether two sink parti- cles that encounter one another inside a grid cell collide or pass by each other. Nevertheless, we find the latter to be more prob- able. Taking into account gravitational focusing, the maximum impact parameter leading to a collision of two sink particles (cid:114) bmax = (r1 + r2)2 + 2G(m1 + m2)(r1 + r2) ∆v2 , (15) where mi are the masses of the two sink particles, ri their radii, which are calculated from mi using a solid body density of 3 g cm−3, and ∆v is their relative velocity (see, e.g., Armitage 2007). The mean maximum impact parameter (cid:104)bmax(cid:105), averaged over all mutual sink particle accretions in all five simulations where the lifetime of the accreted sink particle is greater than 0.1 PK and weighted by the lifetimes of the accreted sink parti- cles, amounts to only 3.5% of the grid cell edge length. Mutual accretions of three or more sink particles at the same simulation time are not included in this statistic because we cannot infer their outcome. We note, though, that the sink particle data are not written out after each simulation time step, but every 0.01 PK. Thus, we can only imprecisely determine the simulation time at Article number, page 5 of 11 A&A proofs: manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability Fig. 2. Pebble column density Σp, integrated over the vertical dimension of the simulation box, as a function of radial location x and azimuthal location y at four different simulation times t = 25 PK (top left panel), t = 30 PK (top right panel), t = 35 PK (bottom left panel), and t = 35.1 PK (bottom right panel) for the simulation with the largest box size, run_0.8_320. In the lower right panel, sink particles are plotted as white dots and three pebble clusters are indicated using white circles. After self-gravity has been initiated at t = 25 PK, the pebbles aggregate into clusters and the axisymmetric filaments disperse (upper panels). When the self-gravity attains its full strength at t = 35 PK, most pebbles are concentrated in clusters and the filaments are no longer visible (lower left panel). At this point, we introduce the formation of sink particles, and the pebbles comprised in all but the three encircled clusters are replaced by sink particles. The sink particles emerging from the same pebble cluster undergo a merging process until only one of them remains. This process is largely completed at t = 35.1 PK (lower right panel). which a sink particle merging occurs and the maximum impact parameter for this encounter. The integrated power law can be expressed as (cid:34)(cid:18) Mmax Mpow (cid:19)1−α − (cid:18) M (cid:19)1−α(cid:35) 4. Initial mass function N>(M) Ntot = 1 1− α Mpow , (16) We fit the cumulative mass distribution of the sink particles that emerge in our simulations using two functional forms: an inte- grated power law and a power law with an exponential cutoff. We choose to fit the cumulative mass distributions because, in particular for small numbers of sink particles, they are less af- fected by noise than the differential mass distributions and can thus be fitted more accurately. where N> is the number of sink particles with masses greater than M, Ntot is their total number and Mmax their maximum mass, and Mpow and α are fitting parameters. Since the formation of planetesimals by the streaming instability is a stochastic process and the actual Mmax in a simulation might differ significantly from the Mmax of the ensemble-averaged mass distribution of the sink particles for this model, we treat Mmax as a fitting parameter. Article number, page 6 of 11 t = 25.0 PK-0.4-0.20.00.20.4x [Hg]-0.4-0.20.00.20.4y [Hg]-3.0-2.5-2.0-1.5-1.0-0.50.0log10(Σp/Σg,0)t = 35.0 PK-0.4-0.20.00.20.4x [Hg]-0.4-0.20.00.20.4y [Hg]-3.0-2.5-2.0-1.5-1.0-0.50.0log10(Σp/Σg,0)t = 30.0 PK-0.4-0.20.00.20.4x [Hg]-0.4-0.20.00.20.4y [Hg]-3.0-2.5-2.0-1.5-1.0-0.50.0log10(Σp/Σg,0) Schäfer et al.: Initial mass function of planetesimals formed by the streaming instability Table 2. Best-fitting parameters Name run_0.2_320 run_0.4_320 run_0.8_320 run_0.2_640 run_0.4_640 g P−2 Mmin [G−1 H3 K ]a (1.9± 3.8)× 10−7 (1.9± 0.3)× 10−5 (1.5± 0.2)× 10−5 (4.7± 2.8)× 10−6 (1.2± 0.1)× 10−6 (cid:104)α(cid:105)b 0.0002± 0.0097 0.0057± 0.0094 0.0045± 0.0022 0.0147± 0.0085 0.0097± 0.0011 g P−2 (cid:104)Mexp(cid:105) [G−1 H3 K ]b (1.85± 0.32)× 10−4 (2.31± 0.14)× 10−5 (1.32± 0.02)× 10−4 (5.05± 0.41)× 10−5 (1.82± 0.02)× 10−5 (cid:104)β(cid:105)b 0.324± 0.025 0.284± 0.004 0.375± 0.003 0.294± 0.007 0.352± 0.002 c Nf 1 3 4 1 3 Notes. Listed errors are standard errors. a Best-fitting values at t = 36 PK in the case of run_0.2_320 and at t = 40 PK in the case of the other simulations. (b) Mean values, aver- aged over 36 PK ≤ t ≤ 40 PK. (c) Number of filaments. Table 3. Sink particle statistics Fig. 3. Radial locations x of sink particles which are not accreted before t = 36 PK, color-coded according to the locations at which they emerge, as functions of the simulation time t for the simulation with the largest box size, run_0.8_320. On average, the motions of the sink particles span more than half of the radial extent of the simulation box, a couple of them even migrate through the entire radial dimension of the box. Name run_0.2_320 run_0.4_320 run_0.8_320 run_0.2_640 run_0.4_640 g P−2 K ] g P−2 Ntot Mmin [G−1 H3 5.9× 10−5 4 7.8× 10−6 15 8.3× 10−6 42 1.0× 10−5 7 1.9× 10−6 62 K ] Mmax [G−1 H3 1.6× 10−3 4.3× 10−3 9.3× 10−3 1.5× 10−3 2.2× 10−3 g P−2 K ] (cid:104)M(cid:105) [G−1 H3 4.8× 10−4 4.9× 10−4 6.9× 10−4 2.6× 10−4 1.2× 10−4 The exponentially tapered power law is given by (cid:18) M (cid:19)−α Mpow N>(M) Ntot = (cid:34) exp − (cid:19)β(cid:35) (cid:18) M Mexp , (17) where Mexp and β are fitting parameters. The condition N>(Mmin)/Ntot = 1, where Mmin is the minimum sink particle mass, can be used to eliminate one of the fitting parameters in Eq. (17). We choose to eliminate the characteristic mass of the power-law part, Mpow, which is then given by (cid:34) 1 α (cid:18) Mmin (cid:19)−α (cid:19)β(cid:35) (cid:34)(cid:18) Mmin Mexp . exp Mexp Mpow = Mmin exp (cid:18) M Mmin N>(M) Ntot = Substituting Eq. (18) into Eq. (17) yields (18) (cid:19)β(cid:35) (cid:18) M Mexp . (19) (cid:19)β − We use Eq. (19) to fit the sink particle data because we find the resulting fits to be better than the ones we obtain applying Eq. (17). For the same reason as Mmax in Eq. (16), Mmin is treated as a fitting parameter. To investigate the dependence of the shape of the initial mass function on the resolution and particularly the dimensions of the simulation box, we determine an individual initial mass func- tion for every simulation. At first, we employ the least-squares method to fit Eqs. (16) and (19) to the sink particle data at each Keplerian orbital period between t = 36 PK and the end of the simulations, t = 40 PK. In all five simulations, all sink particles emerge at once at the simulation time at which their formation is initiated, t = 35 PK, but we begin the fitting at t = 36 PK to give the sink particles that emerge from the same pebble cluster time to merge into one. Averaged over this period, we then calculate mean values of the fitting parameters that do not vary signifi- cantly with time and are thus probably relatively unaffected by artificial sink particle merging. In the left panel and the right panel of Fig. 4 we show the sink particle mass distributions as well as the fitted integrated power laws and power laws with exponential tapering for the two sim- ulations with the largest numbers of sink particles, the one with the largest box size, run_0.8_320, and the one with the middle box size and the higher resolution, run_0.4_640, respectively, at t = 40 PK. In the legends for both fits, the standard deviation σ of the actual N>(M)/Ntot for the sink particle masses M from the fitted N>(M)/Ntot is given. We find that the exponential tapered power laws fit the mass distributions better than the integrated power laws. The standard deviations for the power-law fits without exponential cutoff are larger, not only at t = 40 PK as shown in the figure, but also at other simulation times. We further note that the shallower expo- nential cutoff represents the high-mass end of the distributions better than the steeper cutoff of the integrated power law, and that the exponential tapering better reproduces the smooth change of the slope of the distributions. Hence, we limit our further analy- sis to the power-law fits with exponential cutoff. 4.1. Best-fitting parameters 4.2. Power-law fits with and without exponential tapering In Fig. 5 the best-fitting parameters Mmin, α, Mexp, and β in Eq. (19) are shown for t = 36 PK to t = 40 PK for all of our sim- ulations. In the case of the simulation with the smallest box di- mensions and the lower resolution, run_0.2_320, the small num- ber of sink particles persisting after t = 36 PK does not permit to properly fit a power law and an exponential cutoff. The param- eters α, Mexp, and β are approximately constant in time for all simulations apart from Mexp for the simulation with the middle box dimensions and the lower resolution, run_0.4_320, which seems to converge to an upper limit. In Table 2, we list mean values of these three parameters, averaged over the simulation time span from t = 36 PK to t = 40 PK. Similar to Mexp for run_0.4_320, the fitted minimum mass Mmin increases with time for the simulation with the largest box size, run_0.8_320, and the one with the middle box size and the higher resolution, run_0.4_640, but appears to saturate at an up- per limit. The increase of Mmin and Mexp is likely due to both pebble accretion by the sink particles -- as less and less peb- bles remain towards the end of the simulations (see Fig. 1), both masses converge to an upper limit -- and possibly artificial merg- ing of sink particles. In contrast to this, for run_0.4_320 and the simulation with the smallest box dimensions and the higher resolution, run_0.2_640, Mmin is roughly constant. To provide a comparable value for each simulation despite these differences in the time dependence, in Table 2 we list the best-fitting val- Article number, page 7 of 11 x [Hg]353637383940t [PK]-0.4-0.20.00.20.4 A&A proofs: manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability Fig. 4. Cumulative mass distributions of sink particles (black crosses) and fitted integrated power laws (Eq. 16, green lines) and power laws with exponential cutoff (Eq. 19, red lines) at t = 40 PK for the two simulations with the largest numbers of sink particles: the one with the largest box dimensions, run_0.8_320, (left panel) and the one with the middle box dimensions and the higher resolution, run_0.4_640 (right panel). Standard deviations of the actual N>/Ntot (black crosses) from the fitted N>/Ntot are given in the legends. The standard deviations for the integrated power laws are up to twice as large as those for the exponentially tapered power laws. In addition, it can be seen that the shallower exponential cutoffs fit the actual cutoffs more accurately than the steeper cutoffs of the integrated power laws and better replicate the smooth change of the slope of the mass distributions. We thus find the power laws with exponential tapering to represent the mass distributions better than the integrated power laws. ues at t = 40 PK for all simulations (except for run_0.2_320, for which only the value at t = 36 PK is available). A compar- ison of these values shows that Mmin declines with both increas- ing box dimensions and increasing resolution if the value for run_0.2_320 is disregarded. The best-fitting values of the exponent of the power-law component α for all simulations vanish. This is because a lower limit for the sizes of the pebble clusters, and thus the sink par- ticle masses, is set by the resolution, which even in the case of the higher resolution is too large for low-mass sink particles that would constitute the power-law part to emerge. Both Johansen et al. (2015) and Simon et al. (2016) fit the differential mass distribution with a power-law exponent of about −1.6, corre- sponding to α = 0.6. Since the higher resolution we employ, 640 H−1 g , is the lowest one that is considered in these papers, to properly study the power-law distribution higher resolutions than 640 H−1 g seem to be required. We find the mass scale of the exponential cutoff Mexp to correlate with the mass budget in every filament. The parame- ter Mexp should increase with the distance between the filaments because, if the distance is larger, more pebbles can be accreted by the planetesimals forming in each filament. In Col. 6 of Ta- ble 2, the numbers of filaments Nf we observe in our simula- tions are listed. One, three, and four filaments form in the sim- ulation boxes with radial and azimuthal dimensions of 0.2 Hg, 0.4 Hg, and 0.8 Hg, respectively. Therefore, the mass reservoir of pebbles in each filament is similar for the smallest and the largest box sizes, but smaller for the middle box sizes. We in- deed see that the mean values of Mexp for run_0.2_320 and run_0.8_320 are similar, but larger than the one for run_0.4_320. Likewise, the mean value for run_0.2_640 is greater than that for run_0.4_640. Even though the mean values for the two simula- tions with the middle box size differ by more than one standard deviation, we note that the best-fitting values for run_0.4_320 increase with time with a range enclosing the mean value for run_0.4_640. Hence, we find Mexp to be largely independent of the resolution. With the mean values of the exponent of the exponential cutoff β ranging from 0.28 to 0.38, the exponential cutoff is Article number, page 8 of 11 rather smooth. Johansen et al. (2015) fit their data by eye us- ing β = 4/3, which is a significantly steeper cutoff, but this steep cutoff might be an artefact of the small box size they em- ployed, in which only a few massive planetesimals formed. The best-fitting values for the two simulations with the largest num- ber of sink particles, run_0.8_320 and run_0.4_640, are nearly equal, but somewhat greater than the values for run_0.4_320 and run_0.2_640, which are also roughly equal. This indicates that only in the former two simulations enough sink particles emerge to completely capture the high-mass end of the mass distribu- tion. However, the mean values for all simulations lie in a rather small range of 0.1, hence we find β to be relatively independent of the box size and the resolution. Substituting the best-fitting parameters listed in Table 2 into the cumulative (Eq. (19)) or the differential mass distribution, dN dM = 1 M α + β (cid:34) ×exp (cid:18) M (cid:34)(cid:18) Mmin Mexp (cid:19)β(cid:35)(cid:18) M (cid:18) M (cid:19)β − Mpow (cid:19)−α (cid:19)β(cid:35) , Mexp Mexp (20) yields an initial mass function for each simulation. The cumu- lative mass distribution can be converted to a cumulative size distribution, (cid:18) R (cid:19)−3α Rmin N>(R) Ntot = (cid:34)(cid:18) Rmin Rexp (cid:19)3β − (cid:18) R Rexp (cid:19)3β(cid:35) exp , (21) where R is the radius of every sink particle and the minimum radius Rmin and the radius scale of the exponential tapering Rexp can be calculated from Mmin and Mexp, respectively, using a solid density of 3 g cm−3. In Table 3, we list the number Ntot, minimum mass Mmin, maximum mass Mmax, and mean mass (cid:104)M(cid:105) of the sink particles at the end of our five simulations, t = 40 PK. In the case of the two simulations with the smallest box dimensions, we observe less than ten sink particles, and the best-fitting parameters for 10-610-510-410-310-2M [G-1 Hg3PK-2]0.010.101.00N> / Ntot1022102310241025M [g]run_0.8_320integrated power law, σ = 0.089exponentially tapered power law, σ = 0.05910-610-510-410-310-2M [G-1 Hg3PK-2]0.010.101.00N> / Ntot1022102310241025M [g]run_0.4_640integrated power law, σ = 0.071exponentially tapered power law, σ = 0.039 Schäfer et al.: Initial mass function of planetesimals formed by the streaming instability Fig. 5. Best-fitting parameters Mmin (upper left panel), α (upper right panel), Mexp (lower left panel), and β (lower right panel) of the exponentially tapered power law (Eq. (19)) at every Keplerian orbital period between t = 36 PK and t = 40 PK for all five simulations. Standard errors are plotted as error bars. In the case of the simulation with the smallest box size and the lower resolution, run_0.2_320, only parameter values for t = 36 PK are plotted because afterwards only four sink particles persist, which prevents us from properly fitting the mass distributions with a power law and an exponential tapering. these simulations are therefore afflicted with comparably large errors (see Fig. 5). We find the actual Mmin and (cid:104)M(cid:105) to be of the order of the fitted Mmin and Mexp, respectively (compare with Cols. 2 and 4 of Table 2). For run_0.8_320 and run_0.4_640, the 10% of the sink particles which are most massive contain 66% and 70%, respectively, of the total sink particle mass. That is, in our simulations the most massive sink particles dominate the total mass. As stated above, a higher resolution enables us to observe the formation of smaller pebble clusters, and thus less massive sink particles. Hence, Ntot increases and both Mmin and (cid:104)M(cid:105) de- cline with increasing resolution. Like Johansen et al. (2015) and Simon et al. (2016), we find the maximum mass Mmax to be rel- atively independent of the resolution. We expect the number of planetesimals to increase with the number of filaments, and thus with the radial box dimension, and with the length of the filaments, i.e. the azimuthal box di- mension. Analogously to the mass scale of the exponential cut- off Mexp, as discussed above, we further expect (cid:104)M(cid:105) to increase with the distance between the filaments. Our findings are con- sistent with these expectations, with the exception of (cid:104)M(cid:105) for run_0.2_320. One and three filaments form in run_0.2_320 and run_0.4_320, respectively (see Col. 6 of Table 2), therefore the value for the former simulation should be greater than the one for the latter simulation, yet it is slightly smaller. This shows that, at least at the lower resolution, the smallest boxes, in which only one filament forms, might be too small to accurately capture the mass budget of each filament. Furthermore, Mmin and Mmax in general increase with the box dimensions, but we find Mmin to considerably decrease if the box size is increased from 0.2 Hg to 0.4 Hg in the radial and azimuthal directions. This also indicates that the smallest box di- mensions might not capture the sink particle mass distribution as well as the larger boxes. However, it may also be a stochastic ef- fect because, especially for a small number of sink particles, the ensemble-averaged values of Mmin and Mmax might differ signif- icantly from the actual values. 5. Summary and discussion We have investigated the formation of planetesimals by the streaming instability in numerical simulations with three differ- ent box sizes and two different resolutions. In particular, we have studied the initial mass function of these planetesimals, employ- ing the largest box dimensions to date with radial and azimuthal sizes of up to 0.8 gas scale heights. These large box sizes have enabled us to study planetesimal formation in multiple axisym- metric filaments formed by the streaming instability and to yield better statistics because more planetesimals emerge in simula- tions with larger box sizes. In the absence of self-gravity, the streaming instability con- centrates pebbles into axisymmetric filaments. After self-gravity Article number, page 9 of 11 3637383940t [PK]10-810-710-610-510-4Mmin [G-1 Hg3PK-2]1020102110221023Mmin [g]run_0.2_320run_0.4_320run_0.8_320run_0.2_640run_0.4_6403637383940t [PK]10-510-410-3Mexp [G-1 Hg3PK-2]10231024Mexp [g]run_0.2_320run_0.4_320run_0.8_320run_0.2_640run_0.4_6403637383940t [PK]0.000.020.040.060.08αrun_0.2_320run_0.4_320run_0.8_320run_0.2_640run_0.4_6403637383940t [PK]0.150.200.250.300.350.40βrun_0.2_320run_0.4_320run_0.8_320run_0.2_640run_0.4_640 A&A proofs: manuscript no. Initial_mass_function_of_planetesimals_formed_by_the_streaming_instability has been introduced, these filaments disperse within about ten Keplerian orbital periods because the pebbles accumulate into clusters that undergo gravitational collapse and form planetes- imals. We have observed that, after their formation, the plan- etesimals on average migrate through more than half of the ra- dial dimension of the simulation box owing to mutual gravita- tional scattering. The extent of the radial migration does not con- verge for the box sizes we have taken into consideration. Further studies could provide insights regarding the implications of the migration through multiple filaments for the dependence of the composition of planetesimals on the orbital distance. The radii of the planetesimals formed in our simulations, which depend on the strength of the self-gravity and thus on the solid particle column density, range from 80 km to 620 km. We have compared power-law fits to their cumulative mass dis- tribution with and without exponential tapering and have found that a rather shallow exponential cutoff fits the distribution bet- ter than the steeper cutoff of an integrated power law. Johansen et al. (2015) also find the initial mass function to be represented best by an exponentially tapered power law, although they stud- ied planetesimals that are smaller than the ones formed in our simulations. In their simulation with the highest resolution, the planetesimal radii amount to between 30 km and 120 km. In con- trast to this, Simon et al. (2016) find that a power law without exponential tapering is suitable to fit the birth mass distribution of planetesimals with radii between 50 km and a few hundred kilometers. We have found a value of the exponent of the exponen- tial cutoff of about 0.3 to 0.4, which is largely invariant un- der changes in the box size and the resolution, but considerably smaller than the value of 4/3 that Johansen et al. (2015) deter- mine, which is based on a much smaller number of massive plan- etesimals that formed in a small simulation domain. However, the resolutions we have considered are insufficient to constrain the shape of the power-law part and to investigate its dependence on the box dimensions because the planetesimals that formed in our simulations are too large to constitute a power-law distribu- tion at the low-mass end of the initial mass function. Both the characteristic mass of the exponential cutoff and the mean mass of the planetesimals correlate with the pebble mass budget in every filament. In this regard, we have found indica- tions that a simulation box with a size of 0.2 gas scale heights in the radial, azimuthal, and vertical directions, in which only one filament emerges, may be too small to properly capture the mass reservoir. This is consistent with the observation by Yang & Jo- hansen (2014) that box dimensions of 0.2 scale heights in the radial and azimuthal directions are too small to capture all scales relevant for the streaming instability, and is further supported by Li et al.'s (in prep.) finding that the density distribution function of solid particles is consistent only for boxes with radial and az- imuthal sizes of at least 0.4 scale heights. The current size distribution of the asteroids in the aster- oid belt with diameters between 120 km and several hundred kilometers, corresponding to the sizes of the planetesimals that emerge in our simulations, is well-fitted with a power law. This power law, which Bottke et al. (2005) argue represents the pri- mordial asteroid size distribution, is in contrast to the exponen- tial cutoff we (and Johansen et al. 2015) find. Subsequent pebble accretion therefore appears to be necessary to convert the expo- nential tapering of the birth size distribution into the power law observed in the asteroid belt (Johansen et al. 2015). It is interesting to compare the initial mass function of plan- etesimals to the classical concept of an initial mass function of stars. The formation of stars is comparable to the formation of Article number, page 10 of 11 planetesimals by the streaming instability insofar as both stars and planetesimals form by gravitational collapse, the former from molecular cloud cores, and the latter from pebble clusters. The differential mass distribution of stars with masses greater than about 1 M(cid:12) is given by a power law with an exponent of ap- proximately −2.3 (Salpeter 1955; Massey 1998; Chabrier 2003). That is, the total stellar mass is dominated by the least massive stars, in contrast to the total mass of the planetesimals we have observed in our simulations, which is dominated by the most massive ones. It has been argued that there is a physical upper mass cutoff of the stellar initial mass function, but massive stars are rare and short-lived, and their mass distribution is therefore difficult to observe (Zinnecker & Yorke 2007). Johansen et al. (2015) and Simon et al. (2016) investi- gated the dependence of the shape of the planetesimal initial mass function on the resolution, the pebble column density, the strength of the self-gravity, and the simulation at which self-gravity is initiated. We have complemented these parameter studies with an analysis of the box-size dependence, but how, for instance, the solid-to-gas ratio, the friction time of the pebbles, the radial gas pressure gradient, and the vertical box size affect the shape of the birth mass distribution remains to be investi- gated. Finally, the streaming instability has been shown to oper- ate in protoplanetary disks with turbulence driven by the mag- netorotational instability (Johansen et al. 2007). It remains to be seen how turbulence and magnetic fields influence the shape of the initial mass function. Acknowledgements. We thank Piero Ranalli for his advice on how to calculate the standard errors of the fitting parameters. We further thank the anonymous referee for providing comments and questions that helped to improve the paper. The simulations presented in this paper were performed on resources provided by the Swedish National Infrastructure for Computing (SNIC), the Alarik system at Lunarc at Lund University and the Beskow system at the PDC Center for High Performance Computing at the KTH Royal Institute of Technology. This research was supported by the European Research Council under ERC Starting Grant agreement 278675-PEBBLE2PLANET. A. J. is thankful for financial support from the Knut and Alice Wallenberg Foundation and from the Swedish Research Council (grant 2014-5775). References Adachi, I., Hayashi, C., & Nakazawa, K. 1976, Progress of Theoretical Physics, Armitage, P. J. 2007, ArXiv Astrophysics e-prints [astro-ph/0701485] Bai, X.-N. & Stone, J. M. 2010a, ApJS, 190, 297 Bai, X.-N. & Stone, J. M. 2010b, ApJ, 722, 1437 Bai, X.-N. & Stone, J. M. 2010c, ApJL, 722, L220 Birnstiel, T., Ormel, C. W., & Dullemond, C. P. 2011, A&A, 525, A11 Bitsch, B., Johansen, A., Lambrechts, M., & Morbidelli, A. 2015, A&A, 575, 56, 1756 A28 741 Bottke, W. F., Durda, D. D., Nesvorný, D., et al. 2005, Icarus, 175, 111 Brandenburg, A., Nordlund, Å., Stein, R. F., & Torkelsson, U. 1995, ApJ, 446, Brauer, F., Dullemond, C. P., Johansen, A., et al. 2007, A&A, 469, 1169 Carrera, D., Johansen, A., & Davies, M. B. 2015, A&A, 579, A43 Chabrier, G. 2003, PASP, 115, 763 Gammie, C. F. 2001, ApJ, 553, 174 Goldreich, P. & Lynden-Bell, D. 1965, MNRAS, 130, 125 Güttler, C., Blum, J., Zsom, A., Ormel, C. W., & Dullemond, C. P. 2010, A&A, 513, A56 Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1995, ApJ, 440, 742 Hayashi, C. 1981, Progress of Theoretical Physics Supplement, 70, 35 Hockney, R. W. & Eastwood, J. W. 1981, Computer Simulation Using Particles (New York: McGraw-Hill) Johansen, A., Klahr, H., & Henning, T. 2011, A&A, 529, A62 Johansen, A., Mac Low, M.-M., Lacerda, P., & Bizzarro, M. 2015, Science Ad- vances, 1 [arXiv:1503.07347] Johansen, A., Oishi, J. S., Mac Low, M.-M., et al. 2007, Nature, 448, 1022 Johansen, A. & Youdin, A. 2007, ApJ, 662, 627 Johansen, A., Youdin, A., & Klahr, H. 2009a, ApJ, 697, 1269 Schäfer et al.: Initial mass function of planetesimals formed by the streaming instability Johansen, A., Youdin, A., & Mac Low, M.-M. 2009b, ApJL, 704, L75 Johansen, A., Youdin, A. N., & Lithwick, Y. 2012, A&A, 537, A125 Lambrechts, M. & Johansen, A. 2012, A&A, 544, A32 Massey, P. 1998, in Astronomical Society of the Pacific Conference Series, Vol. 142, The Stellar Initial Mass Function (38th Herstmonceux Conference), ed. G. Gilmore & D. Howell, 17 Ormel, C. W. & Klahr, H. H. 2010, A&A, 520, A43 Salpeter, E. E. 1955, ApJ, 121, 161 Simon, J. B., Armitage, P. J., Li, R., & Youdin, A. N. 2016, ApJ, 822, 55 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2008, ApJ, 677, Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2009, ApJ, 702, 1296 1490 Weidenschilling, S. J. 1977, MNRAS, 180, 57 Windmark, F., Birnstiel, T., Güttler, C., et al. 2012, A&A, 540, A73 Wurm, G., Paraskov, G., & Krauss, O. 2005, Icarus, 178, 253 Yang, C.-C. & Johansen, A. 2014, ApJ, 792, 86 Yang, C.-C., Johansen, A., & Carrera, D. 2016, A&A, [arXiv:1611.07014] submitted Youdin, A. & Johansen, A. 2007, ApJ, 662, 613 Youdin, A. N. & Goodman, J. 2005, ApJ, 620, 459 Zinnecker, H. & Yorke, H. W. 2007, ARA&A, 45, 481 Zsom, A., Ormel, C. W., Güttler, C., Blum, J., & Dullemond, C. P. 2010, A&A, 513, A57 Article number, page 11 of 11
1803.00215
2
1803
2019-03-03T08:05:06
The continued importance of habitability studies
[ "astro-ph.EP" ]
This is a white paper in response to the National Academy of Sciences "Exoplanet Science Strategy" call. We summarize recent advances in theoretical habitability studies and argue that such studies will remain important for guiding and interpreting observations. Interactions between 1-D and 3-D climate modelers will be necessary to resolve recent discrepancies in model results and improve habitability studies. Observational capabilities will also need improvement. Although basic observations can be performed with present capabilities, technological advances will be necessary to improve climate models to the level needed for planetary habitability studies.
astro-ph.EP
astro-ph
THE CONTINUED IMPORTANCE OF HABITABILITY STUDIES A white paper submitted in response to the NAS call on Exoplanet Science Strategy Lead Author: Ramses M. Ramirez Earth-Life Science Institute, Tokyo Institute of Technology, Tokyo 152-8550 email: [email protected] phone: 03-5734-2183 Co-authors: Dorian S. Abbot (University of Chicago) Vladimir Airapetian (NASA Goddard Space Flight Center and American University) Yuka Fujii (Earth-Life Science Institute and Tokyo Institute of Technology) Keiko Hamano (Earth-Life Science Institute and Tokyo Institute of Technology) Amit Levi (Harvard University) Tyler D. Robinson (Northern Arizona University) Laura Schaefer (Arizona State University) Eric T. Wolf (University of Colorado at Boulder) Robin D. Wordsworth (Harvard University) INTRODUCTION The habitable zone (HZ) is the hypothesized circumstellar region where standing bodies of liquid water could be stable on a planetary surface. The classical HZ [1] suggests that CO2 and H2O are the key greenhouse gases for habitable exoplanets as they are on Earth. This implies that the climates of potentially habitable exoplanets are regulated via an Earth-like carbonate- silicate cycle (or equivalent) that maintains planetary habitability over Gyr timescales [1]. Such assumptions do not suggest that lifeforms elsewhere must be Earth-like or that they could even evolve under such conditions, however. For example, can life evolve on a planet with a 10-bar CO2 atmosphere that is located near its host star's outer edge? Likewise, could the different radiation environment on a M-star planet produce life? Would any such life be "Earth-like"? Is the classical CO2-H2O HZ really the best and only targeting tool that should be used to find potentially habitable planets? Theoretical habitability studies will remain important for addressing such questions. Moreover, a proper theoretical understanding of habitability is necessary to guide and interpret future observations. In Section 1, we summarize recent habitability research, which includes updates to our understanding of habitable zones, ocean worlds, and magma oceans. Some potential research directions are given. We then argue that improving theoretical habitability studies will require 1-D and 3-D climate modelers to discuss and attempt to resolve recent differences in model results (Section 2). Technological improvements in imaging are also needed to further better such studies. Nevertheless, some key observations can be made now and in the near-future (Section 3). We conclude by summarizing the aforementioned points. 1a) Habitable zones: extensions in space SECTION 1: HABITABILITY The outer edge of the HZ is extraordinarily sensitive to the allowable greenhouse gas combinations. For instance, the solar system's outer edge extends from ~1.7 to 10 AU if young planets can accrete hundreds to thousands of bars of hydrogen, H2, from the protoplanetary disk[2]. Habitable conditions (e.g. surface conditions warm enough to support liquid water) for such planets located within the classical HZ may last a few to tens of millions of years [3], possibly longer at farther distances [2] or if biological feedbacks can regulate H2 [4]. Moreover, H2 can be volcanically-outgassed within a background CO2 atmosphere, increasing the classical HZ width by ~ 50% or more [5]. If mantle conditions can remain reducing enough and/or H2 escape rates to space are sufficiently slowed, habitable conditions within the classical HZ may last on Gyr timescales [5][6]. In the latter scenario, such secondary greenhouse gases can confound carbonate-silicate cycle predictions suggesting that CO2 pressures for habitable planets should increase towards the outer edge [7]. This is because additional absorbers (like H2) lower the CO2 amounts needed to support habitable surface conditions[5]. Related concerns: a. Would it be possible for life to subsist on a H2-based form of photosynthesis [8]? b. The effects of secondary greenhouse gases on HZ boundaries are just beginning to be explored. Other greenhouse gas combinations should also be assessed. 1b) Habitable zones: changes with time Although most HZ studies have focused on the main-sequence phase of stellar evolution, the pre-main-sequence is crucial in discussions of planetary habitability. The habitability of M- star planets should be questioned for at least two reasons: a. Such HZ planets (e.g. TRAPPIST-1) should have lost much - if not all - of their surface water inventory following a runaway greenhouse episode during the superluminous pre-main-sequence phase [9] -- [11]unless they could have accreted a few tens % of their mass in water [12], [13]. b. Large fluxes of stellar radiation (X-ray and UV) and stellar winds could significantly erode exoplanetary atmospheres over geological timescales [14]. Thus, M-star planets located in the HZ today may face extra challenges to their habitability. The entire spectrum of A -- M stars should be observed with these caveats in mind. Related concerns: a. Can desiccated M-dwarf HZ planets be partially replenished by later meteoritic impacts or mantle degassing and become potentially habitable [9]? 1c) Magma oceans as keys to understanding atmospheric evolution and habitability In addition, the pre-main-sequence phase is crucial for understanding how worlds retain and lose volatiles during this stage of early accretion [15]. Moreover, high levels of abiotically-produced oxygen, O2, incorrectly linked to life, may accumulate in magma ocean atmospheres [10]. However, the resulting magma ocean and impact- induced soil oxidation processes may be important O2 sinks [16][17]. Thus, the magma ocean stage may explain differences in evolutionary histories between Earth and Venus and could elucidate the conditions necessary for the emergence of life on exoplanets. Related concerns: a. Most magma ocean studies to date have considered mostly H2O- or CO2- dominated atmospheres [15][16][18][19], but many atmospheric compositions are possible depending on impactor chemistry [20]. These should all be explored. b. This early stage of planetary evolution could be the key to providing a more definitive HZ inner edge, improving our ability to target worlds that are likely to have sizeable water inventories [15]. 1d) The possible habitability of ocean worlds It has been argued that the carbonate-silicate cycle requires some land to operate, which suggests that ocean worlds (which have no land and possess total water inventories that are many times that of Earth's) may not be habitable [21]. However, sea ice enriched in clathrates should readily form within the resulting high CO2 atmospheric environments predicted for such planets, providing conditions favorable for the initiation of life [22]. Plus, the carbonate-silicate cycle may be replaced by one in which net outgassing of CO2 is balanced by a net ocean influx [13]. High CO2 levels may also cool the stratosphere sufficiently to prevent a moist greenhouse [22]. SECTION 2: CLIMATE COMMUNITY COLLABORATIONS 2a) 1-D and 3-D climate models as planetary habitability assessment tools Until recently, virtually all exoplanet habitability studies had been performed using 1-D climate models. However, the types of climate models used to perform habitability assessments have significantly diversified over the past ~5 years. Exoplanet habitability studies are now routinely performed using a wide array of 1-D (including energy balance models) (e.g. [5][9] [15][16][18][19][22] -- [24]) and 3-D (e.g. [25][26][27][28][29][30][31][32][33]) climate models, with some studies employing both [34][35] or even hybrid approaches [36]. This trend has occurred because both 1-D and 3-D climate models have complementary advantages and disadvantages. 1-D models are computationally cheap, allowing quick exploration of parameter space, more complex radiative transfer and atmospheric chemistry, and are more easily coupled to other (e.g. interior, escape, stellar, photochemical) models. For instance, atmospheric and magma ocean processes can be readily coupled in 1-D models [16], allowing volatile retention and O2 build-up to be quantified. Also, coupled climate- photochemical calculations are routinely performed [24]. In contrast, 3-D models calculate more atmospheric processes self-consistently than can be done with 1-D models, which have to make more simplifying assumptions regarding dynamics, relative humidity, and clouds. Thus, 3-D models are used to assess the subtleties of atmospheres in greater detail, which includes evaluating complex circulation patterns. 2b) The need for collaborations between 1-D and 3-D modelers Unfortunately, there is currently a lack of information exchange between the 1-D and 3-D climate modeling communities. However, community interactions are necessary for addressing recent unresolved differences in model predictions such as: a. Does relative humidity increase with surface temperature as 1-D models predict [37], [38]or can it decrease as suggested by some 3-D models [39]? b. Is the moist greenhouse for M-star inner edge planets triggered at the low mean surface temperatures predicted in some 3-D models [28][31] or at the higher temperatures calculated in other 3-D and 1-D models [1][40]? c. Do surface temperature inversions in very warm atmospheres truly occur as predicted in recent 3-D simulations [41]? d. How valid is the 1-D moist adiabatic lapse rate assumption for planetary atmospheres [1][41]? SECTION 3: CURRENT LIMITATIONS AND SUGGESTED OBSERVATIONS All models have to make simplifications, however. With 3-D models, this usually means tuning poorly-understood processes or parameterizing them based on how they operate on Earth. With 1-D models, this requires making assumptions about the relative humidity and lapse rate. Unfortunately, such simplified approaches are almost certainly inadequate to study exoplanets. Plus, current observations are also woefully limited. For example, present observational capabilities are insufficient to characterize terrestrial HZ planets and their atmospheres in great detail, requiring modelers to "invent" atmospheric compositions and guess what the values for specific planetary parameters may be. Thus, major technological advances in observational techniques are necessary to produce data that is of sufficient quality to inform and substantially improve exoplanet climate models. In spite of these limitations, some examples of observations that can still be made today and in the near-future include: a. Do M-star HZ planets have atmospheres? If they do, does O2 build up? Does O2 build up on HZ planets orbiting other star types? HabEx and LUVOIR could observe atmospheres for signs of O2 buildup at VIS/NIR wavelengths. ELTs can also detect O2 through multiple transits [42]. b. OST and JWST [43]could detect O3 in O2-rich atmospheres. c. The ELTs may be able to detect thermal emission from magma oceans [44][45]. d. The ELTs may observe nearby M-stars for planets located in the pre-main- sequence HZ since they can work at small inner working angles [9]. e. High predicted CO2 pressures for outer edge classical HZ planets lower scale heights and increase the difficulty to observe bioindicators in transmission spectroscopy. However, H2 can increase atmospheric scale height in sufficiently high quantities, which improves bioindicator detection [5][46]. JWST could target the TRAPPIST-1 planets and LHS-1140b [47]. f. Ocean worlds with sufficient water may be deduced from other types of habitable planets. Lower computed densities for ocean worlds would distinguish them from drier ones, breaking the mass-radius degeneracy. Moreover, should such atmospheres be CO2-rich [13], lower scale heights would distinguish them from worlds with H/He outer envelopes. Once enough planetary atmospheres inside and outside the HZ are characterized, statistics will be available to test and improve model predictions [1], [7]. In turn, improved climate models would be used to provide better interpretations of observations. FINAL SUMMARY As the direct links between theory (Section 1) and observations (Section 3) suggest, theoretical habitability studies will remain indispensable for observations. This is because such studies, which include understanding the planetary (e.g. atmospheric, geologic) and stellar processes1 that make a planet more or less likely to support life, provide the roadmap for making proper observations. Continued work on habitability is essential even if life elsewhere happens to be exactly like it is on Earth- which would still have appeared very differently to an extraterrestrial observer examining our world at different points in geologic history [48]. 1 Although we have focused on climate modeling in this white paper, we also support more theoretical habitability studies in other areas (e.g. interiors, photochemistry, volatile delivery, stellar atmospheres, aeronomy, and atmospheric escape). The likelihood of finding life can be increased by suspending some of our most cherished Earth-centric notions of habitability, whatever they may be (e.g. carbonate-silicate cycles, CO2-H2O atmospheres, oxygen biochemistries), no matter how compelling they may seem, and considering assessing alternate scenarios as well. We do not know what extraterrestrial life may be like, so it is self-restricting to generalize life elsewhere solely based on a limited understanding about this planet. Healthy scientific speculation will be key to making progress given limited observations. Both 1-D and 3-D climate models will be essential for advancing theoretical habitability studies. Community partnerships aimed at resolving model differences will lead to improved exoplanet climate models, maximizing the utility of such theoretical habitability studies. Such efforts would ultimately lead to improved observational interpretations as well. lack of detailed exoplanetary observations. Although drawing upon knowledge from various solar system bodies to infer exoplanetary processes has been a useful tactic [1][5][6][15][49] having access to better observations would lead to even bigger advances in our understanding. This situation would vastly improve with technological (e.g. engineering) advancements in observational techniques, including direct imaging. Even with current limitations, we have suggested some observations that can still be made with upcoming and next generation missions. REFERENCES [1] However, improvements to theoretical modeling efforts are currently stymied by the J.F. Kasting, D. Whitmire, and R. Raynolds, "Habitable Zones Around Main Sequence Stars," Icarus, vol. 101, pp. 108 -- 128, 1993. [2] R. Pierrehumbert and E. Gaidos, "Hydrogen Greenhouse Planets Beyond the Habitable Zone," Astrophys. J. Lett., vol. 734, no. L13, Apr. 2011. [3] R.D. Wordsworth, "Transient conditions for biogenesis on low-mass exoplanets with escaping hydrogen atmospheres," Icarus, vol. 219, no. 1, pp. 267 -- 273, 2012. [4] D. S. Abbot, "A Proposal for Climate Stability on H2-greenhouse Planets," Astrophys. J. Lett., vol. 815, no. 1, 2015. [5] R. M. Ramirez and L. Kaltenegger, "A Volcanic Hydrogen Habitable Zone," Astrophys. J. Lett., vol. 837, no. 1, 2017. [6] R. M. Ramirez, R. Kopparapu, M. E. Zugger, T. D. Robinson, R. Freedman, and J. F. [7] Kasting, "Warming early Mars with CO2 and H2," Nat. Geosci., vol. 7, no. 1, pp. 59 -- 63, Nov. 2014. E. M.-R. Bean, Jacob L.; Abbot, Dorian S.; Kempton, "A Statistical Comparative Planetology Approach to the Hunt for Habitable Exoplanets and Life Beyond the Solar System," Astrophys. J. Lett., vol. 841, no. 2, 2017. [8] W. Bains, S. Seager, and A. Zsom, "Photosynthesis in Hydrogen-Dominated Atmospheres," pp. 716 -- 744, 2014. [9] R. M. Ramirez and L. Kaltenegger, "The habitable zones of pre-main-sequence stars," Astrophys. J. Lett., vol. 797, no. 2, 2014. [10] R. Luger and R. Barnes, "Extreme water loss and abiotic O2 buildup on planets throughout the habitable zones of M dwarfs," Astrobiology, vol. 15, no. 2, pp. 119 -- 143, 2015. [11] F. Tian and S. Ida, "Water contents of Earth-mass planets around M dwarfs," Nat. Geosci., vol. 8, no. 3, 2015. [12] Y. Alibert and W. Benz, "Formation and composition of planets around very low mass stars," Astron. Astrophys., vol. 598, no. L5, 2017. [13] M. Levi, A.; Sasselov, D.; Podolak, "The Abundance of Atmospheric CO2 in Ocean Exoplanets: a Novel CO2 Deposition Mechanism," Astrophys. J., vol. 838, no. 1, 2017. [14] M. W. Airapetian, Vladimir S.; Glocer, Alex; Khazanov, George V.; Loyd, R. O. P.; France, Kevin; Sojka, Jan; Danchi, William C.; Liemohn, Michael W.Airapetian, Vladimir S.; Glocer, Alex; Khazanov, George V.; Loyd, R. O. P.; France, Kevin; Sojka, Jan; Danchi, Wi, "How Hospitable Are Space Weather Affected Habitable Zones? The Role of Ion Escape," Astrophys. J. Lett., vol. 836, no. 1, 2017. [15] K. Hamano, Y. Abe, and H. Genda, "Emergence of two types of terrestrial planet on solidification of magma ocean," Nature, vol. 497, no. 7451, pp. 607 -- 610, 2013. [16] D. Schaefer, L., Wordsworth, R. D., Berta-Thompson, Z., Sasselov, "Predictions of the atmospheric composition of GJ 1132b," Astrophys. J., vol. 2, no. 63, 2016. [17] K. Kurosawa, "Impact-driven planetary desiccation: The origin of the dry Venus," Earth Planet. Sci. Lett., vol. 429, no. 1, pp. 181 -- 190, 2015. [18] A. Marcq, E.; Salvador, A.; Massol, H.; Davaille, "Thermal radiation of magma ocean planets using a 1-D radiative-convective model of H2O-CO2 atmospheres," J. Geophys. Res. Planets, vol. 122, no. 7, pp. 1539 -- 1553, 2017. [19] E. Salvador, A.; Massol, H.; Davaille, A.; Marcq, E.; Sarda, P.; Chassefière, "The relative influence of H2O and CO2 on the primitive surface conditions and evolution of rocky planets," J. Geophys. Res. Planets, vol. 122, no. 7, 2017. [20] L. Schaefer and B. Fegley Jr., "Chemistry of atmospheres formed during accretion of the Earth and other terrestrial planets," Icarus, vol. 208, no. 1, pp. 434 -- 448, 2010. [21] F. J. Abbot, Dorian S.; Cowan, Nicolas B.; Ciesla, "Indication of Insensitivity of Planetary Weathering Behavior and Habitable Zone to Surface Land Fraction," Astrophys. J., vol. 756, no. 2, 2012. [22] R. D. Wordsworth and R. T. Pierrehumbert, "Water Loss From Terrestrial Planets With Co 2 -Rich Atmospheres," Astrophys. J., vol. 778, no. 2, p. 154, Dec. 2013. [23] A. Vladilo, Giovanni; Silva, Laura; Murante, Giuseppe; Filippi, Luca; Provenzale, "Modeling the Surface Temperature of Earth-like Planets," Astrophys. J., vol. 804, no. 1, 2015. [24] K. Engin, J. L. Grenfell, M. Godolt, S. Barbara, and R. Heike, "The Effect of Varying Atmospheric Pressure upon Habitability and Biosignatures of Earth-like Planets," Astrobiology, vol. 18, no. 2, 2018. [25] J. Yang, N. Cowan, and D. Abbot, "Stabilizing cloud feedback dramatically expands the habitable zone of tidally locked planets," Astrophys. J. Lett., vol. 771, no. 2, 2013. [26] A. P. Kaspi, Yohai; Showman, "Atmospheric Dynamics of Terrestrial Exoplanets over a Wide Range of Orbital and Atmospheric Parameters," Astrophys. J., vol. 804, no. 1, 2015. [27] S. Kopparapu, Ravi kumar; Wolf, Eric T.; Haqq-Misra, Jacob; Yang, Jun; Kasting, James F.; Meadows, Victoria; Terrien, Ryan; Mahadevan, "The Inner Edge of the Habitable Zone for Synchronously Rotating Planets around Low-mass Stars Using General Circulation Models," Astrophys. J., vol. 819, no. 1, 2016. [28] D. S. Fujii, Yuka; Del Genio, Anthony D.; Amundsen, "NIR-driven Moist Upper Atmospheres of Synchronously Rotating Temperate Terrestrial Exoplanets," Astrophys. J., vol. 848, no. 2, 2017. [29] G. Turbet, Martin; Forget, Francois; Leconte, Jeremy; Charnay, Benjamin; Tobie, "CO2 condensation is a serious limit to the deglaciation of Earth-like planets," Earth Planet. Sci. Lett., vol. 476, pp. 11 -- 21, 2017. [30] E. T. Wolf, "Assessing the Habitability of the TRAPPIST-1 System Using a 3D Climate Model," Astrophys. J. Lett., vol. 839, no. 1, 2017. [31] R.K. Kopparapu; Wolf, Eric T.; Arney, Giada; Batalha, Natasha E.; Haqq-Misra, Jacob ; Grimm, Simon L.; Heng, "Habitable moist atmospheres on terrestrial planets near the inner edge of the habitable zone around M dwarfs," Astrophys. J., vol. 1, no. 5, 2017. [32] J. D. Haqq-Misra, E. T. Wolf, M. Josh, X. Zhang, and R. K. Kopparapu, "Demarcating Circulation Regimes of Synchronously Rotating Terrestrial Planets within the Habitable Zone," Astrophys. J., vol. 852, no. 2, 2018. [33] Y. Kodama, T.; Nitta, A.; Genda, H.; Takao, Y.; O'ishi, R.; Abe-Ouchi, A.; Abe, "Dependence of the Onset of the Runaway Greenhouse Effect on the Latitudinal Surface Water Distribution of Earth-Like Planets," J. Geophys. Res. Planets, vol. 123, no. 2, pp. 559 -- 574, 2018. [34] D. S. Mills, Sean M.; Abbot, "Utility of the Weak Temperature Gradient Approximation for Earth-like Tidally Locked Exoplanets," Astrophys. J. Lett., vol. 774, no. 2, 2013. [35] J. Checlair, K. Menou, and D. S. Abbot, "No snowball on habitable tidally locked planets," Astrophys. J., vol. 845, no. 2, 2017. [36] L. Carone, R. Keppens, L. Decin, and T. Henning, "Stratosphere circulation on tidally locked ExoEarths," Mon. Not. R. Astron. Soc., vol. 473, no. 4, pp. 4672 -- 4685, 2017. [37] C. Goldblatt, T. D. Robinson, K. J. Zahnle, and D. Crisp, "Low simulated radiation limit for runaway greenhouse climates," Nat. Geosci., vol. 6, no. 8, pp. 661 -- 667, Jul. 2013. [38] R. M. Ramirez, R. K. Kopparapu, V. Lindner, and J. F. Kasting, "Can increased atmospheric CO2 levels trigger a runaway greenhouse?," Astrobiology, vol. 14, no. 8, pp. 714 -- 31, Aug. 2014. [39] G. L. Russell, A. a. Lacis, D. H. Rind, C. Colose, and R. F. Opstbaum, "Fast atmosphere- ocean model runs with large changes in CO 2," Geophys. Res. Lett., vol. 40, no. 21, pp. 5787 -- 5792, Nov. 2013. [40] J. Leconte, F. Forget, B. Charnay, R. Wordsworth, and A. Pottier, "Increased insolation threshold for runaway greenhouse processes on Earth-like planets.," Nature, vol. 504, no. 7479, pp. 268 -- 71, Dec. 2013. [41] E. T. Wolf and O. B. Toon, "Journal of Geophysical Research : Atmospheres the brightening Sun," J. Geophys. Res. Atmos., vol. 120, no. 12, pp. 5775 -- 5794, 2015. [42] F. Rodler and M. Lopez-Morales, "Feasibility studies for the detection of O2 in an Earth- like exoplanet," Astrophys. J., vol. 781, no. 1, 2014. [43] J. K. Barstow and P. G. J. Irwin, "Habitable worlds with JWST: transit spectroscopy of the TRAPPIST-1 system?," Mon. Not. R. Astron. Soc., vol. 461, no. 1, pp. L92 -- L96, 2016. [44] E. Miller-Ricci, M. R. Meyer, S. Seager, and L. Elkins-Tanton, "On the emergent spectra of hot protoplanet collision afterglows," Astrophys. J., vol. 704, no. 1, 2009. [45] K. Hamano, H. Kawahara, Y. Abe, M. Onishi, and G. L. Hashimoto, "Lifetime and spectral evolution of a magma ocean with a steam atmosphere:its detectability by future direct imaging," Astrophys. J., vol. 806, no. 2, 2015. [46] R. Seager, S.; Bains, W.; Hu, "Biosignature Gases in H2-dominated Atmospheres on Rocky Exoplanets," Astrophys. J., vol. 777, no. 2, 2013. [47] C. V. Morley, L. Kreidberg, Z. Rustamkulov, T. D. Robinson, and J. J. Fortney, "Observing the Atmospheres of Known Temperate Earth-sized Planets with JWST," Astrophys. J., vol. 850, no. 2, 2017. [48] L. Kaltenegger; Traub, Wesley A.; Jucks, "Spectral Evolution of an Earth-like Planet," Astrophys. J., vol. 658, no. 1, pp. 598 -- 616, 2007. [49] T. D. Robinson and D. C. Catling, "Common 0.1 bar tropopause in thick atmospheres set by pressure-dependent infrared transparency," Nat. Geosci., vol. 7, no. 1, pp. 12 -- 15, Dec. 2013.
1909.03001
2
1909
2019-10-23T16:32:14
Hydrogen isotopic evidence for early oxidation of silicate Earth
[ "astro-ph.EP" ]
The Moon-forming giant impact extensively melts and partially vaporizes the silicate Earth and delivers a substantial mass of metal to Earth's core. Subsequent evolution of the magma ocean and overlying atmosphere has been described by theoretical models but observable constraints on this epoch have proved elusive. Here, we report calculations of the primordial atmosphere during the magma ocean and water ocean epochs and forge new links with observations to gain insight into the behavior of volatiles on the early Earth. As Earth's magma ocean crystallizes, it outgasses the bulk of the volatiles into the primordial atmosphere. The redox state of the magma ocean controls both the chemical composition of the outgassed volatiles and the hydrogen isotopic composition of water oceans that remain after hydrogen loss from the primordial atmosphere. Whereas water condenses and is retained, molecular hydrogen does not condense and can escape, allowing large quantities (~10^2 bars) of hydrogen - if present - to be lost from Earth in this epoch. Because the escaping inventory of H can be comparable to the hydrogen inventory in the early oceans, the corresponding deuterium enrichment can be large with a magnitude that depends on the initial H2 inventory. By contrast, the common view that terrestrial water has a carbonaceous chondrite source requires the oceans to preserve the isotopic composition of that source, undergoing minimal D-enrichment via H2 loss. Such minimal enrichment places upper limits on the amount of primordial H2 in contact with early water oceans (pH2<20 bars), implies oxidizing conditions for outgassing from the magma ocean, and suggests that Earth's mantle supplied the oxidant for the chemical resorption of metals during late accretion.
astro-ph.EP
astro-ph
Hydrogen isotopic evidence for early oxidation of silicate Earth Kaveh Pahlevan1*, Laura Schaefer1, Marc M. Hirschmann2 1. School of Earth and Space Exploration, Arizona State University, Tempe, AZ, 85287, USA 2. Department of Earth Sciences, University of Minnesota, Minneapolis, MN, 55455, USA *To whom correspondence should be addressed: Email: [email protected] Tel: +1 (480) 401 8584 Fax: +1 (480) 965 8102 6 Figures Submitted to EPSL Abstract The Moon-forming giant impact extensively melts and partially vaporizes the silicate Earth and delivers a substantial mass of metal to Earth's core. The subsequent evolution of the magma ocean and overlying atmosphere has been described by theoretical models but observable constraints on this epoch have proved elusive. Here, we report thermodynamic and climate calculations of the primordial atmosphere during the magma ocean and water ocean epochs respectively and forge new links with observations to gain insight into the behavior of volatiles on the Hadean Earth. As accretion wanes, Earth's magma ocean crystallizes, outgassing the bulk of its volatiles into the primordial atmosphere. The redox state of the magma ocean controls both the chemical composition of the outgassed volatiles and the hydrogen isotopic composition of water oceans that remain after hydrogen escape from the primordial atmosphere. The climate modeling indicates that multi-bar H2-rich atmospheres generate sufficient greenhouse warming and rapid kinetics resulting in ocean-atmosphere H2O-H2 isotopic equilibration. Whereas water condenses and is mostly retained, molecular hydrogen does not condense and can escape, allowing large quantities (~102 bars) of hydrogen -- if present -- to be lost from the Earth in this epoch. Because the escaping inventory of H can be comparable to the hydrogen inventory in primordial water oceans, equilibrium deuterium enrichment can be large with a magnitude that depends on the initial atmospheric H2 inventory. Under equilibrium partitioning, the water molecule concentrates deuterium and, to the extent that hydrogen in other forms (e.g., H2) are significant species in the outgassed atmosphere, pronounced D/H enrichments (~1.5-2x) in the oceans are expected from equilibrium partitioning in this epoch. By contrast, the common view that terrestrial water has a carbonaceous chondritic source requires the oceans to preserve the isotopic composition of that source, undergoing minimal D-enrichment via equilibration with H2 followed by hydrodynamic escape. Such minimal enrichment places upper limits on the amount of primordial atmospheric H2 in contact with Hadean water oceans and implies oxidizing conditions (logfO2>IW+1, H2/H2O<0.3) for outgassing from the magma ocean. Preservation of an approximate carbonaceous chondrite D/H signature in the oceans thus provides evidence that the observed oxidation of silicate Earth occurred before crystallization of the final magma ocean, yielding a new constraint on the timing of this critical event in Earth history. The seawater-carbonaceous chondrite "match" in D/H (to ~10-20%) further constrains the prior existence of an atmospheric H2 inventory -- of any origin -- on post-giant-impact Earth to <20 bars, and suggests that the terrestrial mantle supplied the oxidant for the chemical resorption of metals during terrestrial late accretion. Keywords: silicate Earth; magma ocean; Hadean; oxidation; water; hydrogen 1. Introduction The composition and origin of Earth's early atmosphere has been debated since at least the mid-twentieth century (Brown, 1949). Recent interest arises from a desire to understand climate on the early Earth (Wordsworth and Pierrehumbert, 2013a) as well as the environment that led to abiogenesis (Kasting, 2014), and because the volatile history of Earth gives insight into the origin and evolution of the planet more generally. Here, we develop the connection between the history of the fluid envelope and that of the silicate Earth. We use ideas about Earth accretion and insights that they yield for the origin and history of terrestrial volatiles. The focus is on Earth's primordial atmosphere, a unique reservoir in Earth history that links the energetic process of planetary accretion via giant impact to the emergence of Earth's oceans and Hadean climate. We use "steam atmosphere" to describe any atmosphere prevented from condensation by internal heat in which H2O is an important component, even those in which other gases (e.g., H2) are more abundant by number. A question closely related to the atmospheric composition of the early Earth is the oxygen fugacity (fO2) of the magma ocean from which the primordial atmosphere was outgassed (Elkins-Tanton, 2008). Plausible fO2 of magma ocean outgassing range from reducing (H2-CO-rich) to oxidizing (H2O-CO2-rich) atmospheres (Hirschmann, 2012). In analogy with the volatile inventory on modern Earth, prior work has assumed that the primordial atmosphere of early Earth was H2O-CO2-rich (Abe and Matsui, 1988; Kasting, 1988; Lebrun et al., 2013; Salvador et al., 2017). However, given a lack of knowledge about the oxygen fugacity of outgassing, reducing primordial outgassed atmospheres are difficult to rule out (Abe et al., 2000; Hirschmann, 2012). Models and measurements of early Earth oxygen fugacity yield contradictory evidence: whereas the fO2 of metal- silicate equilibration is necessarily reducing (logfO2<IW-2) due to the co-existence of metals and implies H2-CO-rich gas mixtures (Gaillard et al., 2015; Wade and Wood, 2005), the oldest terrestrial samples are characterized by much higher fO2 consistent with the modern oxidized mantle and suggest an H2O-CO2-rich atmosphere (Delano, 2001; Nicklas et al., 2018; Trail et al., 2011). Some process apparently oxidized the silicate Earth during or shortly after core formation. Because the nature and timing of this process are as yet unclear, the chemical composition and oxidation state of the early Earth's atmosphere remain unknown. Here, by linking the hydrogen isotopic composition of the terrestrial oceans to the chemical composition of the primordial atmosphere, we articulate new constraints on the oxygen fugacity of magma ocean outgassing and subsequent processes on the Hadean Earth. Because this model is constrained by isotopic observations, it has the potential to yield new insights into early Earth evolution. Several features of Earth evolution imply that the D/H composition of the oceans reflects early atmospheric processes: (1) most of the water gained by Earth during planet formation was already accreted at the time of the Moon-forming giant impact and therefore participated in the terminal terrestrial magma ocean (Fischer-Gödde and Kleine, 2017; Greenwood et al., 2018), (2) most (>70%) of the water initially dissolved in the magma ocean was outgassed during solidification, rendering the steam atmosphere the dominant exchangeable water reservoir on early Earth (Elkins-Tanton, 2008), (3) the steam atmosphere was too short-lived for significant water loss via hydrodynamic escape before the condensation of the oceans (Hamano et al., 2013; Massol et al., 2016), (4) the residence time of water in the terrestrial oceans with respect to subduction and the deep water cycle is long, of order ~1010 years (van Keken et al., 2011). Together, these observations suggest that most hydrogen atoms currently residing in the oceans experienced the magma ocean and its aftermath and carry isotopic memory from early epochs. The inference that the modern oceans reflect the isotopic composition of Hadean oceans -- and, indeed, the bulk silicate Earth -- is supported by measurements on early Archean samples with D/H values for seawater identical to modern values to within a few percent (Pope et al., 2012). Percent-level variations in ocean D/H can arise due to exchange of water with the solid Earth due to plate tectonic processes (Kurokawa et al., 2018; Lécuyer et al., 1998) but here we are interested in large magnitude D/H variations (~1.5-2x) that can arise due to early atmospheric processes (see §4.3). The isotopic composition of the oceans is determined by deuterium partitioning between H2O and H2 and the contrasting histories of these molecules in the planetary environment. Hydrogen (1H) and deuterium (2D) in water and molecular hydrogen experience distinct vibrational frequencies due to different bond strengths associated with the O-H and H-H stretch. These distinct bonding environments result in deuterium being concentrated into water molecules with molecular hydrogen deuterium-depleted at equilibrium, especially at low temperatures, i.e., isotopic fractionation occurs between the water ocean and H2- rich atmosphere. Hence, even in the absence of HD/H2 mass fractionation during the escape process (Zahnle et al., 1990), loss of molecular hydrogen from the early Earth can enrich planetary water in deuterium due to equilibrium partitioning because light isotopes are preferentially concentrated into the escaping atmosphere relative to the oceans (Genda and Ikoma, 2008). As on modern Earth, water vapor is expected to be confined to the lower atmosphere via condensation below the tropopause (Wordsworth and Pierrehumbert, 2013b) and retained on the Hadean Earth whereas hydrogen in non- condensable forms (e.g., H2) can traverse the tropopause and undergo large-scale escape (Catling et al., 2001; Hunten and Strobel, 1974; Lammer et al., 2018; Tian et al., 2005). To the extent that the isotopic composition of the escaping gas was distinct from that of the oceans, isotopic evolution would have taken place. Figure 1 -- Origin of deuterium-to-hydrogen on Earth. Measured values of D/H in carbonaceous chondrites (CC) (Robert, 2003), the inferred values of the bulk silicate Earth (149+/-3 ppm) (Lécuyer et al., 1998) and the solar nebula (SN) (26+/-7 ppm) (Mahaffy et al., 1998). A nebular source for terrestrial water requires a ~six-fold D/H enrichment due to H2 escape over several Ga (Genda and Ikoma, 2008), whereas a carbonaceous chondrite source (shaded region) disallows significant (~2x) deuterium enrichment via hydrogen escape if the source signature is to be preserved. Here, we develop the consequences of such preservation for the early Earth. That the hydrogen isotopic composition of the terrestrial oceans retains memory of early epochs permits its use in constraining early atmospheric processes, pending knowledge of the D/H of the source. Indeed, the source of Earth's oceans are commonly inferred from its D/H composition in comparison with early Solar System reservoirs such as comets, asteroids, and the solar nebula. Based on the hydrogen and nitrogen isotopic evidence, the major terrestrial volatiles (C, N, H) are commonly inferred to be sourced primarily from carbonaceous chondrites (Alexander et al., 2012; Halliday, 2013; Marty, 2012). The close match (to 10-20%, Fig. 1) between terrestrial D/H and the carbonaceous chondrite distribution peak -- if not genetic -- must be relegated to coincidence with low a priori probability (Lécuyer et al., 1998). Moreover, other potential sources for Earth's major volatiles face severe difficulties. The impact probability of comets with Earth is small (~ppm) such that the inferred mass of cometary material can only contribute <10% of a terrestrial ocean, even with complete volatile retention during impact (Dauphas and Morbidelli, 2014). A nebular origin for terrestrial water has been discussed (Ikoma and Genda, 2006) but the solar neon inventory of the Earth only implies ingassing of a small fraction (<10%) of Earth's water inventory into an early magma ocean in the presence of the solar nebula (Wu et al., 2018). Moreover, a nebular source for terrestrial water would require a ~6x deuterium enrichment in the oceans (Fig. 1) due to gentle H2 escape over several Ga (Genda and Ikoma, 2008), whereas the nearly modern ocean D/H at 3.8 Ga (Pope et al., 2012) points to more limited primordial isotopic fractionation and a minor role for any nebular contribution. Finally, the nitrogen isotopic evidence argues for a carbonaceous chondrite source for Earth's major volatiles (Alexander et al., 2012; Halliday, 2013; Marty, 2012). We adopt the common view that the terrestrial oceans have a primarily carbonaceous chondritic source and show that the preservation of such a chondritic D/H signature in the oceans places an upper limit on H2 inventories on the Hadean Earth, requires oxidizing conditions for magma ocean outgassing (H2/H2O<0.3), and suggests a limited role for H2 production via the iron-water reaction during late accretion. The outline of the paper is as follows. In §2, we estimate equilibration timescales between the magma ocean and primordial atmosphere and present calculations that relate the chemical composition of the outgassed atmosphere to the oxidation state of the magma ocean. In §3, we motivate and introduce a climate model to calculate greenhouse warming by the primordial H2 inventory in the subsequent water ocean epoch. In §4, we describe the results of the climate model for the isotopic evolution of the oceans due to equilibration and loss of a primordial H2 inventory. In §5, we discuss the implications of these results for the oxidation state of the magma ocean and the oxidant for late accretion, and in §6, we summarize and conclude. The envisioned sequence explored in this paper is summarized in Figure 2. Figure 2 -- Behavior of hydrogen on Earth after the Moon-forming giant impact. (a) A deep magma ocean dissolves most of the hydrogen accreted to Earth, (b) crystallization of the magma ocean leads to outgassing of most of exchangeable hydrogen with H2/H2O determined by oxygen fugacity of last equilibration (§2.2), (c) condensation of the oceans (§3.1) and low-temperature (~300-600K) D/H equilibration (§4.3) leads to deuterium- enrichments (H2O) and depletions (H2) in co-existing species, (d) retention of water via condensation and loss of H2 via hydrodynamic escape produces deuterium enrichment in the oceans whose magnitude depends on the initial H2/H2O of the outgassed atmosphere. 2. Magma ocean outgassing Magma oceans arise in the early Solar System through various processes (Elkins-Tanton, 2012). The Moon-forming giant impact extensively melts the silicate Earth and leaves the accreting planet with ~99% of its final mass (Pahlevan and Morbidelli, 2015). During the ensuing magma ocean crystallization period, terrestrial water transitions from being predominantly dissolved in the magma ocean to primarily outgassed into the steam atmosphere, subsequent condensation of which forms the early terrestrial oceans (Elkins- Tanton, 2008; Hamano et al., 2013) (see §3.1). This event is considered the major volatile processing event in Earth history, after which the abundance and initial distribution of water is largely established and planetary processes transition from the accretionary to the geological regime. A growing body of evidence suggests that most water in the silicate Earth was initially concentrated near the surface and that Earth's deep water cycle has been characterized by a net influx of water into the mantle over geologic time, consistent with extensive early outgassing (Kendrick et al., 2017; Korenaga, 2008; Kurokawa et al., 2018; van Keken et al., 2011). In this section, we first justify the use of equilibrium thermodynamics in calculating outgassed atmospheric compositions (§2.1) and then discuss the dependence of the outgassed gaseous composition on the redox state of the magma ocean at the time of last equilibration with the atmosphere (§2.2). 2.1. Magma ocean-atmosphere equilibration timescales To justify the use of equilibrium thermodynamics, we first examine equilibration times between the magma ocean and the overlying atmosphere. On the modern Earth, the timescale for pCO2 equilibration with the oceans is ~102 years (Archer et al., 2009), but no equivalent empirical estimate exists for magma ocean-atmosphere equilibration. Using a boundary-layer analysis (Hamano et al., 2013), we estimate the timescale for magma ocean-atmosphere equilibration assuming diffusion through the magma surface boundary layer, rather than ascent of bubble plumes, dominates the equilibration process. We consider a schematic sequence in which thermal boundary layers form at the magma ocean surface and are peeled away by negative buoyancy. The equilibration timescale can then be estimated: !"#=!%&'=()*+/-)×(0/)1) where !%& is the timescale for formation of a thermal boundary layer by thermal diffusion (1) and N is the number of formation and buoyant destruction cycles before the entire magma ocean mass is processed through the surface boundary layer, dT and dC are the thermal and chemical boundary layer thicknesses, respectively, k is the thermal diffusivity of the magma (cm2 s-1), and z is magma ocean depth. We can relate the thickness of the chemical boundary layer (dC) to that of the thermal boundary layer (dT) via scaling: ()1/)*)=(3/-)4/+ where D is the atomic diffusivity for water in magma (cm2 s-1). For parameter choices, we adopt a thermal diffusivity (k=2x10-3 cm2 s-1) appropriate to peridotite liquid (Lesher and Spera, 2015) and an atomic diffusivity (Dw=1.4x10-4 cm2 s-1) appropriate for a basaltic magma with ~1 wt% water at 2000 K (Zhang et al., 2007). Substitution of these parameters yields an estimate for equilibration timescales: !"#=3×1089:;<= >?@4ABC> DEFF GBC (2) This result suggests that equilibration with the atmosphere by processing the magma ocean through a chemical boundary layer is rapid relative to the crystallization timescale, which is estimated to be ~106 years (Lebrun et al., 2013). Although more work is required to quantify the competing role of outgassing via bubble plumes in accommodating the supersaturation near the magma surface (Ikoma et al., 2018), this calculation suggests that boundary layer diffusion alone may be sufficiently rapid to motivate the equilibrium assumption. We therefore expect the magma ocean and primordial atmosphere to evolve as a coupled thermochemical system such that the properties of the system at equilibrium (e.g., the oxygen fugacity) reflect the properties of the components (e.g., the atmospheric composition). 2.2. Primordial atmospheric compositions A fundamental parameter governing equilibrium compositions of outgassed atmospheres is the oxygen fugacity (fO2). As long as the redox buffering capacity of the magma ocean exceeds that of the outgassed atmosphere, the magma ocean determines the fO2 of the atmosphere with which it equilibrates (Hirschmann, 2012). Volatile outgassing continues until the end of magma ocean crystallization, when the formation of a meters-thick solid chill crust isolates the newly formed atmosphere from rapid exchange with the silicate Earth. Therefore, the initial composition of the atmosphere during the Hadean is dictated by last equilibration with the magma ocean. Despite its importance to planetary evolution, the fO2 at the magma ocean-atmosphere interface is not well-constrained. There are two end-members: (1) in analogy with the modern Earth mantle, the magma ocean-primordial atmosphere system may be chemically oxidized (logfO2≈QFM), with water vapor and carbon dioxide dominant (Abe and Matsui, 1988; Kasting, 1988; Lebrun et al., 2013; Salvador et al., 2017). However, (2) the terrestrial magma ocean -- having held metallic droplets in suspension -- may also be much more reducing (logfO2≈IW-2) at the surface where equilibration with the atmosphere takes place. Under reducing conditions, H2 and CO are the dominant H- and C-bearing gaseous species (Fig. 3). A remarkable feature of magma oceans is that the expected range of possible fO2 values spans the transition from the reducing (H2-CO- rich) to oxidizing (H2O-CO2-rich) gas mixtures, indicating a rich volatile-processing history during early planetary evolution. Figure 3 -- High-temperature equilibrium outgassed atmospheres. The mole fraction of vapor species is calculated as a function of oxygen fugacity (fO2) relative to the iron-wüstite (IW) buffer at an equilibrium temperature of 1,400 K. Parameters for the IW buffer are given in (Frost, 1991). Thermodynamic data for gaseous species (H2O-H2-CO-CO2) are adopted from (Chase et al., 1985). The lower end of the range of logfO2 (IW-2) characterizes oxygen fugacity of a magma ocean in equilibrium with a suspension of metallic droplets, whereas the upper end of the range (IW+4≈QFM) corresponds to the redox state of the modern Earth mantle. Because hydrogen and carbon speciation reactions depend on fO2 alone, different H/C ratios yield similar results for H2/H2O and CO/CO2. Whereas an oxidizing (H2O-CO2-rich) outgassed gaseous envelope experiences minimal chemistry upon cooling and maintains its molecular composition, evolution of a reducing (H2O-CO-H2-rich) envelope may involve more significant chemical transformations. As the reducing envelope cools following magma ocean crystallization, the equilibrium reaction (3H2+COóCH4+H2O) shifts to the right, potentially converting the outgassed mixture into a methane-rich atmosphere before condensation of the oceans (Schaefer and Fegley, 2010). Such internal transformation changes the molecular composition but not atomic abundances of the envelope. However, we consider CH4 to be -- at most -- a transient molecule in the primordial atmosphere because it is unstable with respect to photolysis via Lyman a emission. The photolysis of methane under the influence of the UV flux of the young Sun occurs rapidly, at a rate of ~1 bar/Myrs (Kasting, 2014), with the carbon oxidized to CO/CO2 and the hydrogen reverting to molecular form (H2) before escaping, a somewhat more oxidizing analog to the modern atmosphere of Titan. The photochemical stability of H2 suggests that the reducing power inherited from the magma ocean is primarily carried by -- and lost via the escape of -- molecular hydrogen. Oxidizing (H2O-CO2-rich) atmospheres and their associated climates have previously been described (Wordsworth and Pierrehumbert, 2013b) but equivalent reducing (H2O- CO-H2-rich) atmospheres and climates have not. Our approach in the rest of the paper is to consider climates for a range of atmospheres more reducing than the oxidizing end- member (§3), link the chemical composition of outgassed atmospheres to the ocean D/H, and articulate constraints on reduced gas abundances in Earth's Hadean atmosphere (§4). 3. Primordial climate Calculation of planetary climate after magma ocean crystallization requires specification of the mass and molecular composition of the early atmosphere and presence or absence of water oceans. In this section, we discuss the timescales for formation of water oceans (§3.1) and introduce a model for calculating primordial climates (§3.2). 3.1. Rapid formation of water oceans Following magma ocean crystallization, the outgassed atmosphere is no longer thermally buffered and its short-term evolution can be considered decoupled from the solid Earth. We consider outgassed water inventories of 1-2 modern oceans equivalent, which is within the range of inferred abundances for the bulk silicate Earth (Hirschmann, 2006) and which can arise from a magma ocean with ~0.1 wt% H2O (Elkins-Tanton, 2008). Crystallization reduces the geothermal heat flux and leads to an atmospheric heat flux below the runaway greenhouse threshold (Hamano et al., 2013), at which point the stable climate state requires condensation of the steam into oceans. On a rapid timescale relative to escape (>0.1-10 Ma, Fig. 5), the atmosphere behaves as a closed system. Heat inherited from the magma ocean powers atmospheric secular cooling for a timescale: !AHHI=J@1K∆* MNOP*QR (3) with PT the total atmospheric pressure, CP the thermal heat capacity, ∆T the atmospheric temperature difference between the magma ocean and water ocean epochs, g the surface gravity, sSB the Stefan-Boltzmann constant, and TE the effective temperature for cooling. In a steam atmosphere where opacity is dominated by water vapor, the outgoing radiation flux (sSBTE4) has a lower limit (~300 W/m2) due to condensation (Nakajima et al., 1992). Adopting PT=270-540 bars appropriate for a steam atmosphere with 1-2 modern oceans of water and ∆T=103 K for cooling from a steam atmosphere (1,500 K) (Matsui and Abe, 1986) to a warm ocean (500 K) (Fig. 4) yields a secular cooling time of ~103 years. Latent heat of condensation of water only contributes comparable heat as the sensible heat to the heat budget. Condensation of a steam atmosphere is rapid relative to the timescale of escape (Fig. 5). Following magma ocean crystallization, outgassed H2O rapidly condenses and the Hadean Earth relaxes into a climate with oceans whose temperature is determined via greenhouse warming by the primordial atmosphere. 3.2. Reducing climate model In this section we introduce a climate model to calculate surface temperatures due to H2 greenhouse warming, which we use in the next section to determine the behavior of hydrogen and deuterium on the Hadean Earth (§4.3). To describe the earliest Hadean climate, we adopt a 2-component (H2O-H2) model for the atmosphere and ocean and take the oxygen fugacity of last equilibration between the magma ocean and primordial atmosphere -- and hence the outgassed H2/H2O -- as a free parameter (Fig. 3). We select a 2-component (H2O-H2) model rather than more complex models (e.g. H2O-CO-H2) to describe a reduced outgassed ocean and atmosphere because: (i) carbon monoxide is not an effective greenhouse gas and is expected to have a secondary influence on calculated surface temperatures (§4.1), (ii) a hydrodynamic hydrogen wind might drag CO to space via molecular collisions and prolong the lifetime of the primordial atmosphere against escape (Zahnle and Kasting, 1986), only strengthening the conclusion that ocean- atmosphere isotopic equilibration was maintained during the escape process (§4.2), (iii) outgassed CO that does not escape is oxidized to CO2 (Kasting, 1990) producing more molecular hydrogen (CO+H2OèCO2+H2). The net effect of incorporating carbon would be to increase the inventory of escaping H2 and deuterium enrichment at any given fO2 of outgassing. However, below we show that even equilibrium deuterium-enrichments calculated by neglecting carbon have sufficient magnitude to yield new constraints on the conditions of primordial outgassing (§4.3). Because water condenses in the lower atmosphere and is retained but molecular hydrogen can escape, the free parameter governing early climate can be expressed as the inventory of molecular hydrogen in the atmosphere (pH2). To translate H2/H2O ratios derived from outgassing to atmospheric H2 inventories, we scale by the mass of the ocean reservoir, which we assume to be 1-2 modern ocean equivalents (§3.1) Due to water condensation and cold-trapping in the lower atmosphere and collision- induced infrared opacity of H2 at moderately high (>0.1 bar) pressures, the emission level in H2O-H2 model atmospheres is determined by the opacity of H2 (Wordsworth, 2012). The emission temperature (TE) is given by top-of-the-atmosphere radiative balance with the early Sun: &S(1−U)=VW%XYS (4) with L the solar constant, A the visible bond albedo, and sSB the Stefan-Boltzmann constant. For L=103 W/m2 appropriate for the early Sun and a bond albedo A=0.3, an emission temperature of 235 K is obtained for the primordial Earth. For simplicity, solar luminosity in these calculations is held constant. Results are qualitatively similar for a range of emission temperatures (215-255 K), as might be expected based on an evolving Sun, cloud feedback and/or by adjusting the planetary albedo to account for Rayleigh scattering in thicker atmospheres. These effects are known to alter the radiation budget by tens of percent (Gough, 1981; Wordsworth, 2012). At infrared wavelengths, the mean optical depth unity surface of a pure H2 atmosphere has been calculated for a several Earth-mass planet (g=20 m/s2, Tph=100 K) and is ~0.2 bars (Wordsworth, 2012). Combining the expressions for photospheric pressure (Pph∝g/k) and collision-induced opacity (k∝r) of an ideal gas (r∝Pph/Tph) yields a scaling relation between photospheric pressure, gravity, and photospheric temperature (Pphµg1/2Tph1/2). Applying this scaling relation to emission temperatures (Tph=235K) relevant to the Hadean Earth (g=9.8 m/s2) yields an H2 photospheric pressure ~0.21 bars, which we adopt. Since the atmosphere at the emission level is cold and dry, this pressure -- appropriate for a pure H2 atmosphere -- is taken as the emission pressure to which a moist adiabatic structure must be stitched. For any given H2 inventory, we use the "all-troposphere" approximation to solve for the surface temperature that yields the emission temperature at the photosphere required by radiative balance (see Supplementary Information). This approximation has been shown to be accurate in describing greenhouse warming in H2-rich atmospheres (Pierrehumbert and Gaidos, 2011). 4. Results In this section, we describe the results of the climate model (§4.1) and use the results to compare the timescales for ocean-atmosphere isotopic equilibration with atmospheric escape (§4.2) to motivate equilibrium isotopic partitioning to describe the behavior of deuterium on the early Earth. Finally, we discuss the hydrogen isotopic evolution of the Hadean oceans and the derived upper limit on pH2 on the early Earth (§4.3). 4.1. Equilibrium surface temperatures Greenhouse warming by inventories of ~1-102 bars of equivalent H2 on the Hadean Earth yields surface temperatures ~300-550 K (Fig. 4). Despite some uncertainty arising from cloud feedback or higher albedo due to Rayleigh scattering, multi-bar H2 inventories are sufficient to keep Earth out of a snowball state via H2 opacity alone. The calculation assumes radiative transfer in the terrestrial planet regime, where a sufficient fraction of the visible flux penetrates the atmosphere and powers convection throughout the troposphere, whereas for pH2>20 bars visible photons might not penetrate, creating a deep atmospheric structure governed by geothermal rather than solar heating as in the interiors of the giant planets (Wordsworth, 2012). Calculated surface temperatures for pH2>20 bars are therefore upper limits. These calculations suggest that the dominant control on surface temperatures in the early Hadean is the H2 inventory and that earliest climate on Earth was warm and governed by the physics of atmospheric escape. Figure 4 -- Ocean surface temperature as a function of H2 inventory. Temperatures are shown for different values of the outgoing longwave radiation (OLR) flux for emission temperatures (TE) of 215, 235, and 255 K, corresponding to a range (0-0.5) of bond albedos. The primordial climate depends primarily on the H2 inventory. As has also been found in the case of H2-CO2 early Martian atmospheres (Ramirez et al., 2014), one bar partial pressure of H2 is entirely sufficient to stabilize a warm and wet early climate. The dominant control on earliest climate in these scenarios is the physics of atmospheric escape. 4.2. Atmospheric equilibration and escape timescales Surface temperatures derived from the climate model can be used to assess H isotopic equilibration between the oceans and atmosphere. A key comparison is between the equilibration time and the residence time of the atmospheric H2 inventory. H2O-H2 deuterium exchange may be rate-limited by exchange in the atmosphere, because H2 dissolves negligibly in the oceans. Deuterium exchange between water vapor and molecular hydrogen is rapid and highly temperature-dependent (Lécluse and Robert, 1994). After condensation of the steam into oceans (§3.1), a significant H2 inventory (>10-100 bars H2) and high surface temperatures (>400-550 K, Fig. 4) result in extremely rapid atmospheric reactions such that the timescale for ocean-atmosphere equilibration is limited by the rate of water evaporation and circulation in a hydrological cycle. This timescale ((cid:2)depth of oceans/evaporation rate) is rapid (≈103 years), and nominally independent of H2 inventory and surface temperature (e.g. pH2>3 bars, Fig. 5). Ocean- atmosphere isotopic equilibration is therefore expected to initially proceed rapidly. As climate evolves due to H2 escape, reaction rates decline until ocean-atmosphere equilibration becomes rate-limited by exchange reactions in the atmosphere. In this ![\=(!"]/!^"_)∙ !Aa^ regime, the equilibration time between the oceans and atmosphere is given by: with tex the isotopic exchange time between atmospheric H2 and H2O, !^"_ the residence time of atmospheric water vapor, and !Aa^ the timescale to circulate the entire oceans (5) through the atmosphere via evaporation and precipitation in a hydrological cycle (Genda and Ikoma, 2008). As the H2 inventory is lost, the ocean-atmosphere equilibration time (tAO) evolves for two reasons: (1) the isotopic exchange timescale (tex) increases due to the cooling temperatures, (2) the residence time of water vapor in the atmosphere ((cid:2) depth of equivalent water layer/precipitation rate) decreases due to the lower water vapor pressures associated with lower temperature. Both effects prolong the equilibration time between the ocean and atmosphere, which increases rapidly at low temperatures (pH2<3 bars, Fig. 5), approaching ~1 Myrs for H2 inventories of ~1 bar. Figure 5 -- Ocean-atmosphere isotopic equilibration and hydrodynamic escape versus H2 inventory. Isotopic equilibration between the water oceans and a primordial H2-rich atmosphere is calculated for three values of top-of-the- atmosphere emission temperature (TE) as a function of H2 inventory. At high H2 inventories (e.g. 10-100 bars equivalent H2), equilibrium temperatures are high (~400-500K, Fig. 4), and isotopic exchange reactions are extremely fast. Parameters characterizing the escape times are the thermal efficiency (e) and the planetary EUV absorption radius (x) (§4.2). Timescales for ocean- atmosphere equilibration are generally shorter than the residence time of atmospheric H2 with respect to escape suggesting continuous equilibration during the loss process. Ocean-atmosphere isotopic equilibration requires that the equilibration time be shorter than the residence time of H2 in the atmosphere. To determine whether or not this is the case, we calculate extreme ultraviolet (EUV) powered escape rates (Watson et al., 1981) for H2 inventories, and assume that loss to space is the sole H2 sink, as expected on the (Pa s-1) prebiotic Earth. Escape rates can be calculated using the energy-limited approximation: with b the hydrogen escape flux expressed as atmospheric pressure loss per unit time (6) bcd=efQgh Sij k+ (Wordsworth, 2012), FEUV the extreme ultraviolet flux of the young sun, e the thermal efficiency or the fraction of incident EUV used to power the planetary wind (e=0.1-0.2), Rp the planetary radius, and x((cid:2)REUV/Rp) the effective absorption radius in planetary radii (x=2-3), which characterizes the distended nature of EUV absorption in escaping atmospheres. FEUV is assumed equal to 100 times the modern extreme ultraviolet flux, i.e. 464 erg cm-2 s-1 (Ribas et al., 2005). Other parameter choices (e,x) are adopted from simulations of hydrogen-rich atmospheres exposed to comparable EUV fluxes (Erkaev et al., 2016; Shematovich et al., 2014). Calculated residence times ((cid:2)pH2/fH2) for H2 are ≈ 106-107 years for H2 inventories ≈ 10-100 bars (Fig. 5). Timescales for ocean- atmosphere equilibration are generally shorter than the residence time of atmospheric H2, suggesting continuous equilibration with ocean-atmosphere quenching occurring at H2 inventories of a few bars or less. 4.3. Hydrogen isotopic evolution of the Hadean oceans Finally, we quantify the behavior of hydrogen (1H) and deuterium (2D) as tracers of early Earth evolution to articulate constraints on the chemical composition of the primordial atmosphere. In brief, water vapor is retained via condensation but hydrogen in non- condensable form (e.g. H2) interacts with water vapor in the lower atmosphere and is transported to the upper atmosphere and lost to space. Equilibrium D/H partitioning between water and hydrogen is calculated from standard prescriptions (Richet et al., 1977).1 Although this prescription strictly relates water vapor to molecular hydrogen, it can be used to characterize equilibrium between water oceans and H2-rich atmospheres because the vapor pressure isotope effect relating liquid water to water vapor is an order of magnitude smaller and can be neglected to first order. H2O-H2 equilibration is among the largest equilibrium fractionations between two molecules in nature, with a clearly resolvable magnitude at planetary temperatures. D/H composition of planetary oceans reflect the mass of early H2 reservoirs and the temperatures of isotopic equilibration. To the extent that an atmospheric H2 reservoir comparable to the hydrogen in the terrestrial oceans was present, the ocean D/H could have evolved dramatically due to equilibrium partitioning and removal of isotopically light H2. The hydrogen isotopic evolution of Hadean oceans can be calculated using equilibrium ocean-atmosphere partitioning. Following magma ocean crystallization, most of the water vapor condenses into oceans, while most H2 partitions into the atmosphere, generating the earliest climate (§3). The magnitude of greenhouse warming for a freshly outgassed atmosphere is significant: only a few percent of the outgassed inventory need be in the form of molecular hydrogen to prevent a snowball Earth and to stabilize water oceans via H2 greenhouse warming alone (Fig. 4). The existence of such an early greenhouse climate permits isotopic equilibration between the ocean and atmosphere (§4.2) with the temperature-dependent partitioning between reservoirs determined self- consistently via climate with a given H2 inventory (§3). As the atmospheric H2 inventory is depleted via escape, the greenhouse effect also diminishes, accentuating the temperature-dependent partitioning between ocean and atmosphere. Hence, deuterium is further concentrated into the oceans due to the cooling radiative balance accompanying 1 The exchange reaction is H2O+HDóHDO+H2 with equilibrium constant K(T)=1+0.22*(103/T)2. H2 loss. In this way, D/H evolution of the oceans can be calculated as an equilibrium distillation sequence, converging to a value determined by the initial inventory of H2. Water molecules have a strong tendency to concentrate deuterium, and, to the extent that H2 is abundant in the primordial atmosphere, pronounced D/H enrichments (~1.5-2x) can arise from equilibration and H2 escape in this epoch (Fig. 6), an enrichment not evident in Earth's chondritic oceans. Such deuterium enrichment is minimized for a pure steam atmosphere outgassed from an oxidizing magma ocean. Preservation of a chondritic D/H signature in the oceans to within <20% thereby constrains H2/H2O of Earth's outgassed atmosphere to <0.3. For a given H2O abundance, e.g., 1-2 modern ocean equivalents (Korenaga, 2008), upper limits on the initial H2 inventory of Earth (pH2<10-20 bars) can be derived (Fig. 6). Figure 6 -- Ocean deuterium enrichment versus oxygen fugacity of primordial outgassing. Oxygen fugacity determines the H2/H2O of the outgassed atmosphere. Oxidizing conditions lead to nearly pure steam atmospheres and minimal ocean D/H enrichment whereas reducing outgassing (e.g., logfO2<IW) generates higher H2/H2O (>1) and stronger ocean D/H enrichments (>1.5-2x) due to equilibration and loss of large quantities of deuterium-depleted H2. Minimal (<20%) deuterium-enrichment in terrestrial water relative to the source (§5) constrains H2/H2O of the outgassed atmosphere to <0.3 and logfO2 of outgassing to >IW+1 (Permitted region). Tolerable H2 abundances can be expressed in absolute terms (top axis, pH2<10-20 bars) by fixing water abundances to 1-2 modern ocean equivalents. Temperatures of ocean- atmosphere equilibration are calculated via the climate model (§3.2), with the enrichment curves corresponding to different emission temperatures (TE=215, 235, 255K), demonstrating the robustness of the result to plausible variations in early climate. Here, we have only considered ocean-atmosphere partitioning followed by the removal of atmospheric H2 with no isotopic fractionation during the escape process. Kinetic processes (e.g. HD/H2 mass-fractionation) could fractionate H isotopes further (Zahnle et al., 1990) and cause additional ocean deuterium enrichment for a given H2 inventory although a sufficiently vigorous hydrogen escape flux would produce a subdominant or negligible kinetic contribution to the D/H enrichment relative to equilibrium partitioning (Genda and Ikoma, 2008). The constraints on primordial H2 abundances shown in Figure 6 are therefore strictly speaking upper limits. We compare these results to those obtained previously. For initial outgassed envelopes with H2/H2O=1-10, ocean D/H enrichments of 2-3x due to equilibrium partitioning alone and 2-9x by inclusion of HD/H2 kinetic mass fractionation during escape have been reported (Genda and Ikoma, 2008). These previous results were obtained by assuming ocean-atmosphere equilibration occurred at 300 K for all H2 inventories (pH2≃30-300 bars for initial H2/H2O=1-10). By contrast, we find higher temperatures of ocean-atmosphere equilibration (Ts=475-550 K for pH2=30- 100 bars, Fig. 4). By calculating climate and temperature-dependent isotopic partitioning self-consistently, we expect our model to yield a more accurate description of the equilibrium behavior of deuterium on the Hadean Earth. We find ocean deuterium enrichments of 1.5-1.8x (Fig. 6) for initial H2 inventories of 30-300 bars (H2/H2O≈1-10) corresponding to reducing conditions for primordial outgassing (logfO2=IW to IW-2). We do not attempt to estimate kinetic mass fractionation during escape but note that the near constancy of terrestrial D/H in the last 3.8 Ga (Pope et al., 2012) restricts any primordial H2 inventory to the Hadean Earth and implies a vigorous hydrogen wind with a subdominant role for kinetics (Genda and Ikoma, 2008). Calculated enrichments are therefore a conservative but robust constraint on minimum ocean deuterium enrichment for a given past H2 inventory. In summary, calculated ocean D/H enrichments (1.5-1.8x, Fig. 6) are somewhat smaller than those predicted by (Genda and Ikoma, 2008) via equilibrium partitioning alone (2-3x, their Fig. 6) because greenhouse warming reduces equilibrium isotopic fractionation and permits more deuterium to partition into escaping H2 in our calculations. Neither our calculated equilibrium enrichments (1.5-1.8x) nor those reported previously (2-3x) (Genda and Ikoma, 2008) for H2-rich outgassing are sufficiently high to reconcile a nebular source with the terrestrial oceans, supporting the carbonaceous chondrite origin of the terrestrial hydrosphere. Nevertheless, calculated ocean deuterium enrichments are sufficiently large to yield new constraints on Hadean evolution, which we discuss next. 5. Discussion The molecular composition of Earth's primordial atmosphere is not well-constrained. Nevertheless, on the basis of their isotopic compositions, Earth's major volatiles (H, N, C) are thought to be sourced primarily from carbonaceous chondrites (Alexander et al., 2012; Halliday, 2013; Marty, 2012). This widely-held view of the source of major terrestrial volatiles requires preservation of the source signature in the terrestrial oceans and implies minimal D/H enrichment via equilibration and escape of primordial H2. To quantify the constraint that this comparison places on primordial outgassing, we compare the isotopic composition (dD=-25 +/- 5‰) of the Archean oceans (Pope et al., 2012) with the lowest bulk chondritic values (dD=-226 +/- 4‰) measured to date (Alexander et al., 2012). On this basis, a primarily chondritic source for terrestrial water requires <20% deuterium-enrichment via H2 loss. According to our calculations, this level of isotopic preservation requires most early outgassed hydrogen from the Earth to appear in the form of water vapor (H2/H2O<0.3, pH2<10-20 bars, Fig. 6). By connecting the conditions of outgassing to the observable isotopic signatures in ancient seawater, we can articulate new constraints on the composition of the Earth's primordial atmosphere. These results show that reducing gases such as H2 and CH4 made up only a minor fraction of the Earth's outgassed atmosphere and require the terminal magma ocean to be oxidized by the time of last equilibration with the atmosphere. In this section, we discuss the implications of these results for redox evolution of the magma ocean (§5.1) and the oxidant involved in terrestrial late accretion (§5.2). 5.1. The redox state of the terrestrial magma ocean The redox state of a magma ocean determines both the chemical composition of the outgassed atmosphere and the isotopic composition of water oceans following primordial H2 loss. Oxidizing magma oceans outgas water-rich primordial atmospheres, which condense into oceans, experiencing minimal hydrogen escape and deuterium enrichment. Reducing magma oceans, by contrast, outgas substantial quantities of hydrogen as H2 in addition to water molecules whose equilibration with water oceans before loss can enrich the oceans in deuterium by ~1.5-2x relative to initial values, a feature not evident in the terrestrial isotopic record (Pope et al., 2012). The persistence of a chondritic signature in the terrestrial oceans requires a low outgassed H2/H2O (<0.3) and oxidizing conditions for last atmospheric equilibration with the magma ocean (logfO2>IW+1) (Fig. 6). Given that the convective magma ocean initially held metallic droplets in suspension (Stevenson, 1990) and was therefore chemically reducing (logfO2<IW-2) at early times, these results suggest that the silicate Earth was oxidized during the evolution of the magma ocean. Three mechanisms have been discussed for this primordial oxidation: (1) the terrestrial magma ocean could have been oxidized via Fe disproportionation at high pressure (3Fe+2è2Fe+3+Fe0) with separation of the newly generated metallic iron to the core leaving an oxidized mantle residue (Hirschmann, 2012; Wade and Wood, 2005), (2) the primordial atmosphere was reducing (H2-rich) but the process of H2 escape during the lifetime of the magma ocean oxidized both the atmosphere and the co-existing silicate Earth (Hamano et al., 2013), and (3) the Fe+3/Fe+2 value of the terrestrial magma ocean was low (~0.01) but the more incompatible nature of ferric iron (Fe+3) in mantle minerals enriched it in evolving liquids such that the late-stage magma ocean was more oxidizing than that at the outset of crystallization (Schaefer et al. submitted). The relative importance of these processes for the redox evolution of magma oceans is subject to future study. For now, we conclude that the oxidation of the silicate Earth occurred during the crystallization of the magma ocean, independently corroborating the conclusion from geological data for early (>3.5-4 Gya) establishment of oxidizing conditions in Earth's mantle (Delano, 2001; Nicklas et al., 2018; Trail et al., 2011). 5.2. The oxidant for terrestrial late accretion Before outgassing of the primordial atmosphere, the magma ocean potentially facilitates the last major episode of core formation via separation of metallic droplets accompanying deep magma ocean convection on rapid (~102 year) timescales (Stevenson, 1990). Such metallic droplets strongly concentrate and efficiently scavenge highly siderophile elements (HSEs) from the terrestrial magma ocean and sequester them into the metallic core. Mantle relative abundances of HSEs resemble the chondrites, leading to the notion that these elements were delivered during the final ~1% of Earth accretion, after cessation of core formation (Kimura et al., 1974), now interpreted as accretion after the Moon- forming giant impact. Isotopic characteristics of the Earth's mantle suggest delivery by bodies with metals either as undifferentiated metallic grains or as planetesimal cores (Marchi et al., 2018). However, the terrestrial upper mantle is currently unsaturated in metallic Fe, instead exhibiting a more oxidizing redox state, indicated by higher Fe+3/Fe+2 values than those characterizing co-existence with metallic iron. Accordingly, accreted metals must have been oxidized and dissolved into Earth's mantle, prompting the question of the nature of the oxidant involved in late accretion. Since the terrestrial magma ocean crystallized on ~106 year timescales (Lebrun et al., 2013) while the leftovers of planetary accretion were swept up over ~107-108 years (Morbidelli et al., 2012), late accretion likely occurred onto a silicate Earth with water oceans. Possibilities for oxidizing the metals delivered during late accretion are: (1) Earth's fluid envelope, e.g., water in the terrestrial oceans, via the iron-water reaction (Fe+H2OèFeO+H2) followed by hydrodynamic escape of H2 (Genda et al., 2017) and (2) the oxidizing power of Earth's mantle, epitomized by ferric iron (2Fe+3+Fe0è3Fe+2), lowering the ferric iron abundance to the modern upper mantle value (Fe+3/Fe+2=0.03-0.04) (Canil et al., 1994). Using calculated enrichments in the D/H of water oceans coexisting with significant (~10-100 bar) early H2 atmospheric inventories, we can limit the extent of the iron-water reaction in oxidizing the metals of late accretion. Given that the oxidation of metals at that time would consume more than a modern ocean worth of water via the iron-water reaction (producing >30 bars H2) and that the Hadean H2 inventory was apparently <10- 20 bars (Fig. 6), we conclude that the iron-water reaction had a subdominant role during late accretion. A more dominant role for this reaction would have produced deuterium- enriched oceans not observed in the terrestrial record. This reasoning suggests that the terrestrial mantle supplied oxidants during late accretion, a feature that may yield insights into the physics and chemistry of this early process. 6. Conclusions The isotopic composition of the oceans provides a unique constraint for early planetary evolution. It is widely accepted that most water accreted by the Earth was delivered before the Moon-forming giant impact and that most water dissolved in the subsequent magma ocean was excluded from crystallizing minerals and outgassed into a primordial atmosphere (Elkins-Tanton, 2008; Greenwood et al., 2018). Given that the residence time of water in Earth's oceans with respect to the deep water cycle is comparable to, or greater than, the current age of the Earth (van Keken et al., 2011), most of the hydrogen atoms in the oceans today are inferred to be outgassed from the magma ocean, retaining isotopic memory of early epochs. The oxygen fugacity of terrestrial magma ocean outgassing -- and therefore the chemical composition of the primordial atmosphere -- has not been independently constrained. By linking the oxygen fugacity of primordial outgassing to the deuterium content of Earth's hydrosphere, we articulate new constraints on these critical parameters governing early Earth evolution. We find that preservation of a carbonaceous chondritic D/H signature in the terrestrial oceans (to 10-20%) requires Earth's outgassed envelope be hydrogen- poor (H2/H2O<0.3), indicating oxidizing conditions (logfO2>IW+1) for last equilibration with the magma ocean. We infer that oxidation of the silicate Earth took place during the evolution of Earth's final magma ocean, and may require no geological oxidation processes (e.g. subduction) to be consistent with an oxidized mantle observed in the earliest terrestrial record (Trail et al., 2011). The inferred absence of massive (>20 bar) H2 inventories of any origin on the Hadean Earth constrains the oxidant for terrestrial late accretion. Whereas the likely existence of early water oceans has previously been taken to imply that the iron-water reaction was responsible for oxidizing the metals delivering HSEs to early Earth (Genda et al., 2017), we find that such massive production of molecular hydrogen would have disturbed the carbonaceous-chondrite-like signature of the terrestrial oceans. We therefore infer that oxidants in the terrestrial mantle (e.g., Fe+3) were responsible for oxidative resorption of late-accreting metals delivered to the silicate Earth. Indeed, the oxidative potential in Earth's modern mantle is comparable to the reducing potential in ~0.5% of an Earth mass of chondritic late accretion, a feature that may yield insight into this early terrestrial process. Acknowledgements K.P. acknowledges support from a grant from the W.M. Keck foundation. The authors are grateful to Peter Buseck for detailed comments on an early draft, and to Fabrice Gaillard and two anonymous reviewers for thorough reviews that greatly helped to improve the manuscript. References Abe, Y., Matsui, T., 1988. Evolution of an impact-generated H2O -- CO2 atmosphere and formation of a hot proto-ocean on Earth. Journal of the Atmospheric Sciences 45, 3081-3101. Abe, Y., Ohtani, E., Okuchi, T., Righter, K., Drake, M., 2000. Water in the early Earth. Origin of the Earth and Moon 1, 413-433. Alexander, C.O.D., Bowden, R., Fogel, M., Howard, K., Herd, C., Nittler, L., 2012. The provenances of asteroids, and their contributions to the volatile inventories of the terrestrial planets. Science 337, 721-723. Archer, D., Eby, M., Brovkin, V., Ridgwell, A., Cao, L., Mikolajewicz, U., Caldeira, K., Matsumoto, K., Munhoven, G., Montenegro, A., Tokos, K., 2009. Atmospheric Lifetime of Fossil Fuel Carbon Dioxide. Annual Review of Earth and Planetary Sciences 37, 117-134. Brown, H., 1949. Rare gases and the formation of the Earth's atmosphere, The Atmospheres of the Earth and Planets, p. 258. Canil, D., O'Neill, H.S.C., Pearson, D., Rudnick, R., McDonough, W., Carswell, D., 1994. Ferric iron in peridotites and mantle oxidation states. Earth and Planetary Science Letters 123, 205-220. Catling, D.C., Zahnle, K.J., McKay, C.P., 2001. Biogenic Methane, Hydrogen Escape, and the Irreversible Oxidation of Early Earth. Science 293, 839-843. Chase, M.W., Davies, C.A., Downey, J.R., Frurip, D.J., Mcdonald, R.A., Syverud, A.N., 1985. Janaf Thermochemical Tables - 3rd Edition .2. Journal of Physical and Chemical Reference Data 14, 927-1856. Dauphas, N., Morbidelli, A., 2014. Geochemical and planetary dynamical views on the origin of Earth'atmosphere and oceans. Treatise Geochem 6, 1-35. Delano, J.W., 2001. Redox history of the Earth's interior since∼ 3900 Ma: implications for prebiotic molecules. Origins of Life and Evolution of the Biosphere 31, 311-341. Elkins-Tanton, L.T., 2008. Linked magma ocean solidification and atmospheric growth for Earth and Mars. Earth and Planetary Science Letters 271, 181-191. Elkins-Tanton, L.T., 2012. Magma Oceans in the Inner Solar System. Annual Review of Earth and Planetary Sciences 40, 113-139. Erkaev, N., Lammer, H., Odert, P., Kislyakova, K., Johnstone, C., Güdel, M., Khodachenko, M., 2016. EUV-driven mass-loss of protoplanetary cores with hydrogen-dominated atmospheres: the influences of ionization and orbital distance. Monthly Notices of the Royal Astronomical Society 460, 1300-1309. Fischer-Gödde, M., Kleine, T., 2017. Ruthenium isotopic evidence for an inner Solar System origin of the late veneer. Nature 541, 525-527. Frost, B.R., 1991. Introduction to oxygen fugacity and its petrologic importance. Reviews in Mineralogy and Geochemistry 25, 1-9. Gaillard, F., Scaillet, B., Pichavant, M., Iacono-Marziano, G., 2015. The redox geodynamics linking basalts and their mantle sources through space and time. Chemical Geology 418, 217-233. Genda, H., Brasser, R., Mojzsis, S.J., 2017. The terrestrial late veneer from core disruption of a lunar-sized impactor. Earth and Planetary Science Letters 480, 25- 32. Genda, H., Ikoma, M., 2008. Origin of the ocean on the Earth: Early evolution of water D/H in a hydrogen-rich atmosphere. Icarus 194, 42-52. Gough, D., 1981. Solar interior structure and luminosity variations. Solar Physics 74, 21-34. Greenwood, R.C., Barrat, J.-A., Miller, M.F., Anand, M., Dauphas, N., Franchi, I.A., Sillard, P., Starkey, N.A., 2018. Oxygen isotopic evidence for accretion of Earth's water before a high-energy Moon-forming giant impact. Science Advances 4. Halliday, A.N., 2013. The origins of volatiles in the terrestrial planets. Geochimica Et Cosmochimica Acta 105, 146-171. Hamano, K., Abe, Y., Genda, H., 2013. Emergence of two types of terrestrial planet on solidification of magma ocean. Nature 497, 607-610. Hirschmann, M.M., 2006. Water, Melting, and the Deep Earth H2O Cycle. Annual Review of Earth and Planetary Sciences 34, 629-653. Hirschmann, M.M., 2012. Magma ocean influence on early atmosphere mass and composition. Earth and Planetary Science Letters 341 -- 344, 48-57. Hunten, D.M., Strobel, D.F., 1974. Production and escape of terrestrial hydrogen. Journal of the Atmospheric Sciences 31, 305-317. Ikoma, M., Elkins-Tanton, L., Hamano, K., Suckale, J., 2018. Water Partitioning in Planetary Embryos and Protoplanets with Magma Oceans. Space Science Reviews 214, 76. Ikoma, M., Genda, H., 2006. Constraints on the Mass of a Habitable Planet with Water of Nebular Origin. The Astrophysical Journal 648, 696. Kasting, J.F., 1988. Runaway and moist greenhouse atmospheres and the evolution of Earth and Venus. Icarus 74, 472-494. Kasting, J.F., 1990. Bolide impacts and the oxidation state of carbon in the Earth's early atmosphere. Origins of Life and Evolution of the Biosphere 20, 199-231. Kasting, J.F., 2014. Atmospheric composition of Hadean -- early Archean Earth: The importance of CO, in: Shaw, G.H. (Ed.), Earth's Early Atmosphere and Surface Environment. Geological Society of America. Kendrick, M., Hémond, C., Kamenetsky, V., Danyushevsky, L., Devey, C.W., Rodemann, T., Jackson, M., Perfit, M., 2017. Seawater cycled throughout Earth's mantle in partially serpentinized lithosphere. Nature Geoscience 10, 222. Kimura, K., Lewis, R.S., Anders, E., 1974. Distribution of gold and rhenium between nickel-iron and silicate melts: implications for the abundance of siderophile elements on the Earth and Moon. Geochimica et Cosmochimica Acta 38, 683-701. Korenaga, J., 2008. Plate tectonics, flood basalts and the evolution of Earth's oceans. Terra Nova 20, 419-439. Kurokawa, H., Foriel, J., Laneuville, M., Houser, C., Usui, T., 2018. Subduction and atmospheric escape of Earth's seawater constrained by hydrogen isotopes. Earth and Planetary Science Letters 497, 149-160. Lammer, H., Zerkle, A.L., Gebauer, S., Tosi, N., Noack, L., Scherf, M., Pilat-Lohinger, E., Güdel, M., Grenfell, J.L., Godolt, M., 2018. Origin and evolution of the atmospheres of early Venus, Earth and Mars. The Astronomy and Astrophysics Review 26, 2. Lebrun, T., Massol, H., Chassefière, E., Davaille, A., Marcq, E., Sarda, P., Leblanc, F., Brandeis, G., 2013. Thermal evolution of an early magma ocean in interaction with the atmosphere. Journal of Geophysical Research: Planets 118, 1155-1176. Lécluse, C., Robert, F., 1994. Hydrogen isotope exchange reaction rates: Origin of water in the inner solar system. Geochimica et Cosmochimica Acta 58, 2927-2939. Lécuyer, C., Gillet, P., Robert, F., 1998. The hydrogen isotope composition of seawater and the global water cycle. Chemical Geology 145, 249-261. Lesher, C.E., Spera, F.J., 2015. Thermodynamic and transport properties of silicate melts and magma, The Encyclopedia of Volcanoes (Second Edition). Elsevier, pp. 113-141. Mahaffy, P., Donahue, T.M., Atreya, S., Owen, T., Niemann, H., 1998. Galileo probe measurements of D/H and 3He/4He in Jupiter's atmosphere. Space Science Reviews 84, 251-263. Marchi, S., Canup, R., Walker, R., 2018. Heterogeneous delivery of silicate and metal to the Earth by large planetesimals. Nature Geoscience 11, 77. Marty, B., 2012. The origins and concentrations of water, carbon, nitrogen and noble gases on Earth. Earth and Planetary Science Letters 313, 56-66. Massol, H., Hamano, K., Tian, F., Ikoma, M., Abe, Y., Chassefière, E., Davaille, A., Genda, H., Güdel, M., Hori, Y., 2016. Formation and evolution of protoatmospheres. Space Science Reviews 205, 153-211. Matsui, T., Abe, Y., 1986. Evolution of an impact-induced atmosphere and magma ocean on the accreting Earth. Nature 319, 303-305. Morbidelli, A., Marchi, S., Bottke, W.F., Kring, D.A., 2012. A sawtooth-like timeline for the first billion years of lunar bombardment. Earth and Planetary Science Letters 355, 144-151. Nicklas, R.W., Puchtel, I.S., Ash, R.D., 2018. Redox state of the Archean mantle: Evidence from V partitioning in 3.5 -- 2.4 Ga komatiites. Geochimica et Cosmochimica Acta 222, 447-466. Pahlevan, K., Morbidelli, A., 2015. Collisionless encounters and the origin of the lunar inclination. Nature 527, 492-494. Pierrehumbert, R., Gaidos, E., 2011. Hydrogen Greenhouse Planets Beyond the Habitable Zone. The Astrophysical Journal Letters 734, L13. Pope, E.C., Bird, D.K., Rosing, M.T., 2012. Isotope composition and volume of Earth's early oceans. Proceedings of the National Academy of Sciences 109, 4371-4376. Ramirez, R.M., Kopparapu, R., Zugger, M.E., Robinson, T.D., Freedman, R., Kasting, J.F., 2014. Warming early Mars with CO2 and H2. Nature Geosci 7, 59-63. Ribas, I., Guinan, E.F., Güdel, M., Audard, M., 2005. Evolution of the solar activity over time and effects on planetary atmospheres. I. High-energy irradiances (1-1700 Å). The Astrophysical Journal 622, 680. Richet, P., Bottinga, Y., Javoy, M., 1977. Review of Hydrogen, Carbon, Nitrogen, Oxygen, Sulfur, and Chlorine Stable Isotope Fractionation among Gaseous Molecules. Annual Review of Earth and Planetary Sciences 5, 65-110. Robert, F., 2003. The D/H ratio in chondrites. Space Science Reviews 106, 87-101. Salvador, A., Massol, H., Davaille, A., Marcq, E., Sarda, P., Chassefière, E., 2017. The relative influence of H2O and CO2 on the primitive surface conditions and evolution of rocky planets. Journal of Geophysical Research: Planets 122, 1458-1486. Schaefer, L., Fegley, B., 2010. Chemistry of atmospheres formed during accretion of the Earth and other terrestrial planets. Icarus 208, 438-448. Shematovich, V.I., Ionov, D.E., Lammer, H., 2014. Heating efficiency in hydrogen- dominated upper atmospheres. Astronomy & Astrophysics 571, A94. Stevenson, D.J., 1990. Fluid Dynamics of Core Formation, Origin of the Earth, pp. 231-249. Tian, F., Toon, O.B., Pavlov, A.A., De Sterck, H., 2005. A Hydrogen-Rich Early Earth Atmosphere. Science 308, 1014-1017. Trail, D., Watson, E.B., Tailby, N.D., 2011. The oxidation state of Hadean magmas and implications for early Earth's atmosphere. Nature 480, 79. van Keken, P.E., Hacker, B.R., Syracuse, E.M., Abers, G.A., 2011. Subduction factory: 4. Depth(cid:1)dependent flux of H2O from subducting slabs worldwide. Journal of Geophysical Research: Solid Earth 116. Wade, J., Wood, B.J., 2005. Core formation and the oxidation state of the Earth. Earth and Planetary Science Letters 236, 78-95. Watson, A.J., Donahue, T.M., Walker, J.C.G., 1981. The dynamics of a rapidly escaping atmosphere: Applications to the evolution of Earth and Venus. Icarus 48, 150-166. Wordsworth, R., 2012. Transient conditions for biogenesis on low-mass exoplanets with escaping hydrogen atmospheres. Icarus 219, 267-273. Wordsworth, R., Pierrehumbert, R., 2013a. Hydrogen-Nitrogen Greenhouse Warming in Earth's Early Atmosphere. Science 339, 64-67. Wordsworth, R.D., Pierrehumbert, R.T., 2013b. Water loss from terrestrial planets with CO2-rich atmospheres. The Astrophysical Journal 778, 154. Wu, J., Desch, S.J., Schaefer, L., Elkins-Tanton, L.T., Pahlevan, K., Buseck, P.R., 2018. Origin of Earth's Water: Chondritic Inheritance Plus Nebular Ingassing and Storage of Hydrogen in the Core. Journal of Geophysical Research: Planets 123, 2691-2712. Zahnle, K., Kasting, J.F., Pollack, J.B., 1990. Mass fractionation of noble gases in diffusion-limited hydrodynamic hydrogen escape. Icarus 84, 502-527. Zahnle, K.J., Kasting, J.F., 1986. Mass fractionation during transonic escape and implications for loss of water from Mars and Venus. Icarus 68, 462-480. Zhang, Y., Xu, Z., Zhu, M., Wang, H., 2007. Silicate melt properties and volcanic eruptions. Reviews of Geophysics 45, 1-27. Supplementary Information -- Calculation of surface temperatures and moist adiabats Planetary surface temperature is controlled by the structure of the troposphere. At the base of the troposphere, vapor pressure equilibrium with the ocean controls the water vapor abundance. We assume a troposphere saturated in water vapor throughout. Accordingly, the partial pressure of water vapor is given by !"#$=exp (−∆-//0), with ∆G=∆H-T∆S and ∆H=40.58 kJ/mol, ∆S=0.1082 kJ/mol.K (Chase et al., 1985). As in any multi-component atmosphere, the inventory of one gas influences the partial pressure of other gases through vertical redistribution (Wordsworth and Pierrehumbert, 2013). The total pressure is assumed given by the expression !2=3"#$gµ/6"#$7"#$ with sH2O the surface density of water vapor, g the gravitational acceleration at the surface, µ the mean molecular weight, and µH2O and xH2O the atomic mass and the mole fraction of water vapor. This expression is strictly valid in a well-mixed atmosphere, which we take as an adequate approximation for an H2O-H2 atmosphere. With the above relations, we write an expression for the entropy of atmospheric gas in contact with the oceans, which we use to calculate surface temperature. The entropy of the troposphere is that of the basal gaseous (A1) which is the expression for entropy of an ideal mixture of ideal gases, with the first term the sum over species as pure gases at standard pressure, the second term an entropy of mixing term deriving from the fact that the gas parcel is a mixture of randomly distributed gas molecules, and the third term a pressure correction due to the volume available to each molecule. In this way, the entropy of the atmospheric parcel at any temperature (T), pressure (P) and composition (xi) can be calculated. We consider ideal gas theory an adequate approximation for the primordial atmosphere because intermolecular distances are large relative to the size of the molecules. The mixture: 8(0)=∑7:;:(T) : −R∑7:ln7: : −RlnP thermodynamic data for the entropy of pure substances is taken from standard sources (Chase et al., 1985). Because a moist adiabat is also isentropic, the tropospheric entropy can be used to relate the conditions at the base to those characterizing the radiative emission level where the mode of energy transport transitions to radiation. A major influence on tropospheric structure is the condensation of water vapor into clouds through adiabatic expansion and cooling. The specific entropy at the radiative emission level at the top of the troposphere must therefore consider condensates: 8(0)=AB8B(T)+AD8D(T) (A2) with FV and FL the fraction of total molecules in the parcel in the gas and condensates, respectively. Atmospheric P-T paths determined by calculating pseudoadiabats (i.e., with rainout) are similar to adiabatic equivalents (i.e., with condensates suspended) (Ingersoll, 1969), permitting use of a two- phase isentrope even when rainout might be expected and upper tropospheric opacity determined by gas-phase (H2) opacity alone. The procedure for calculating surface temperature entails: (1) an initial estimate for surface temperature, yielding the partial pressure of water vapor and, with a given H2 inventory, total surface pressure, (2) calculation of the entropy of the convective atmosphere, (3) evaluation of the thermodynamic state, including temperature, of the atmospheric parcel at a pressure of 0.21 bars representing the radiative emission level. In this way, each value of surface temperature corresponds to an emission temperature, and iteration allows identification of the surface temperature corresponding to the emission temperature required by top-of-the- atmosphere radiative balance (Equation 4). Because the convective troposphere is isentropic, we can solve for the surface temperature as a function of emission temperature without explicitly resolving the vertical structure in the intervening atmosphere. In this way, we calculate surface temperatures in the computationally efficient "all-troposphere" approximation (Pierrehumbert, 2010) and iterate to find solutions. References Chase, M.W., Davies, C.A., Downey, J.R., Frurip, D.J., Mcdonald, R.A., Syverud, A.N., 1985. Janaf Thermochemical Tables - 3rd Edition .2. Journal of Physical and Chemical Reference Data 14, 927-1856. Ingersoll, A.P., 1969. The Runaway Greenhouse: A History of Water on Venus. Journal of the Atmospheric Sciences 26, 1191-1198. Pierrehumbert, R.T., 2010. Principles of planetary climate. Cambridge University Press. Wordsworth, R., Pierrehumbert, R., 2013. Hydrogen-Nitrogen Greenhouse Warming in Earth's Early Atmosphere. Science 339, 64-67.
1212.2848
1
1212
2012-12-12T15:46:24
Impact of stellar companions on precise radial velocities
[ "astro-ph.EP", "astro-ph.IM" ]
Context: With the announced arrival of instruments such as ESPRESSO one can expect that several systematic noise sources on the measurement of precise radial velocity will become the limiting factor instead of photon noise. A stellar companion within the fiber is such a possible noise source. Aims: With this work we aim at characterizing the impact of a stellar companion within the fiber to radial velocity measurements made by fiber-fed spectrographs. We consider the contaminant star either to be part of a binary system whose primary star is the target star, or as a background/foreground star. Methods: To carry out our study, we used HARPS spectra, co-added the target with contaminant spectra, and then compared the resulting radial velocity with that obtained from the original target spectrum. We repeated this procedure and used different tunable knobs to reproduce the previously mentioned scenarios. Results: We find that the impact on the radial velocity calculation is a function of the difference between individual radial velocities, of the difference between target and contaminant magnitude, and also of their spectral types. For the worst-case scenario in which both target and contaminant star are well centered on the fiber, the maximum contamination for a G or K star may be higher than 10 cm/s, on average, if the difference between target and contaminant magnitude is $\Delta m$ < 10, and higher than 1 m/s if $\Delta m$ < 8. If the target star is of spectral type M, $\Delta m$ < 8 produces the same contamination of 10 cm/s, and a contamination may be higher than 1 m/s
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. Artigo1DCstructabstract2 June 20, 2018 c(cid:13) ESO 2018 Impact of stellar companions on precise radial velocities D. Cunha1,2, P. Figueira1, N. C. Santos1,2, C. Lovis3, and G. Bou´e1,4 1 Centro de Astrof´ısica da Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal e-mail: [email protected] 2 Departamento de F´ısica e Astronomia, Faculdade de Ciencias, Universidade do Porto, Rua do Campo Alegre 687, 4169-007 Porto, Portugal 3 Observatoire Astronomique de l'Universit´e de Gen`eve, 51 Ch. des Maillettes, - Sauverny - CH1290, Versoix, Suisse 4 ASD, IMCCE-CNRS UMR8028, Observatoire de Paris, UPMC, 77 avenue Denfert-Rochereau, 75014 Paris, France 2 1 0 2 c e D 2 1 . ] P E h p - o r t s a [ 1 v 8 4 8 2 . 2 1 2 1 : v i X r a Received June 20, 2018; accepted November 29, 2012 ABSTRACT Context. With the announced arrival of instruments such as ESPRESSO one can expect that several systematic noise sources on the measurement of precise radial velocity will become the limiting factor instead of photon noise. A stellar companion within the fiber is such a possible noise source. Aims. With this work we aim at characterizing the impact of a stellar companion within the fiber to radial velocity measurements made by fiber-fed spectrographs. We consider the contaminant star either to be part of a binary system whose primary star is the target star, or as a background/foreground star. Methods. To carry out our study, we used HARPS spectra, co-added the target with contaminant spectra, and then compared the resulting radial velocity with that obtained from the original target spectrum. We repeated this procedure and used different tunable knobs to reproduce the previously mentioned scenarios. Results. We find that the impact on the radial velocity calculation is a function of the difference between individual radial velocities, of the difference between target and contaminant magnitude, and also of their spectral types. For the worst-case scenario in which both target and contaminant star are well centered on the fiber, the maximum contamination for a G or K star may be higher than 10 cm/s, on average, if the difference between target and contaminant magnitude is ∆m < 10, and higher than 1 m/s if ∆m < 8. If the target star is of spectral type M, ∆m < 8 produces the same contamination of 10 cm/s, and a contamination may be higher than 1 m/s if ∆m < 6. Key words. Planets and satellites: detection - Techniques: radial velocities 1. Introduction The search for extrasolar planets is currently a very active field of research in astronomy. Since the first discovery of Mayor & Queloz (1995), more than 800 planets were discov- ered 1, of which 80 % were detected with the radial velocity method (RV). Because it is the workhorse for planetary detec- tions, this method received much attention from the community, which constantly increased its precision (for some of the latest results check Mayor et al. 2011) and tried to characterize its lim- itations. One of the fundamental drawbacks is that it is an indirect method, and therefore one should be extremely careful with false positives, i.e., signals created by other planetary RV signals. The literature provides many examples and techniques used to pin- point these signals (e.g. Queloz et al. 2001; Santos et al. 2002; Melo et al. 2007; Hu´elamo et al. 2008; Figueira et al. 2010) or to average them out (Dumusque et al. 2011b,a). With the ad- vent of very high precision spectrographs, such as ESPRESSO or CODEX (Pepe et al. 2010), the RV precision enters a new do- main, that of cm/s, which in turn will increase the quality of our characterization of these unwanted yet ubiquitous signals (e.g. Cegla et al. 2012). Here we address a so far unexplored mechanism that is capa- ble of distorting RV signals and creating false ones: the contam- ination of the main spectrum by that of a (usually unresolved or 1 http://exoplanet.eu/, as of 27/09/2012 otherwise undetectable) companion. We consider two different cases: 1. that of a faint gravitationally bounded companion and 2. that of an unbound star that is aligned with the target at the moment of observation. We assumed the target stars to be placed at the Galaxy disk edge to consider a representative case that encompasses several sce- narios of contaminant stars. We used real high signal to noise ra- tio (S/N) observations obtained with HARPS to create the com- posite spectra and process these with the RV calculation pipeline to evaluate the impact of the contamination in the most realistic way possible. While we aim at characterizing this effect down to the precision level of ESPRESSO, 10 cm/s, it can be assessed at an even lower precision level. Although this study is performed to characterize the impact on the RV when using the cross-correlation function (CCF), point spread function (PSF) modeling will also be affected by stellar companions, if in a different way. The results of this study are particularly interesting for deep surveys in crowded fields (Kepler). If the contaminant RV remains constant, stellar blends will only cause a undetectable contamination in the RV calculation. However, if the contaminant RV is changing, if the seeing varies, or if there are guiding problems during observations, the effect of the contamination will no longer be constant and it might mimic the presence of a planet. As an example of a false positive caused 1 D. Cunha et al.: Impact of stellar companions on precise radial velocities by blended stars we refer to the work of Santos et al. (2002), in which RV measurements derived from CORALIE blended spec- tra of HD 41004 AB have unveiled a radial-velocity variation with a period of 1.3 days and a small amplitude of 50 m/s, com- patible with the signal expected to be caused by a planetary com- panion to HD 41004 A. Another example is WASP-9b, which was discarded as a planet after it was discovered that the signal was due to a fortuitous alignment (Triaud, priv. comm.). In Sect. 2 we explain which stars we consider most likely to contribute to a shift in the RV calculation, if the star is ei- ther in a G-, K- or M-dwarf binary system, or is affected by a fortuitous alignment. Then, in Sect. 3 we describe the properties of the spectra used in our simulations. In Sect. 4 we describe the method used to scramble spectra from target star and stel- lar companion. Results from our work, assuming the worst-case scenario in which both target and contaminant are well centered on the fiber, are shown in Sect. 5. The results are followed by a statistical analysis in Sect. 5.4, which is based on these results, but assumes a more realistic situation, in which the contaminant star may not be fiber centered. In Sect. 6 we discuss our work, and we present our conclusions in Sect. 7. 2. Contaminant cases When one observes a target star, a secondary star may be present, even without one's knowledge. This secondary star may be a source of contamination for the calculation of the target star's RV. Secondary/contaminants stars may appear as two different types: Real binaries or fortuitous alignments. 2.1. Realbinaries In the case of a real binary system, the contaminant star will be gravitationally bound. In this section we present the probability distributions of real binary properties that we used in our work. Here, for G- and K primaries we based our study on the re- sults of Duquennoy & Mayor (1991), who studied the multiplic- ity among solar-type stars in the solar neighborhood. They con- sidered a subsample of 164 primary stars with spectral types be- tween F7 and G9. Using Table 7 of Duquennoy & Mayor (1991) on the mass-ratio distribution among G-dwarfs binaries, we can calculate the probability of observing a binary with a mass-ratio q = M2/M1, where M2 is the mass of the secondary star and M1 the mass of the primary. These probabilities are shown in Table 1, in which we can see that there is a maximum for the mass- ratio range ]0.2,0.3]. The distribution of orbital periods (P) of their study is approximated to a Gaussian with log P = 4.8 and σlogP = 2.3. Duquennoy & Mayor also presented the eccentricity distri- bution for P < 1000 and P > 1000 days, shown here in Table 2. We stress that there is a large uncertainty for VLMC (very low mass companion) binaries (with q < 0.1), and that this study does not include values of q below 0.01. We also emphasize that this mass ratio distribution is calculated for binary stars, but we recall that, also according to Duquennoy & Mayor (1991), there are 57% of binary systems with a mass ratio higher than 0.1, and 43% of apparently single stars. Of these 43%, (8 ± 6)% most probably have a VLMC, which puts the percentage of single stars at ∼ 30%. For M-dwarfs we used the distributions of mass ratio, q, and semi-major axis, a, from the work of Janson et al. (2012) on M- dwarf multiplicity. The authors suggest an M-dwarf total multi- 2 Table 1: Probability of observing a binary with a mass ratio q = M2/M1 among G-dwarf binaries for periods P < 104 days and P < 104 (from Duquennoy & Mayor 1991). qmax 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 P < 104days P(q) [%] 12.0 16.3 13.1 12.7 12.3 12.9 6.5 1.9 2.6 3.4 6.2 P(q) [%] 11.1 11.9 18.4 15.8 10.5 11.2 6.6 6.6 4.0 4.0 0 P > 104days P(q) [%] 11.4 13.6 16.4 14.6 11.2 11.8 6.5 4.8 3.4 3.8 2.4 Total Table 2: Probability of observing a binary with an eccentricity e among G-dwarfs binaries for periods P < 103 days and P < 103 (from Duquennoy & Mayor 1991). emax 0.15 0.3 0.45 0.6 0.75 0.9 1 P < 103days P(e) [%] 12.50 43.75 31.25 6.25 6.25 0.0 0.0 P(e) [%] 5.88 11.76 23.53 14.71 20.59 23.53 0.0 P > 103days Table 3: Probability of observing an M-dwarf binary with a mass ratio q = M2/M1 (from Janson et al. 2012). qmax 1.0 P(q) [%] 0.0 1.61 5.95 10.45 12.93 13.84 14.48 13.03 14.58 13.13 0.1 0.2 0.3 0.6 0.7 0.4 0.5 0.8 0.9 Table 4: Probability of observing an M-dwarf binary with a cer- tain log a (from Janson et al. 2012). logamax 2.25 2.5 P(loga) [%] 0.51 2.60 9.02 15.39 17.13 17.65 14.73 11.74 7.85 3.37 0.25 0.5 0.75 1.25 1.75 1.5 2. 1. plicity fraction of 34% . Table 3 shows the probability of observ- ing an M-dwarf binary with a mass ratio q. Janson et al. found a uniform distribution to be more consistent with the mass ratio of their observed sample, than a rising distribution. Accordingly, we assumed their uniform distribution in our work. For the semi- major axis distribution Janson et al. (2012) also compared the Sun-like and the narrow distribution, finding the latter to be more consistent with their sample. We therefore used this distribution, which is shown in Table 4. As no information was given on the eccentricity distribution for M-dwarf binaries, we assumed the distribution from Duquennoy & Mayor (1991). These distributions were used in our statistical analysis (Sect. 5.4). 2.2. Fortuitousalignment Even if the target star is a single star, an alignment with a back- ground/foreground object may occur. To study the probability of such an event we used the Besanc¸on model (Robin et al. 2003). This allows one to compute the probable stellar con- tent on a given direction of the Galaxy, permitting one to in- fer the most probable contaminants. To do so, we simulated an observation in the direction of the HD 85512 coordinates, D. Cunha et al.: Impact of stellar companions on precise radial velocities Table 5: Most probable contaminants for the fortuitous align- ments. Spectral type [F0, F5[ [F5, G0[ [G0, G5[ [G5, K0[ [K0, K5[ [K5, M0[ [M0, M5[ [M5, M8[ [M8, M9] [13, 16] ∧ [20, 23] mV [13, 24] [13, 25] [13, 25] [12, 26] [16, 30] [17, 30] [20, 30] [21, 30] P [%] 0.08 1.25 1.74 1.32 2.5 4.48 16.99 12.12 13.17 Table 6: Stars used in our simulations, their spectral type, visual magnitude (mv) and S/N at the center of the spectral order 50 (SN50). Object HD20852 HD103774 Sun Tau Ceti HD69830 Alpha Cen B HD85512 Gl436 Gl581 Spec.Type F2 F5 G2 G8 K0 K1 K5 M1 M3 (1) mv 6.92 7.12 3.65 (2) 3.50 5.95 1.33 7.65 10.68 10.57 SN50 225; 98 102; 146 225; 260 300; 450 253; 380 385; 399; 302 156; 86 15; 31; 46 37; 22 (1) from Simbad (http:simbad.u-strasbg.fr/simbad/sim-fid) (2) mv of the moon within the HARPS 0.5" radius fiber. (l, b) = (271.6759, 8.1599), which is positioned at the border of the Galactic disk. We also considered a distance interval of 50 kpc (approximately the diameter of the Milky Way), magnitudes between 0 and 30, and a solid angle for a radius of 3◦. The objec- tive was to have a large number of stars for each bin of magni- tude and spectral type and then to divide it by the ratio between the area considered and that for the HARPS fiber (radius of 0.5 arcsec), thus obtaining the density of stars or the probability to have a fortuitous alignment within the HARPS fiber. In Fig. 1 we present a contour plot of the total number of stars obtained in the simulation for each bin of magnitude and spectral type. For more details on these values and the values for the star density when normalized for a HARPS size fiber, see Appendix A. The most probable contaminants for the accidental alignments are shown in Table 5. 3. Data We contemplated target stars from spectral type G to late M, with visual magnitudes, mv, between 5 and 15. As described in the previous section (Sect. 2.2), the most probable contaminant stars are of spectral types FGKM. To represent both target and contaminant spectral types, we chose spectra from nine stars, which are shown in Table 6. We used spectra with an S/N at the center of the spectral order number 50 (which corresponds to a wavelength of 437.28 nm), varying between 15 for an M star, and 450 for a G star. The S/N values of these spectra can be found in Table 6. The stars were chosen because they had the highest S/N in their spectral type range, as observed with HARPS. 4. Method The goal of this work is to study possible stellar contamination of a target star. To do so, we co-added target and contaminant spectra. We not only considered contaminant stars of spectral types F2, F5, G2, G8, K0/K1, K5, M1, and M3, but also different ratios of magnitude between target star and contaminant as well as differences in RV. To sum two spectra with these particular conditions is not trivial. We need to control the magnitudes and RVs. We set the magnitude V of the target (mt) star according to the data available in Simbad 2 and then calculated the magnitude of the contaminant (mc) with respect to the target star using the apparent magnitude - flux relation. We also considered the rel- ative RV between the two spectra as a tunable knob. However, when processing our spectra using the HARPS pipeline the re- duction takes into account the header (and properties) of the par- ent spectra. To overcome this situation we calculated the contam- inant RV in the target reference frame and shifted it afterwards. We then co-added the target spectrum with the modified con- taminant spectra and processed the result with the HARPS RV pipeline. More details of the method are presented in Appendix B. 4.1. The"realbinaries"starstrategy To study the impact of a companion on the RV calculation of a binary primary star, we reproduced the cases discussed in the work of Duquennoy & Mayor (1991). We considered a primary G2 star with six different secondary/contaminant stars: G2, G8, K1, K5, M1, and M3. The first case had a G2 as contaminant with positive and neg- ative values of ∆RV = RVC′c − RVCt, where RVCt is the target barycentric radial velocity (drift corrected) without contamina- tion, and RVC′c is the contaminant barycentric radial velocity (drift corrected) in the target reference frame. In Fig. 2 we show the impact, RVCt − RVC′t , of the stellar companion on the target RV, where RVC′t is the target barycentric radial velocity (drift corrected) with contamination. For each target-contaminant pair we considered a range of ∆RV between 0 and 30[Km/s] (in steps of 1 [Km/s]), except for a secondary star of spectral type K5 in which the maximum ∆RV was of 24 [Km/s] (see Appendix B for details). 4.2. Fortuitousalignmentstrategy The fortuitous alignment case is more generic. Here we consid- ered six spectral types of primary stars (G2, G8, K0, K5, M1, and M3) in combination with the eight spectral types as contam- inants -- the same six used as target stars plus an F2 (HD20852) and an F5 (HD103774) star. For each spectral type of primary stars we considered contaminants with magnitudes mv between 5 and 15 (in steps of one). The contaminants considered were those described in Table 5. Although the Besanc¸on model includes oxygen and carbon AGB (asymptotic giant branch) stars and white dwarfs in the search, we only considered main-sequence (MS) stars. Very metal-poor stars were also excluded from our study. For an M1 target star a K0 star (HD69680) was used instead of the K1 men- tioned above. This was because the barycentric Earth radial ve- locity (BERV) value form the K1 star did not allow the desired shift in RV. For each target-contaminant pair we considered the 2 http://simbad.u-strasbg.fr/simbad/sim-fid 3 D. Cunha et al.: Impact of stellar companions on precise radial velocities Fig. 1: Number of stars in each bin of spectral type and magnitude (radius of 3◦). 5. Results 5.1. Realbinaries The results for the contamination by the secondary star of a bi- nary with a primary star of spectral type G2 as a function of ∆RV can be found in Fig. 3. Obviously, the impact on the RV calcula- tion is a function of the difference between target and contami- nant RV (∆RV). One can interpret a CCF as an "average" line of a spectrum, that is fitted by a Gaussian curve to measure the cen- ter of the CCF. In the presence of a second star, the CCF is com- posed of two curves. For ∆RV = 0 the two curves fully overlap, and so the impact on the Gaussian fit and on the RV calculation is minimum. As ∆RV increases, the impact on the Gaussian fit increases until the point when the two CCF curves overlap, with the contaminant RV, RVC′c, close to 0.6 times the FWHM of the target CCF. When the CCF curves become distinct, the impact on the measured RV of the primary decreases. For ∆RV > 20 km/s the peak of the contaminant CCF curve exits the CCF window that is usually computed with a value of 20 km/s, and for values ∆RV ≥ 25 km/s the contaminant CCF curve completely exits the target CCF window. Nevertheless, the contamination is not null. There are bumps in the wings that slightly contaminate the target spectra. Figure 4 shows the maximum contamination for each type of contaminant. As we can see, as the secondary stars become fainter and of later spectral type, the maximum contamination decreases. Fig. 2: Shift of a G2 target RV caused by a contaminant of the same spectral type and magnitude, mv, in function of the differ- ence between individual radial velocities. differences between its RV, ∆RV, under 20 Km/s, which is half the width of the CCF window. Although giant stars will often be contaminant stars, we assumed that spectra from MS stars are similar to those of giants and used only spectra of MS stars in our simulations. 5.2. Contaminationbythesky The sky brightness is not zero and its light results from the re- flection and scattering of Sun spectra. The sky brightness has a 4 D. Cunha et al.: Impact of stellar companions on precise radial velocities Fig. 3: Impact on the RV of a G2 target star caused by a companion of spectral type G2, G8, K1, K5, M1, and M3 (left to right, top to bottom). For each case a close-up with a bump for the impact for 10 < ∆RV < 24 km/s is shown. magnitude per arcsec2 in V band of 21.8 on a new-moon night, and of 20.0 on a full-moon night. To study the impact of this con- tamination we assumed that sky spectrum is the solar one. This is equivalent to say that for a fiber of 0.5" radius, we have a sys- tematic contamination of a G2 star with mV ∼ 20 on a full-moon night to a limit of a G2 star with mV ∼ 22 (new moon). The results of the maximum contamination by the sky bright- ness study are shown in Fig. 5. We can see that when observing a G2, G8, K0, or a K5 target star fainter that mV = 11, or M1, M3 fainter that mV = 14, the errors induced on the observed RV can be larger than 10 cm/s, which can be a problem for future instruments such as ESPRESSO. For a night of observation with a full moon, induced errors become larger than 10 cm/s for G- and K target stars fainter that mV = 9 and for an M star fainter than mV = 12. 5.3. Fortuitousalignment In this section we present a more exhaustive study of all pos- sible sources of stellar contamination. Beyond the moon as a contaminator, there is a vast range of possible contaminators, as was discussed in Sect. 2.2. Our study is simplified because our analysis holds for similar spectra with different magnitude for target and contaminant star, but with same difference between them (∆m = mc − mt), as we can see in Fig. 6. This way, we only carried out our study on the impact for a target star in which mv differs ∆m from the contaminant star. The result of our study for contamination as a function of spectral type is shown in Fig. 7. We first note that there is a clearly defined relation between ∆m and the maximum of the contamination, except for some cases where the difference in magnitude between target and con- taminant is null. A misidentification of the companion correla- tion peak with that of the primary may happen due to a (slightly) larger contrast. If we take only ∆m > 1, we can describe the relation between ∆m and the maximum of the contamination as MaxRVCt − RVC′t = 10a·∆m+b, where a and b are the coefficients presented in Table 7. In this table we also present the value for the root mean square (rms) for each fit. By comparing the panels of Fig. 7 we can also see that the contamination depends not only on ∆m, but also on the target-contaminant spectral type combination. This be- comes more evident in Fig. 8. This figure shows the maximum contamination as a function of the contaminant spectral type for a G2 and M3 target star with ∆m = 10. We find two reasons for this behavior. First, less contamination occurs with increas- ingly different spectra. Also, the contamination will depend on (1) 5 D. Cunha et al.: Impact of stellar companions on precise radial velocities (a) (c) (e) (b) (d) (f) Fig. 5: Maximum contamination of the sky brightness on a a) G2, b) G8, c) K0, d) K5, e) M1, and f) M3 target star spectra. The blue triangles show the impact on the RV caused by the sky brightness on a full-moon night, while black diamonds show the impact on a new-moon night. the depth of spectral line. Thus, for similar spectra, cooler stars will cause a higher contamination, because their spectral lines depth is higher. 5.4. Statisticalanalysisofthecontamination After calculating the impact caused by a contaminant star in the RV calculation, it is important to known how often this occurs. Therefore we calculated the probability of measuring the target RV with a given contamination. In the previous two subsections (Sects. 5.1 and 5.3) we pre- sented a study for the impact to have a contaminant star within 6 D. Cunha et al.: Impact of stellar companions on precise radial velocities Table 7: Coefficients a an b for each fitting of the relation be- tween ∆m and the maximum impact on the RV calculations: MaxRVCt − RVC′t = 10a·∆m+b, and rms resulting from each combination of target spctral type (TST) and contaminant spec- tral type (CST). TST CST a b G2 G8 K0 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 F2 F5 G2 G8 K1 K5 M1 M3 -0.4008 -0.4019 -0.4022 -0.4023 -0.4026 -0.4009 -0.4001 -0.3996 -0.4008 -0.4019 -0.4027 -0.4021 -0.4027 -0.4007 -0.3995 -0.3984 -0.3812 -0.4008 -0.3933 -0.3929 -0.3954 -0.3914 -0.3799 -0.3777 -0.3988 -0.3999 -0.4035 -0.4032 -0.4053 -0.4030 -0.4011 -0.4009 -0.4056 -0.3932 -0.4018 -0.4011 -0.3990 -0.4014 -0.3985 -0.3994 -0.4052 -0.3924 -0.4035 -0.4019 -0.4024 -0.4021 -0.3991 -0.3999 5.0733 5.1601 5.4288 5.4829 5.4556 5.1859 4.8096 4.7301 4.9379 5.0222 5.3440 5.3154 5.3641 5.11 4.7357 4.651 4.7320 4.9521 5.2712 5.2613 5.3704 5.1827 4.7540 4.6783 4.9559 5.0865 5.4592 5.5354 5.5580 5.3793 4.9458 4.9257 3.5177 3.4673 4.1212 4.2786 4.1604 4.5525 5.0993 5.1581 3.4770 3.4421 4.0854 4.2469 4.2649 4.5317 5.0440 5.1436 rms [cm/s] 815.0813 1644.9045 4491.1375 4956.4768 5136.0605 1500.0546 23.3793 197.1604 681.0614 1075.0923 3968.9098 2697.4956 4215.8544 1139.4446 3.0972 51.5545 1566.6309 871.3506 5307.563 4246.1257 7390.8788 3347.1766 566.5518 532.1239 200.036 724.4446 7366.4185 8587.4134 14146.5377 4935.42 71.0313 98.6394 18.1497 24.6295 10.1997 13.0058 24.4166 19.9346 33.3093 56.2493 14.7829 25.7393 30.8372 23.5459 35.7392 32.2298 13.2551 63.9946 either, and also just considered that only ∆RV ≤ 20Km/s will produce a significant impact. We ran the simulations for each case of spectral type target considered in previous sections, at a distance d = 1, 5, 10, 50, and 100 [pc] from Earth. Before presenting our results of the statistical analysis we describe in more detail the considerations to statistically analyze the real binaries (Sect. 5.4.1) and fortuitous alignments (Sect. 5.4.2). 7 Fig. 4: Maximum contamination of a G2 target star caused by each companion spectral type in m/s. Fig. 6: Contour plot of the maximum induced RV shift in the tar- get RVCt − RVC′t depending on the target (mt) and contaminant (mc) magnitudes. the fiber on the RV calculations, assuming that the fiber was well centered on target and contaminant. Although this is a rel- evant study, the most probable scenario for real observations is that of a contaminant star that not centered in the fiber. In this case, we cannot consider that both target and contaminant fluxes will completely enter the fiber, and therefore we calculated how much of each flux enters the fiber. The fraction of binary and for- tuitous alignments and their properties also follows established probability laws. To take these parameters into account in a real- istic way, we ran Monte Carlo simulations with 10000 trials, to calculate the distribution function of the impact of the presence of a contaminant star on the RV calculation. Once again, we had to consider both cases of contamination: binaries and fortuitous alignment. If in our simulations one star was in a binary sys- tem and had a background/foreground at the same time, we only considered the stronger RV effect. In these simulations we did not take into account the contamination by the sky brightness D. Cunha et al.: Impact of stellar companions on precise radial velocities is present, the probability that the star is standing at a distance r from the target is proportional to r2, and the probability to have the fortuitous contaminant star between r and r + dr is given by equation 3: P(r, r + dr) = π(2rdr + dr2), (3) in which the members are self-explanatory. To calculate ∆RV we used the probability density function for the systemic RV, which was computed by fitting a Gaussian to the systemic velocity of stars from HARPS. This yielded an average and a dispersion of (µ, σ) = (7.87, 29.87). 5.4.3. Statistical analysis results The probability to have a given impact on the RV calculation for each case is presented in Fig. 9. We can see that the most proba- ble scenario is to have a contamination of less than 10 cm/s. But contamination may happen and is more probable for far-away target stars, either because they become fainter, or because, in the case of a binary star, it is more probable that the compan- ion star flux enters the fiber. We can examine the target star G8, for instance. In this case, if the target star is 100 pc away, there is only a probability of 67% for a contamination below 10 cm/s, and 14% probability for a contamination between 10 and 100 m/s. There is a peak around a contamination of 103 cm/s, which is, very probably, caused by physical binary contamina- tion. This level of contamination needs to be taken into account when choosing targets for planet searches. This becomes even more important when choosing target stars to be observed with ESPRESSO. 6. Discussion We showed that we should not neglect the possibility that a stel- lar companion contaminates the RV calculations. This is spe- cially true for far-away stars, which have a higher probability to have two stars within the fiber. Nevertheless, we stress some assumptions we made that may have some impact on our work. Because spectra with high SN are scarce for late-M stars, in our statistical analysis we used the M1 spectra to represent the [M0, M5[ bin and the M3 spectra for the [M5, M9]. We also assumed that the star will have at most one stellar companion, because we found that the probability to have two contaminant stars is too low. We also did not take into account the effect of the sky brightness if there was no stellar companion. We also stress that the Besanc¸on model is a probabilistic model, and thus is not free of errors. We consider it reasonable that the distribution of fortuitous alignments is a Poisson distri- bution of stars. Thus, the associated error is equal to √N, where N is the number of stars per area element. The results presented in Fig. 7 may lead one to think that contamination may be stronger than it actually is, but we should not forget that this figure only shows the maximum impact of contamination as a function of ∆m, which occurs when RV(contaminant) ∼ 0.6 × FWHM(target). We also consid- ered that both stars are well-centered in the fiber and that all flux enters the fiber. This justifies our statistical analysis for the impact of the stellar companion within the fiber. With our statis- tical analysis we are able to conclude that it is important to take into account the possibility of contamination by a stellar com- panion within the fiber, especially when the RV precision enters the domain of the cm/s. Although the probability to have a con- tamination lower than 10 cm/s is above 50% for every target star Fig. 8: Maximum contamination RVCt − RVC′t as a function of the contaminant spectral type for a G2 (blue dots) and M3 (green diamonds) target star with ∆m = 10. 5.4.1. Statistical analysis of binaries For our statistical analysis of contamination in binary systems, we considered the mass ratio distributions, q, mentioned in Sect. 2.1. We used Table 1 based on the study of Duquennoy & Mayor (1991) for target stars of spectral types G and K, and Table 3 based on the study of Janson et al. (2012) for target stars of spec- tral type M. Because we had M1 and M3 spectra for M stars, we used the M1 spectra for [M0, M4[ stars, and for the later types we used the M3 spectrum. From the work of Duquennoy & Mayor (1991) and Janson et al. (2012) we also used the probability density function for the orbital period and for the eccentricities (see Table 2) for G dwarfs, and the probability density function for the semi-major axis for M-dwarfs (see Table 4). The pro- jected distance R and the ∆RV were drawn performing a Monte Carlo simulation of Kepler's laws and its projection, consider- ing a uniformly random mean anomaly (M ∈ [0, 2π]), inclina- tion (cos i ∈ [−1, 1]), longitude of the node (Ω ∈ [0, 2π]), and argument of the periapsis (ω ∈ [0, 2π]). To compute the luminosity of the stellar companion given the mass of the star, we used relation 2 from Duric (2003): if M > 0.43M ⊙ , if M < 0.43M ⊙. ; (2) M ⊙(cid:17)4.0 ⊙ ∝ (cid:16) M ⊙ ∝ 0.23(cid:16) M M ⊙(cid:17)2.3 L L L L  5.4.2. Statistical analysis of fortuitous alignments For the fortuitous alignment we are only interested in contami- nant stars whose flux can enter the fiber. With a seeing of 0.93, which is the mean value for the La Silla observatory, we mea- sured that the fraction of flux that enters the fiber of a star at 2.6 arcsec from the fiber center is 2×10−8. If target and contaminant star have the same magnitude, this corresponds to an effective ∆m of 19 within the fiber, and so its impact on the target RV can be considered null for our purpose. We then created a probability density function to have a contaminant star of a certain spectral type and magnitude within a radius of up to 2.6 arcsec around the target star, using the density tables from Appendix A. For background/foreground stars within the 2.6 arcsec ra- dius we assumed a uniform distribution, and if a contaminant 8 D. Cunha et al.: Impact of stellar companions on precise radial velocities considered, we should pay special attention to far-away target stars, for which the impact of contamination can be higher than 10cm/s for 30 − 40% of the cases. Table 8: Limit ∆m for which the maximum impact (Imax) on the target RV is higher than 0.1, 1, 10, and 100 m/s, for each combi- nation of target spctral type (TST) and contaminant spectral type (CST). 6.1. Impactontypicalcases As a consequence of the results presented in Fig.7, we present in Table 8 a study on the limit ∆m for which the maximum im- pact on the RV calculation is higher than 10 cm/s, 1, 10, and 100 m/s. From this table we can see that for a G star with a contaminant M star, an impact higher than 100 m/s only occurs if ∆m < 2. However, in Table 5 from Sect. 2.2 we have seen that the probability to have a contaminant M star of mv < 17 is very low. Thus, if we are only interested in target stars brighter than mv = 15, an impact higher than 100 m/s is very unlikely. We can also see that when observing an M star, an F star can never cause an impact higher than 100 m/s, and an impact of tens m/s is only possible if target and contaminant have almost the same magnitude (∆m . 1.5). On the other hand, even for actual precisions, we may reach a potentially detectable contam- ination due to a ∆m = 9. In many cases the induced effect is not detectable because it is a constant effect. Nevertheless, there are conditions (variable seeing, centering and/or pointing problems, etc.) in which the induced effect will not be constant. Even so, a very particular configuration is necessary (and is highly un- likely) for the effect to create a periodic signal. A contamination stronger than 1 m/s on the RV calculation of a K0 target star is possible if ∆m < 9. If the K0 star is at 10 pc, it will have a mv ∼ 6, which means that contaminant stars with 6 ≤v≤ 15 can cause such a contamination. The probability to have one con- taminant star with 6 ≤ mv ≤ 15 within the 0.5" radius fiber is 0.02% only. These tabled effects correspond to the maximum case, i.e., to the difference between contamination and no contamination. Discussed cases are extreme in the sense that they are contam- ination and no contamination. Nevertheless, to extensively ana- lyze a variable contamination impact we would need to perform many hypotheses with uncertain parameters, such as seeing or flux counts. 6.1.1. Sky brightness In our analysis of the sky brightness impact we assumed that sky brightness is mainly caused by the moon, and so its spectra will be that of a star with spectral type G2. The new-moon sky bright- ness should then be considered as a lower limit for the contami- nation from the sky brightness with a G2 spectrum. However, we should not forget that sky brightness has other sources and may depend on the sky zone. Sky spectra from the Galactic disk are surely different from the Galactic halo. Moreover, there is also the possibility to have a galaxy as contaminant, which was not considered. Nevertheless, we consider it a good approximation, or at least a good starting point, to treat the sky brightness as a G2 star. If observations are carried out on a night with moon, and if cirrus are present, the contamination by the moon will be en- hanced. Tripathi et al. (2010), reported two outlier RV data points caused by cirrus during their observations of WASP-3, an F7 star of mv = 10.7, on a full-moon night. Their RV measurements present a redshift of 140 m/s and of 49 m/s with respect to their best-fitting models. If we assume that cirrus will reflect some of the moonlight, we have a G2 contaminant. The maximum value G8 K0 K5 M1 CST 5 5 2 11 9 6 4 11 6 11 8 5.5 3 10 8 5.5 3 9.5 7 4.5 2 10.5 9.5 7 4.5 2 10.5 9.5 7.5 4.5 2 10 7.5 5 2.5 TST Imax[m/s] F2 F5 G2 G8 K0 K5 M1 M3 11 11 10 9.5 G2 10 9 9 7 8 8 6 6 4.5 5.5 5.5 4 2 4 3 3 11 9.5 10 11 10 8.5 7.5 7.5 8.5 8.5 7 6 4.5 5 6 3.5 2 2.5 2.5 3.5 3.5 11 9.5 9.5 9.5 11 8 8.5 7.5 7.5 7.5 8.5 8.5 5.5 6 6 4.5 6 4.5 3 2 2.5 3.5 3.5 3.5 11 10 10 10.5 11 11.5 11.5 8.5 7.5 8.5 7.5 6 5 5 6 3.5 2.5 2.5 3.5 2.5 8.5 10.5 10.5 6.5 6.5 7.5 6.5 5.5 4 4 4 1.5 1.5 3 1.5 9 6.5 3.5 1.5 0.1 1 10 100 0.1 1 10 100 0.1 1 10 100 0.1 9 1 6.5 10 4 100 8 0.1 6 1 10 3 100 -- -- 0.5 0.5 6.5 6.5 7.5 0.1 8 5.5 5 3.5 3.5 1 3 10 1 1 2.5 100 -- -- 0 0.5 8 8 5.5 5.5 2.5 2.5 10.5 10.5 7.5 7.5 5 5 2.5 2.5 9 6.5 4 8 5.5 3 0.5 8 5.5 3 0.5 8 5.5 M3 considered for all practical effects for the visual magnitude of this contaminant will be that of the target star (if it were higher than the target magnitude, the target star would be no longer visible). Thus, for a G2 target star with mv = 10.7 and a contam- inant with the same spectral type and magnitude, the maximum contamination will be 2.7 Km/s. This corresponds to 0.15% of the moonlight flux, considering that 100% of the moonlight flux corresponds to a star of magnitude 3.65. If 0.01% of the moon- light reaches the fiber, corresponding to a contaminant of spec- tral type G2 with mv = 13.7, the contamination would be 140 m/s. For a contamination of 49 m/s, 0.003% of the moonlight flux would be enough, corresponding to a G2 contaminant with mv = 14.8. Because the Tripathi et al. (2010) data were taken with the High Resolution Echelle Spectrometer (HIRES) on the Keck I telescope, and we used HARPS spectra in our simula- tion, and also because we did not calculate the impact for a tar- get of spectral type F, these contamination values are presented here just as a reference, since we cannot directly compare them. Nevertheless, we consider that their hypothesis of cirrus contam- ination is valid. The sky brightness contamination is not easily removed. Although its signal may be strong enough to contaminate the RV calculations, it will be too weak to be adequately characterized, fitted, and removed, even for a long-slit spectrograph. 6.1.2. Kepler stars Our work can also be a valuable asset for the follow-up of pro- grams such as Kepler. Kepler stars have 11 . mv . 18 (see Batalha et al. 2012), and consequently, measuring the RV be- comes challenging. Beyond the problem of these being faint stars, there is also the problem of the false-positive planet transit. As an example, we consider the case of Kepler-14b 9 D. Cunha et al.: Impact of stellar companions on precise radial velocities Queloz, D., Henry, G. W., Sivan, J. P., et al. 2001, A&A, 379, 279 Robin, A. C., Reyl´e, C., Derri`ere, S., & Picaud, S. 2003, A&A, 409, 523 Santos, N. C., Mayor, M., Naef, D., et al. 2002, A&A, 392, 215 Tripathi, A., Winn, J. N., Johnson, J. A., et al. 2010, ApJ, 715, 421 (Buchhave et al. 2011), a planet orbiting an F star with a com- panion star of nearly the same magnitude and only 0.3 arcesec of sky-projected angular separation. Buchhave and collaborators were not able to detect a second peak on the CCF and concluded that the two stars probably have nearly the same RV. They com- puted an impact of 280 m/s on the RV caused by the companion star. A direct comparison with their work is not possible, because their observations were carried out with the Fiber-fed ´Echelle Spectrograph (FIES) at the 2.5m NOT at La Palma, and with the High Resolution Echelle Spectrometer (HIRES) mounted on the Keck I on Mauna Kea, Hawaii. Moreover, our simulation only contemplates target stars of spectral GKM, and this is an F star. Still, performing a rough comparison, we see that our simula- tions indicate an impact of ∼ 276 m/s for a G2 target star with a contaminant of the same spectral type, ∆m = 1, and ∆RV = 1. 7. Conclusions This study on the impact of stellar companions within the fiber of RV calculations allows us to conclude that we should not ne- glect the possibility of contaminant flux from stellar companions stars, specially if we are observing far-away stars. On average, if we are observing a G or a K star, the contamination may be higher than 10 cm/s if the difference between target and contam- inant magnitude is ∆m < 10, and higher than 1 m/s if ∆m < 8. If the target star is a star of spectral type M, a ∆m < 8 is enough to obtain the same contamination of 10 cm/s, and a contami- nation may be higher than 1 m/s if ∆m < 6. We also showed that sky brightness should not be discarded, particularly on full- moon nights, if one observes faint target stars (mv > 10). These results will allow us to more wisely choose tar- get stars to be observed with instruments such as ESPRESSO and CODEX, and they provide reference values for the differ- ent cases of contamination possible in fiber-fed high-resolution spectrographs. We also stress that there are diagnosis methods that should be capable of detecting most of the blends that mimick planets. We can detect a blend through bisector anal- ysis of the CCF or from correlations using different templates (Santos et al. 2002). Acknowledgements. This research has made use of the SIMBAD database, op- erated at CDS, Strasbourg, France. DC, PF, NCS, and GB acknowledge the sup- port from the European Research Council/European Community under the FP7 through Starting Grant agreement number 239953, as well as from Fundac¸ ao para a Ciencia e a Tecnologia (FCT), Portugal, in the form of grant reference PTDC/CTE-AST/098528/2008. We also thank J. S. Amaral for a critical reading of the manuscipt. References Batalha, N. M., Rowe, J. F., Bryson, S. T., et al. 2012, ArXiv e-prints Buchhave, L. A., Latham, D. W., Carter, J. A., et al. 2011, ApJS, 197, 3 Cegla, H. M., Watson, C. A., Marsh, T. R., et al. 2012, MNRAS, 421, L54 Dumusque, X., Santos, N. C., Udry, S., Lovis, C., & Bonfils, X. 2011a, A&A, 527, A82 Dumusque, X., Udry, S., Lovis, C., Santos, N. C., & Monteiro, M. J. P. F. G. 2011b, A&A, 525, A140 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Duric, N. 2003, Advanced Astrophysics Figueira, P., Marmier, M., Bonfils, X., et al. 2010, A&A, 513, L8 Hu´elamo, N., Figueira, P., Bonfils, X., et al. 2008, A&A, 489, L9 Janson, M., Hormuth, F., Bergfors, C., et al. 2012, ApJ, 754, 44 Mayor, M., Marmier, M., Lovis, C., et al. 2011, ArXiv e-prints Mayor, M. & Queloz, D. 1995, Nature, 378, 355 Melo, C., Santos, N. C., Gieren, W., et al. 2007, A&A, 467, 721 Pepe, F. A., Cristiani, S., Rebolo Lopez, R., et al. 2010, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 7735, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series 10 D. Cunha et al.: Impact of stellar companions on precise radial velocities (a) (c) (e) (b) (d) (f) Fig. 7: Maximum impact of ∆m on the RVCt for a target star of spectral type a) G2, b) G8, c) K0, d) K5, e) M1, and f) M3, and contaminants of spectral types F2 (star), F5 (hexagon), G2 (squares), G8 (diamonds), K1 (circles), K5 (triangles), M1 ( x crosses), and M3 (+ crosses). 11 D. Cunha et al.: Impact of stellar companions on precise radial velocities (a) (c) (e) (b) (d) (f) Fig. 9: Close-up of the distribution of the expected contamination , RVCt − RVC′t, on the radial velocity, RVCt, of a target star of spectral type a) G2, b) G8, c) K0, d) K5, e) M1, and f) M3, and at distance d, caused by a background or a companion star. The green thick solid line shows the distribution if the target star stands at a distance of d = 1 pc, the blue dotted line for d = 5 pc, black dashed line for d = 10 pc, magenta dash-dotted line for d = 50 pc, and red solid line for 100 pc. Numbers in each bin correspond to the impact probability for each target distance: the number in the top corresponds to 1 pc and the number in the bottom to 100 pc. 12 Appendix A: Density of stars D. Cunha et al.: Impact of stellar companions on precise radial velocities Fig. A.1: Number of stars in each bin of spectral type and magnitude for an area of the sky with a radius of 3◦. Fig. A.2: Density of stars in each bin of spectral type and magnitude for an area of the sky with a radius of 0.5 arcsec. 13 D. Cunha et al.: Impact of stellar companions on precise radial velocities Appendix B: Method details To control the magnitude of the contaminant star we first used the relation B.1 to calculate mv in the target reference frame, mt − mc = −2.5 × log(Ft/Fc) , where Ft and Fc are the fluxes of the target and of the contaminant. The magnitudes are now calibrated and can be changed by multiplying the flux by a factor of (B.1) f = 10−x/2.5, (B.2) where x is the difference between the new chosen magnitude and the true magnitude of the star in the target-star system. To change the RV of the contaminant spectra we also had to modify them. The first challenge when processing our co-added spectra with the HARPS RV pipeline is that it will take into account the header (and proprieties) of the parent spectra. To compensate for this we first need to replace the header of the contaminant spectrum file with that from the target, and then run the cross correlation function (CCF) recipe of the RV pipeline. The next step in this process was to shift the contaminant spectra. To assess by how much the contaminant spectra had to be shifted, we used equation B.3 λc = λt × 1 + RVCc − RVCt − ∆RV c !! , (B.3) where λc are the new wavelengths to derive the desired ∆RV (difference between the RV of the target and the contaminant), λt are the wavelengths from the wave file corresponding to the target, and c is the light speed in vacuum. These new wavelengths were used to calculate, by linear interpolation, the number of pixels by which the contaminant needed to be shifted. This gives a new xx axis that it was used, recurring to another linear interpolation, to shift the contaminant spectra. With the target RV for the contaminant, the next step is the sum of the target with the modified contaminant spectra, which is the sum of their fluxes. The shift of the contaminant spectra should not be larger that the maximum barycentric Earth radial velocity (BERVMX) of the star, otherwise there is the risk that the spectral lines enter and leave the orders when performing the CCF calculation. BERVMX defines the excluded spectral lines from the correlation due to the orbital movement of the Earth toward the star. Our goal is to introduce a shift equivalent to a BERV value lower than the BERVMX, so that we do not need to exclude new lines. 14
1505.07713
1
1505
2015-05-28T15:02:03
Effect of O3 on the atmospheric temperature structure of early Mars
[ "astro-ph.EP" ]
Ozone is an important radiative trace gas in the Earth's atmosphere. The presence of ozone can significantly influence the thermal structure of an atmosphere, and by this e.g. cloud formation. Photochemical studies suggest that ozone can form in carbon dioxide-rich atmospheres. We investigate the effect of ozone on the temperature structure of simulated early Martian atmospheres. With a 1D radiative-convective model, we calculate temperature-pressure profiles for a 1 bar carbon dioxide atmosphere. Ozone profiles are fixed, parameterized profiles. We vary the location of the ozone layer maximum and the concentration at this maximum. The maximum is placed at different pressure levels in the upper and middle atmosphere (1-10 mbar). Results suggest that the impact of ozone on surface temperatures is relatively small. However, the planetary albedo significantly decreases at large ozone concentrations. Throughout the middle and upper atmospheres, temperatures increase upon introducing ozone due to strong UV absorption. This heating of the middle atmosphere strongly reduces the zone of carbon dioxide condensation, hence the potential formation of carbon dioxide clouds. For high ozone concentrations, the formation of carbon dioxide clouds is inhibited in the entire atmosphere. In addition, due to the heating of the middle atmosphere, the cold trap is located at increasingly higher pressures when increasing ozone. This leads to wetter stratospheres hence might increase water loss rates on early Mars. However, increased stratospheric H2O would lead to more HOx, which could efficiently destroy ozone. This result emphasizes the need for consistent climate-chemistry calculations to assess the feedback between temperature structure, water content and ozone chemistry. Furthermore, convection is inhibited at high ozone amounts, leading to a stably stratified atmosphere.
astro-ph.EP
astro-ph
Effect of O3 on the atmospheric temperature structure of early Mars P. von Parisa,b,∗, F. Selsisa,b, M. Godoltc, J.L. Grenfellc, H. Rauerc,d, B. Strackec aUniv. Bordeaux, LAB, UMR 5804, F-33270, Floirac, France bCNRS, LAB, UMR 5804, F-33270, Floirac, France cInstitut fur Planetenforschung, Deutsches Zentrum fur Luft- und Raumfahrt (DLR), Rutherfordstr. 2, 12489 dZentrum fur Astronomie und Astrophysik (ZAA), Technische Universitat Berlin, Hardenbergstr. 36, 10623 Berlin, Germany Berlin, Germany Abstract Ozone is an important radiative trace gas in the Earth's atmosphere and has also been detected on Venus and Mars. The presence of ozone can significantly influence the thermal structure of an atmosphere due to absorption of stellar UV radiation, and by this e.g. cloud formation. Photochemical studies suggest that ozone can form in carbon dioxide-rich atmospheres. Therefore, we investigate the effect of ozone on the temperature structure of simulated early Martian atmospheres. With a 1D radiative-convective model, we calculate temperature-pressure profiles for a 1 bar carbon dioxide atmosphere containing various amounts of ozone. These ozone profiles are fixed, parameterized profiles. We vary the location of the ozone layer maximum and the concentration at this maximum. The maximum is placed at different pressure levels in the upper and middle atmosphere (1-10 mbar). Results suggest that the impact of ozone on surface temperatures is relatively small. However, the planetary albedo significantly decreases at large ozone concentrations. Throughout the middle and upper atmospheres, temperatures increase upon introduc- ing ozone due to strong UV absorption. This heating of the middle atmosphere strongly reduces the zone of carbon dioxide condensation, hence the potential formation of car- bon dioxide clouds. For high ozone concentrations, the formation of carbon dioxide clouds is inhibited in the entire atmosphere. In addition, due to the heating of the middle atmosphere, the cold trap is located at increasingly higher pressures when in- creasing ozone. This leads to wetter stratospheres hence might increase water loss rates on early Mars. However, increased stratospheric H2O would lead to more HOx, which could efficiently destroy ozone by catalytic cycles, essentially self-limiting the increase of ozone. This result emphasizes the need for consistent climate-chemistry calculations to assess the feedback between temperature structure, water content and ozone chemistry. Furthermore, convection is inhibited at high ozone amounts, leading to a stably stratified atmosphere. ∗Corresponding author: email [email protected], tel. +33 (0)557 77 6131 Preprint submitted to Elsevier July 11, 2018 Keywords: early Mars: habitability, atmospheres 1. Introduction The climate and, consequently, the habitability of early Mars are a long-standing question that has not yet been answered conclusively. From an atmospheric modeling point of view, short episodes of a warm, wet climate, followed by long periods of arid, cold climates, seem to be favored for the Noachian period about 3.8 billion years ago (e.g., Segura et al. 2008, Wordsworth et al. 2013, Halevy and Head 2014). An important input for atmospheric modeling studies is the atmospheric compo- sition, together with the surface pressure. Most previous studies (e.g., Kasting 1991, Wordsworth et al. 2013) assumed mixed carbon dioxide (CO2)-water vapor (H2O) at- mospheres. In these studies, the partial pressure of carbon dioxide was of the order of bars, consistent with outgassing model studies (e.g., Phillips et al. 2001, Grott et al. 2011) and in-situ analyses (e.g., Manga et al. 2012, Kite et al. 2014). In 1D models, the atmospheric H2O content is usually controlled by a fixed relative humidity to simulate a hydrological cycle. However, other atmospheric (trace) gases such as molecular nitrogen (N2), molec- ular hydrogen (H2), methane (CH4) or sulphur dioxide (SO2) might have been present in the early Mars atmosphere. The initial N2 inventory is thought to be relatively large (e.g., McKay and Stoker 1989) and can provide modest surface warming (von Paris et al. 2013b). Methane has also been suggested to warm the atmosphere (e.g., Postawko and Kuhn 1986, Ramirez et al. 2014), but the effect on surface temperature is small. Ramirez et al. (2014) find that H2-induced warming could raise surface temperatures above freezing, assuming H2 was a major atmospheric constituent (around 20 % vol- ume mixing ratio). The effect of SO2 has been investigated by a number of studies (e.g., Postawko and Kuhn 1986, Yung et al. 1997, Halevy et al. 2007, Johnson et al. 2008, Tian et al. 2010, Mischna et al. 2013), but results suggest that it most likely does not contribute strongly to warming the surface. One proposed solution to warming early Mars has been the formation of carbon dioxide clouds (e.g., Pierrehumbert and Erlick 1998, Forget and Pierrehumbert 1997). Early 1D modeling studies suggested that the needed warming however strongly de- pends on the assumed cloud cover, which would have to be nearly 100% (e.g., Mischna et al. 2000). Following time-dependent 1D modeling studies by Colaprete and Toon (2003) found this to be unrealistic, a conclusion also supported by recent 3D studies (e.g., Forget et al. 2013, Wordsworth et al. 2013). In addition, Kitzmann et al. (2013) suggested that previous radiative transfer algorithms probably strongly overestimated the warming associated with carbon dioxide clouds. The formation of carbon dioxide clouds depends on the temperature profile in the middle atmosphere. As shown by, e.g., Yung et al. (1997) or Ramirez et al. (2014), UV absorption by SO2 or near-IR absorption by CH4 can effectively inhibit carbon dioxide condensation, or at least reduce the cloud formation region. So far, ozone (O3) has not been considered in the context of early Mars. Like SO2, ozone has strong UV and visible absorption bands. On Earth, these are responsi- 2 ble for the pronounced stratospheric temperature maximum. Ozone has been detected in the CO2-rich atmospheres of both Venus (e.g., Montmessin et al. 2011) and Mars (e.g., Lebonnois et al. 2006). Furthermore, numerous photochemical modeling stud- ies suggests that ozone can be formed from abiotically produced oxygen (e.g., Sel- sis et al. 2002, Segura et al. 2007, Domagal-Goldman and Meadows 2010; Domagal- Goldman et al. 2014). In addition, Wordsworth and Pierrehumbert (2014) propose a (non-chemical) abiotic oxygen formation process based on hydrogen escape and water cold-trapping. Therefore, it is reasonable to assume that on early Mars, ozone could have been also, to a certain extent, present in the atmosphere. Hence, in this study, we explore the effect of ozone on the temperature structure of early Mars The paper is structured as follows: Section 2 presents the atmospheric model and the simulations performed. Results are shown and discussed in Sect. 3 and conclusions in Sect. 4. 2. Computational details 2.1. Model description We use a 1D, steady-state, cloud-free radiative-convective atmosphere model to calculate globally, diurnally averaged temperature and H2O profiles. The model is originally based on Kasting et al. (1984a) and Kasting et al. (1984b). Further code developments are described in, e.g., Kasting (1988), Mischna et al. (2000), von Paris et al. (2008) or von Paris et al. (2010). The model atmospheres are divided into 52 levels. Temperature profiles in the up- per atmosphere are obtained by solving the radiative transfer equation (38 bands for incoming stellar radiation, 25 bands for planetary and atmospheric thermal radiation). For incoming stellar radiation (0.2-4.5 µm), Rayleigh scattering (by N2, H2O, CO2 and CH4) and molecular absorption (by H2O, CO2, CH4 and O3) contribute to the opac- ity. To allow for scattering, the angular integration of the radiative transfer equation is performed using a 2-stream code (Toon et al. 1989). For thermal radiation (1-500 µm), molecular absorption (by H2O, CO2, CH4 and O3, see below, Sect. 2.2) and contin- uum absorption (by N2, H2O and CO2) are considered. N2 continuum data is taken from Borysow and Frommhold (1986) and Lafferty et al. (1996), whereas H2O self and foreign continua as well as the CO2 foreign continuum are incorporated following Clough et al. (1989). The CO2 collision-induced self continuum is described follow- ing Kasting et al. (1984b). In the lower atmosphere, the model performs convective adjustment such that temperature profiles follow the wet adiabat. The formulation of the adiabatic lapse rate takes into account the condensation of H2O or CO2 (Kasting 1991). The super-saturation ratio for the onset of CO2 condensation is unity (but see Glandorf et al. (2002) for a different estimate for early Mars). H2O profiles are re-calculated for each model time step, according to local tem- perature and using a fixed relative humidity (RH) profile. The concentrations of the other species are fixed at the start of the calculations and only adjusted according to condensation of water and carbon dioxide at the surface (von Paris et al. 2013a). For more details, we refer to von Paris et al. (2008) and references therein. 3 Table 1: Spectral intervals for the new IR radiative transfer scheme and species considered contributing species Interval 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 2.2. Model improvements range [cm−1] 7,470 - 10,000 6,970 - 7,470 6,000 - 6,970 5,350 - 6,000 4,600 - 5,350 4,100 - 4,600 3,750 - 4,100 3,390 - 3,750 3,050 - 3,390 2,750 - 3,050 2,400 - 2,750 2,250 - 2,400 2,150 - 2,250 2,000 - 2,150 1,850 - 2,000 1,400 - 1,850 1,100 - 1,400 1,000 - 1,100 905 - 1,000 820 - 905 730 - 820 600 - 730 525 - 600 460 - 525 20 - 460 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4,O3 CO2,H2O,O3 CO2,H2O,O3 CO2,H2O,CH4,O3 CO2,H2O,CH4 H2O,CH4,O3 The original IR radiative transfer scheme MRAC (as used in, e.g., von Paris et al. 2008, von Paris et al. 2010, von Paris et al. 2013b) only considered gaseous absorption by H2O and CO2 for the calculation of opacities. For this work, a new, updated MRAC version has been created that also considers CH4 and O3 for gaseous absorption. Table 1 shows the spectral intervals as well as the species included in these inter- vals. Line positions and line strengths for both CH4 and O3 have been taken from the Hitran 2008 database (Rothman et al. 2009). Line strengths are converted from the reference temperature of 296 K to the desired temperature following Norton and Rins- land (1991). Cross sections are then obtained with the MIRART-Squirrl line-by-line radiative transfer code (Schreier and Schimpf 2001; Schreier and Bottger 2003) using 106 spectral points per band. The line shape is taken to be a Voigt line profile with a cutoff of 10 cm−1. The foreign component of the Lorentz broadening is taken directly from the Hitran database, i.e. for air. The self component of the Lorentz broadening was calculated assuming a volume mixing ratio (vmr) of 10−6 for both CH4 and O3, 4 Table 2: Interpolation and range of the grid variables in the new IR radiative transfer scheme Quantity Temperature Pressure H2O concentration carbon dioxide concentration CH4 concentration O3 concentration Range 100-400 K 10−5-1.5 bar 10−9-1 10−6-1 10−8-10−2 10−8-10−2 Interpolation linear in T linear in log(p) linear in vmr linear in vmr linear in vmr linear in vmr Comments ∆ T=50 K ∆ log(p)=0.1 1-4 points per order of magnitude 1-4 points per order of magnitude 1-4 points per order of magnitude 1-4 points per order of magnitude 3.55·10−4 for CO2 and 10−3 for water. Numerical tests showed that except for high water concentrations ((cid:38)10−2), the effect of the self component on the cross sections was negligible. The chosen line cutoff in this work is relatively short. Tests with a line-by-line radiative transfer code with line cut-offs at 25 cm−1 have shown that the overall flux change rarely exceeds 5 % in a given spectral band. For the total integrated outgo- ing thermal flux, we find a difference of about 1 %, equal to a radiative forcing of a few W m−2. This could lead to a small change in surface temperature of a few K (see also, for example, Wordsworth et al. 2010). However, this is not expected to qualita- tively alter our conclusions. Furthermore, as pointed out by, e.g., Halevy et al. (2009), Wordsworth et al. (2010) or Mischna et al. (2012), assuming a Voigt line profile is probably not a well-justified choice in dense CO2 atmospheres where the far wings of lines are substantially sub-Lorentzian. Sensitivity tests by Wordsworth et al. (2010) or von Paris et al. (2010) showed that the influence on surface temperature can be large (of the order of 10 K). Still, this is again not likely to influence the conclusions regarding the impact of ozone on the thermal structure in the upper and mid atmosphere. The interpolation variables are temperature, log(pressure) and the individual con- centrations ci of the absorbing species. The concentration grid has four dimensions, one for H2O, CO2, CH4 and O3, respectively. Table 2 summarizes the range of the interpolation variables and the type of interpolation. Figure 1 illustrates our approach to calculate the needed k distributions. In every spectral interval, for each temperature-pressure point (T/p), cross sections σi have been calculated for the contributing species i (see Table 1). These are then combined to obtain effective cross sections σeff for the different gas mixtures according to assumed molecular concentrations ci from Table 2: (cid:88) σeff = ci · σi (1) From the effective cross sections σeff, k distributions have been obtained that are then used in the model for the radiative transfer calculations with a correlated-k ap- proach. i 2.3. Model verification We tested the new model against results for early Mars from von Paris et al. (2013b). As an example, Figure 2 shows the temperature profiles and differences for a 1 bar car- 5 Figure 1: Flowchart for calculation of k distributions. bon dioxide case (without N2). It is clearly seen that both models compare relatively well with each other. The small discrepancies are due to a slightly different approach to calculate the line broadening parameters for the k distributions at high pressures, com- pared to earlier work (e.g., von Paris et al. 2008, 2010). In the line-by-line radiative transfer code used to calculate the cross sections (see above), the broadening pressure pb of the gas species i is calculated from the ideal gas law, i.e. pb = ci · p instead of using a column-density correction to calculate the actual broadening pressure (e.g., see Kasting 1987). However, these differences do not affect the conclusions of previous studies. The second model verification uses a modern Earth reference profile. Ozone and methane concentrations were taken from Grenfell et al. (2011). We run simulations with two different thermal radiative transfer schemes. First, we used RRTM (Mlawer et al. 1997), a scheme that is widely used in 1D and 3D exoplanet and Earth climate studies (e.g., Segura et al. 2003, 2005, Roeckner et al. 2006, Grenfell et al. 2007b, Kaltenegger et al. 2011, Stevens et al. 2013). A second simulation was done with the new scheme presented in this work. Figure 3 shows the resulting temperature profiles (left panel), the temperature dif- ference between both models (center panel) and a flux comparison for the upwelling thermal flux (right panel). For the latter, we compared calculated model IR fluxes with fluxes calculated by the high-resolution line-by-line code MIRART-Squirrl. Concerning the temperature profiles, the agreement in the troposphere and the lower stratosphere is very good. In the upper stratosphere, small discrepancies can be seen, reaching up to about 3 K. However, the overall agreement is good, and such small temperature differences are not expected to have a large influence neither on ob- servables (e.g. spectral signatures) nor on temperature-dependent chemistry. For the IR thermal upwelling flux, it is clearly seen that the fluxes do not differ by more than a few % throughout the stratosphere. 6 line databasecross sectionsσH2O, σCO2, σO3, σCH4T, pcH2O, cCO2,cO3, cCH4effective cross section σeffk distributions Figure 2: Early Mars: Comparison of temperature profiles from von Paris et al. (2013b) and this work. Carbon dioxide saturation vapor pressure curve as green dashed line. Based on these verifications, we consider the new radiative scheme to be applicable to further planetary scenarios. 2.4. Early Mars scenarios Simulated model atmospheres were assumed to be composed of CO2, H2O and O3. The CO2 partial pressure is fixed at 1 bar, consistent with upper limits placed on at- mospheric pressure during the Noachian (e.g., Phillips et al. 2001). H2O is calculated according to ambient temperature, assuming a fully saturated atmosphere (i.e., RH=1, see von Paris et al. 2013b). This approach is similar to most 1D studies of early Mars (e.g., Mischna et al. 2000, Colaprete and Toon 2003, Tian et al. 2010), but most likely over-estimates the amount of atmospheric water. The surface albedo was set to 0.21, close to the observed value of present Mars (Kieffer et al. 1977). This value of surface albedo allows the model to reproduce present Mars mean surface temperatures. Plan- etary gravity is 3.73 ms−2, and the orbital distance is set to 1.52 AU. We neglected any effect of eccentricity. However, even though Mars' orbit is eccentric today (e=0.09), and eccentricity cycles would occasionally drive the eccentricity to high values (e.g., 7 Figure 3: Modern Earth model verification: (Left panel) Comparison of temperature profiles calculated with RRTM (Mlawer et al. 1997) and MRAC (this work). (Center panel) ∆T (RRTM-MRAC) between both models, (Right panel) IR upwards flux ratio (climate model)/line-by-line. Laskar et al. 2004), the overall effect would probably be small (at e=0.2, the mean flux increases by only 2 % compared to the circular case). The incoming solar irradiation was set to Noachian conditions 3.8 billion years ago, i.e. 75 % of today's irradiation (e.g., Gough 1981). The input spectrum was taken from Gueymard (2004) and scaled at all wavelengths accordingly. Note that this approach is consistent with previous early Mars studies (e.g., Kasting 1991, Tian et al. 2010, Wordsworth et al. 2013). Other work, primarily on early Earth climate, not only scaled the spectrum, but also incorporate the variation of stellar parameters (e.g., Goldblatt et al. 2009) and enhanced UV radiation (e.g., Kunze et al. 2014). Enhanced UV radiation (see, e.g., Ribas et al. 2005) of the young Sun could have played an especially important role affecting ozone hence the thermal structure of the atmosphere. Ozone profiles strongly depend on the assumed oxygen content of the atmosphere, the incoming stellar UV radiation, and, in addition, on the catalytic cycles (HOx, NOx, etc.) operating in the atmosphere (e.g., Selsis et al. 2002, Segura et al. 2007, Grenfell et al. 2013). Detailed photochemical modeling is warranted for this problem, however such relatively complex models would necessitate many more unconstrained bound- 8 ary conditions and parameters (such as surface fluxes of oxidizing and reducing com- pounds, wet and dry deposition rates, vertical mixing, kinetic data, UV radiation field, etc.). Such complex models (e.g., Selsis et al. 2002, Segura et al. 2007, Hu et al. 2012, Domagal-Goldman et al. 2014, Tian et al. 2014, Grenfell et al. 2014) allow for an esti- mate of possible ranges of ozone concentrations. Motivated by the results from detailed photochemical studies, we perform a sensitivity study of the influence of ozone on tem- perature structure by inserting artificial ozone profiles CO3. These profiles are not the output of a photochemical model, but are parameterized as a function of pressure p based on a Gaussian profile (cid:16) − log p pmax 0.5 (cid:17)2  CO3 (p) = vmrmax · exp (2) where pmax is the pressure of the maximum concentration and vmrmax is the maxi- mum volume mixing ratio. We performed simulations for different combinations of vmrmax and pmax, as shown in Table 3. In Fig. 4, we show a sample of the chosen ozone profiles, together with a modern Earth mean profile. Table 3: Adopted parameter values for eq. 2 in this work. Parameter pmax [mbar] vmrmax [ppm] Values 0.1, 1, 10 0.1, 1, 10, 100 The ozone layer peak (its location and its magnitude) is the result of an interplay be- tween two main effects. On the one hand, ozone formation requires photolytic release (λ (cid:46) 200 nm) of atomic oxygen (O) e.g. from O2, CO2, H2O, which are usually more abundant on the lower levels (followed by fast, three-body formation of ozone). On the other hand, UV radiation is more abundant with increasing altitude, hence there is gen- erally a distinct maximum of ozone concentration at pressures of about 10−4-10−2 bar. This motivates the form of our artificial ozone profiles. Previous calculations by, e.g., Selsis et al. (2002) and Segura et al. (2007) suggest a maximum in ozone number density of up to 1012 cm−3, located at pressures around 1-10 mbar. They found ozone columns between 10−4-2 times the modern Earth column (Table 3 in Selsis et al. 2002 and Table 2 in Segura et al. 2007). Both studies focused on N2- or CO2-rich atmospheres, similar to the ones considered here. The oxygen needed for ozone formation was either provided by H2O or CO2. To cover the entire range of possible ozone concentrations found by Selsis et al. (2002) (i.e., their humid and dry CO2 cases as possible extremes) we varied the maximum ozone concentration between 0.1 and 100 ppm. As shown in Table 4, this leads to relatively thick ozone columns (up to 24x the modern Earth values) for a few cases. 9 Figure 4: Sample of chosen ozone profiles. Earth mean profile from Grenfell et al. (2011) as dash-dotted line. Table 4: O3 columns (in units of terrestrial columns, where one terrestrial column corresponds to 315 Dobson units) for the scenarios listed in Table 3. hhhhhhhhhhhhhhh vmrmax [ppm] pmax [mbar] 0.1 1 10 0.1 1 0.00023 0.00249 0.0248 0.0023 0.0249 0.248 10 0.023 0.249 2.47 100 0.23 2.48 24.7 3. Results and Discussion 3.1. Surface temperature Figure 5 shows the surface temperature as a function of the parameters in eq. 2. For all scenarios, the surface temperatures first slightly increase with increasing vmrmax, due to a slight decrease of planetary albedo and an additional greenhouse effect from ozone. For higher vmrmax, surface temperatures start to decrease with vmrmax (see below). However, for the pmax=0.1 and 1 mbar cases, the overall effect is rather small, of the order of 1-2 K. In contrast, for pmax=10 mbar, the increase at low vmrmax (up to 2 K) and the subsequent decrease towards higher values is much more pronounced, with a maximum temperature drop of about 8 K compared to the zero-ozone case. 3.2. Temperature structure Figure 6 shows the effect of ozone on the calculated temperature-pressure profiles. In the middle and upper atmosphere, up to 60 K of warming is seen, depending on 10 Figure 5: Effect of ozone on surface temperature. the choice of O3 parameters. This is of course due to the massive increase in stellar heating rates associated with the absorption of solar UV absorption by ozone. Heating rates increase by up to a factor of 5 compared to the no-ozone case (not shown). Two other effects are readily inferred from Fig. 6. First, for the pmax=10 mbar cases at vmrmax=10 and 100 ppm, temperature profiles no longer follow the carbon dioxide saturation pressure curve (black dashed line in Fig. 6). This suggests that carbon dioxide condensation, hence also the formation of carbon dioxide clouds, is prohibited throughout the entire atmosphere. For the other cases, the potential cloud formation zone is reduced, although the effect is rather minimal at low vmrmax or small pmax. Second, as suggested by the slope of the vmrmax=100 ppm, pmax=10 mbar case, the lower atmosphere is no longer convective. This is in contrast to the other scenarios, where the lapse rate follows a CO2 adiabat in the mid atmosphere and a wet H2O adiabat in the lower atmosphere. The strong UV absorption by ozone drastically changes the energy balance in the atmosphere. Figure 7 illustrates the effect of ozone on the planetary albedo. The plan- etary albedo is reduced by up to 40 % at large ozone layers. In terms of additional radiative energy deposited into the atmosphere, a reduction of albedo from 0.32 (zero- ozone) to 0.2 (maximum ozone effect) corresponds to a forcing of 13 Wm−2. However, 11 Figure 6: Effect of ozone on temperature structure. Reference case (zero ozone) as black plain line. Carbon dioxide saturation pressure shown as black dashed line. as illustrated in the left panel of Fig. 8, most of this energy is absorbed in the mid- dle atmosphere. As a consequence of the strong absorption, the incoming solar flux at the surface is reduced by about 20 %. Therefore, the total radiative flux in the lower atmosphere (sum of stellar and thermal fluxes) is reduced, coming closer to radiative equilibrium. Hence, less energy is available to drive convection. This then leads to a radiative instead of a convective lower atmosphere (as observed in Fig. 6). A reduced albedo corresponds to a larger thermal flux that must be emitted by the atmosphere to satisfy energy balance and radiative equilibrium at the top of the atmosphere. For optically thin atmospheres such as Earth's, this generally leads to an increase in surface temperature. In the case of dense carbon dioxide-dominated atmospheres considered here, the thermal emission originates mostly from pressures at around 10-100 mbar, since at higher pressures, the atmosphere is optically thick for thermal radiation at all wavelengths (see right panel of Fig. 8). For the considered scenarios with a substantial increase of local temperature (see Fig. 6), the local IR flux also increases. This balances the solar flux deposited in the atmosphere. Most of the outgoing thermal flux stems from emission in the 15 µm band of carbon dioxide which becomes optically thick at pressures around 10-100 mbar. 12 Figure 7: Effect of ozone on calculated planetary albedo. 3.3. Discussion 3.3.1. Oxygen content and catalytic chemistry The assumed ozone profiles are dependent on the oxygen content of the early Mars atmosphere. Given that modern Mars has only little oxygen in the atmosphere (around 10−3 vmr, e.g., Yung and deMore 1999), it seems likely that our high-ozone cases will over-estimate the total ozone column by a generous margin. However, the formation of ozone also depends on the assumed UV radiation field (e.g., Segura et al. 2007, Domagal-Goldman and Meadows 2010, Domagal-Goldman et al. 2014, Grenfell et al. 2014, Tian et al. 2014) which is relatively poorly known for the young Sun (Ribas et al. 2005). Observations of younger solar-type stars suggest that far-UV and X-ray emis- sions strongly decrease with age (e.g., Ribas et al. 2005) whereas the total luminosity is generally assumed to increase with age (e.g., Gough 1981, Caldeira and Kasting 1992). Therefore, the ratio between oxygen (hence ozone) production via photolysis (in the far-UV) and photolytic ozone destruction (in the near-UV) could presumably change dramatically and produce largely varying ozone concentrations (see, e.g., sensitivity studies by Segura et al. 2007, Grenfell et al. 2014 or Tian et al. 2014). Therefore, reli- able quantitative estimates of ozone on early Mars are challenging to obtain. Detailed photochemical models would probably only allow for an order-of-magnitude estimate 13 Figure 8: Effect of ozone on downwards solar (left) and upwelling thermal fluxes (right). Captions corre- spond to both panels. of such quantities, since boundary conditions (such as wet deposition of hydrogenated species and oxidation of reduced volcanic gases) are essentially unconstrained. In addition, ozone chemistry heavily depends on the operating catalytic cycles in the atmosphere (e.g., Crutzen 1970, Grenfell et al. 2013). On Earth, most of the cycles are related to NOx and HOx. Possible NOx sources on early Mars would be cosmic rays (e.g., Jackman et al. 1980) or lightning (e.g., Segura and Navarro-Gonz´alez 2005), but these are hard to estimate. It is unclear whether significant HOx cycles could op- erate on early Mars, since calculated stratospheres are very dry (see Fig. 9), with vmr about 10−8-10−6. This is up to several hundred times drier than in the modern Earth stratosphere (H2O vmr of a few times 10−6 up to 10−5). Note, however, that HOx cycles are responsible for the low oxygen content in the present Mars atmosphere (e.g., Nair et al. 1994, Yung and deMore 1999, Stock et al. 2012a,b) since they quickly recycle the products of CO2 photolysis (CO+O) back to CO2. Without such recycling, the martian atmosphere would be more O2-rich than today, probably at around 5-10 % O2. CO2 would still remain the dominant atmospheric species (e.g., Nair et al. 1994). This can be explained by the fact that at increasingly higher O2 concentrations, O2 UV ab- sorption starts to shield the CO2 from becoming photolyzed since O2 photodissociation 14 cross sections are much larger than corresponding CO2 cross sections (see, e.g., Yung and deMore 1999, Selsis et al. 2002). Even on Venus with a very dry stratosphere, HOx cycles operate, using HCl photolysis as source for H (Yung and deMore 1999). As an alternative destruction cycle, SOx cycles could be possible, since higher amounts of sulphur-bearing species are expected in the early Mars atmosphere (e.g., Farquhar et al. 2000, Halevy et al. 2007, Johnson et al. 2008, Ramirez et al. 2014). Investigating these possibilities is however beyond the scope of this sensitivity study that focuses on climatic effects. 3.3.2. Eddy diffusion and convection As a result of the suppression of convection in the lower atmosphere (see Fig. 6) for the high-ozone case, the atmosphere becomes stably stratified and vertical transport is significantly reduced. This has a potentially large impact on (photo)chemistry in such atmospheres. In general, photochemical networks include photochemical reactions (in- cluding e.g. 3-body reactions, photolysis, decomposition reactions, etc.) as well as a vertical mixing, which is parameterized by so-called Eddy diffusion. For the convective troposphere, 1D photochemical models of Mars or Earth use constant, relatively high, Eddy diffusion coefficients (e.g., Massie and Hunten 1981, Yung and deMore 1999). In a stably stratified atmosphere, the diffusion is most likely less efficient. The magnitude of the effect on convection and transport depends, among others, on the topography. Thus, full 3D dynamic simulations are probably needed to investigate the influence on chemistry. This is, for modern Earth and Earth-like (terrestrial) exoplanets, mostly important for biogenic trace gases such as methane and nitrous oxide. In the lower at- mosphere, these are more strongly influenced by transport rather than chemistry (e.g., Segura et al. 2003, 2005 Grenfell et al. 2007a,b). Exploring this issue, while beyond the scope of this work, would however be inter- esting in the context of fully coupled climate-chemistry simulations. 3.3.3. Water loss from early Mars The very thin atmosphere of current Mars is likely shaped by significant atmo- spheric loss. Isotopic signatures of atmospheric escape have been found for nitrogen and oxygen (e.g., Jakosky and Phillips 2001, Fox and Ha´c 2010, Gillmann et al. 2011). Recent D/H isotopic ratio measurements of the Curiosity rover on Mars suggest that hydrogen escape and water loss has been ongoing over the last few billion years (e.g., Mahaffy et al. 2015) after the (presumed to be dense) Noachian atmosphere has been eroded. Modeling the escape of water is a very important issue for atmospheric evo- lution and habitability (e.g., Kasting 1988, Kulikov et al. 2006). The water loss rate depends on the stratospheric water content, where water is photolyzed and subsequent loss of H atoms occurs (e.g., Kasting 1988, Kasting et al. 1993). Figure 9 shows the calculated H2O profiles. Note that we assume a relative hu- midity of unity throughout the atmosphere hence H2O concentrations are likely to be over-estimated in the mid- to upper atmosphere. On Earth, the stratospheric H2O con- tent is mainly controlled by the efficient cold trap near the tropopause (around 10 km altitude). With increasing temperatures due to the ozone layer, water is efficiently trapped in the troposphere. This leads to the dry stratosphere and, consequently, low water escape rates. 15 Figure 9: Effect of ozone on calculated water profiles. For the early Mars scenarios considered here, cold traps develop even in the absence of temperature inversions. However, given that surface temperatures are very low, and the cold trap is located high in the atmosphere, stratospheric H2O concentrations are very low. This suggests that water loss might be very slow on early Mars. However, when increasing the ozone concentrations in the model atmospheres, the altitude of the cold trap changes dramatically, due to the change in the temperature structure (see Fig. 6). This leads to a significant increase of stratospheric H2O content (up to two orders of magnitude), and possibly impacts the water escape rates. Hence, results suggest that water escape rates might be sensitively influenced by the ozone content. Note, how- ever, that an increase in stratospheric H2O concentrations could also lead to enhanced destruction of ozone through catalytic HOx chemistry (see, e.g., the parametric simu- lations of Tian et al. 2014). Therefore, changes in cold-trap locations might limit the increase of ozone, hence limit also H2O escape. To further investigate this issue would need consistent climate-chemistry models to assess this feedback between vertical tem- perature structure, stratospheric water content and ozone chemistry. To estimate the amount of water lost via atmospheric escape, we adopted the same approach as Wordsworth and Pierrehumbert (2014). Escape was assumed to be diffusion- limited and controlled by the cold-trap temperature Tct and cold-trap water concentra- tions. The diffusion coefficient of H2O through a CO2 atmosphere was taken from approximative equations of Marrero and Mason (1972). These approximations lead to water escape fluxes of, even under the most favorable escape conditions, only about 1 cm of water lost within 109 years (1 Gyr). This has to be compared to the actual atmospheric water column of about a few 10−4 m of water in our simulations. Furthermore, the original Martian water inventory is estimated to be much 16 larger, about 102-103 m (e.g., McKay and Stoker 1989, Lammer et al. 2013, Lasue et al. 2013), Currently, most of the water is thought to either reside in the cryosphere or to be bound to minerals in the crust (see discussion in, e.g., Lammer et al. 2013 and Lasue et al. 2013). In addition, we point out that current best estimates of the hydrogen escape flux suggest about 1-2 m of water lost in 1 Gyr (e.g., Chassefi`ere and Leblanc 2004, Zahnle 2015), about 2 orders of magnitude more than we calculate. This is due to the fact that molecular and atomic hydrogen are the main hydrogen-bearing species in the upper Martian atmosphere, not H2O, as we assumed above. H and H2 have concentrations about 2 orders of magnitude higher than H2O (e.g., Nair et al. 1994, Yung and deMore 1999). H and H2 are derived from H2O photolysis, and since water is controlled by surface conditions and constantly replenished, H and H2 can build up until concen- trations reach an equilibrium between the photolytic source and atmospheric escape. Addressing the connection between cold-trap location, thermal structure and H2O es- cape necessitates coupled climate-chemistry simulations, which is a potential direction of future research. 3.3.4. Spectral appearance In an exoplanetary context, an interesting question with respect to ozone is the possibility of detecting it remotely with spectroscopic observations. It has been pro- posed as a so-called biosignature gas (e.g., L´eger et al. 1996, Selsis 2000, Schindler and Kasting 2000, Des Marais et al. 2002, Selsis et al. 2002), and much work has been invested into atmospheric and spectral modeling regarding detection and interpretation of ozone signatures (e.g., Selsis et al. 2002, Segura et al. 2003, 2007, Kaltenegger et al. 2007, Kaltenegger and Traub 2009, Rauer et al. 2011, Hedelt et al. 2013, Grenfell et al. 2014). Considering early Mars as a potential exoplanet, Fig. 10 shows the resulting emis- sion spectra for five selected cases. They have been calculated with the same high- resolution line-by-line code used for the model verification (Schreier and Bottger 2003) =100. As pointed out by and were then binned to a coarse spectral resolution of R= λ ∆λ Selsis et al. (2002), the ozone band at 9.6 µm coincides with a carbon dioxide hot band and lies in the wings of the strong dimer absorption feature (see, e.g., Wordsworth et al. 2010 for the most recent carbon dioxide continuum absorption data). Therefore, except for the strongest ozone column, the shape of the spectral band is not greatly sensitive to atmospheric ozone amount. To illustrate this, Fig. 11 zooms in on the spectral region around the 9.6 µm fun- damental band of ozone. For comparison, we also show line-by-line calculations with removed ozone (marked "no O3" in Fig. 11) to investigate the effect of ozone on the spectral appearance and a possible masking by CO2. It is clear that an effect is only seen for the highest ozone columns (higher than the modern-day terrestrial column, Ta- ble 4). To detect this, relatively high spectral resolution (here, R=100) and reasonable signal-to-noise ratios (of the order of 4) are needed. These values are not achievable with currently planned instrumentation even for Earth-sized or larger planets, let alone a planet the size of Mars (see, e.g., Hedelt et al. 2013). The strong carbon dioxide fundamental band around 15 µm is very sensitive to middle-atmosphere temperatures (see Fig. 8), and the effect of the changing temper- 17 Figure 10: Emission spectra (spectral resolution R=100) for selected cases. ature structure is clearly apparent in this band. In the hypothetical case that all stel- lar parameters (including UV radiation field) were known, and consistent atmospheric modeling were used, then this band would constitute an excellent indirect probe of at- mospheric ozone content. We note, however, that other radiative gases could also be responsible for a temperature inversion (e.g., SO2, CH4, etc.). This then is an important challenge for retrieval algorithms (e.g., von Paris et al. 2013c, Irwin et al. 2014). In addition to that, the broad 6.3 µm band of water, superimposed by part of a carbon dioxide dimer absorption feature, is also sensitive to middle-atmosphere tem- peratures and thus would be an indirect probe of ozone content. Summarizing, in the cases considered for this work, ozone is hardly seen directly in the spectra, except at high ozone contents. Indirectly, however, due to its impact on the thermal structure, the effect of ozone is clearly seen in the spectra. As outlined in Kitzmann et al. (2011) or von Paris et al. (2013d), we use the model output of the stellar radiative transfer code to produce low-resolution albedo spectra. In Fig. 12, we show the UV-visible part of the spectrum. For the low-ozone reference case, the spectral albedo clearly follows the Rayleigh scattering slope expected for dense carbon dioxide atmospheres. The presence of ozone leads to a strong reduction of albedo in the Hartley (0.24 µm< λ<0.34 µm) and Chappuis (0.4 µm< λ <0.7 µm) bands. Additionally, the location of the ozone maximum also influences the spectral albedo in the Hartley band, as shown in Fig. 12. For cases where the maximum is located at higher pressures, the UV radiation penetrates deeper into the atmosphere and thus is more susceptible to Rayleigh scattering by carbon dioxide. 18 Figure 11: Zoom in the 9.6 µm band of ozone. Effect of including ozone in spectral calculations (R=100). 4. Conclusions Ozone is a radiatively important atmospheric species due to strong UV absorption bands. It can significantly change the atmospheric temperature structure and thus af- fect, e.g., cloud formation or atmospheric transport. We have investigated the influence of ozone on the atmospheric temperature struc- ture of early Mars. We simulated a 1 bar carbon dioxide atmosphere with different, fixed ozone concentration profiles. To do this, we have developed a new IR radia- tive transfer scheme that incorporates molecular absorption of carbon dioxide, water, methane and ozone. Calculated IR cross sections cover a wide range of temperatures, pressures and concentrations. Hence, this scheme can be applied to a wide range of possible planetary scenarios. Results suggest that the impact on surface temperatures at small to moderate ozone concentrations is probably not large, of the order of 1-3 K, depending on the overall ozone column. The increase in surface temperature is mostly due to the decreased albedo, since ozone strongly absorbs the incoming UV and visible solar radiation. For high ozone concentrations, surface temperatures drop by up to 8 K due to a change in energy balance. In the upper and middle atmosphere, temperatures increased by up to 60 K upon introducing ozone. The resulting increase in thermal flux balances the 19 Figure 12: Effect of ozone on spectral albedo. radiative forcing of UV absorption and leads to the observed surface cooling at high ozone concentrations. As a consequence of the stratospheric warming, the cloud forming region is re- duced. In the case of a thick ozone layer (comparable to or larger than the terrestrial ozone layer), cloud formation is inhibited completely. For large ozone columns, convection is inhibited. Instead model atmospheres are fully radiative due to a strong reduction of incoming solar flux in the lower atmosphere. This probably decreases the vertical transport. Upon increasing the ozone content in the model atmospheres, stratospheric water content strongly increased, due to a change in thermal structure and a change in cold trap location. Compared to zero-ozone simulations, water concentrations increased by up to two orders of magnitudes. This is important for assessing possible atmospheric water loss on early Mars, but could be a self-limiting effect due to enhanced catalytic HOx cycles that destroy ozone. Future work aims at performing fully coupled climate-chemistry simulations of early Mars as well as carbon dioxide-dominated atmospheres near the outer boundary of the habitable zone. Acknowledgements This study has received financial support from the French State in the frame of the "Investments for the future" Programme IdEx Bordeaux, reference ANR-10-IDEX- 03-02. This work has been partly supported by the Postdoc Program "Atmospheric dynamics and Photochemistry of Super Earth planets" of the Helmholtz Gemeinschaft 20 (HGF). We thank the two anonymous reviewers for their positive and constructive feed- back. References A. Borysow and L. Frommhold. Collision-induced rototranslational absorption spectra of N2-N2 pairs for temperatures from 50 to 300 K. Astrophys. J., 311:1043–1057, December 1986. doi: 10.1086/164841. K. Caldeira and J. F. Kasting. The life span of the biosphere revisited. Nature, 360: 721–723, December 1992. doi: 10.1038/360721a0. E. Chassefi`ere and F. Leblanc. Mars atmospheric escape and evolution; interaction with the solar wind. Planet. Space Science, 52:1039–1058, September 2004. doi: 10.1016/j.pss.2004.07.002. S. Clough, F. Kneizys, and R. Davies. Line Shape and the Water Vapor Continuum. Atm. Research, 23:229–241, July 1989. A. Colaprete and O. B. Toon. Carbon dioxide clouds in an early dense Martian atmo- sphere. J. Geophys. Res., 108:6–1, April 2003. doi: 10.1029/2002JE001967. P. J. Crutzen. The influence of nitrogen oxides on the atmospheric ozone con- tent. Quart. J. Royal Meteor. Soc., 96:320–325, April 1970. doi: 10.1002/qj. 49709640815. D. J. Des Marais, M. O. Harwit, K. W. Jucks, J. F. Kasting, D. N. C. Lin, J. I. Lunine, J. Schneider, S. Seager, W. A. Traub, and N. J. Woolf. Remote Sensing of Planetary Properties and Biosignatures on Extrasolar Terrestrial Planets. Astrobiology, 2:153– 181, June 2002. doi: 10.1089/15311070260192246. S. Domagal-Goldman and V. Meadows. Abiotic buildup of ozone. ASP Conference Series, 430:152, 2010. S. D. Domagal-Goldman, A. Segura, M. W. Claire, T. D. Robinson, and V. S. Meadows. Abiotic Ozone and Oxygen in Atmospheres Similar to Prebiotic Earth. Astrophys. J., 792:90, September 2014. doi: 10.1088/0004-637X/792/2/90. J. Farquhar, J. Savarino, T. L. Jackson, and M. H. Thiemens. Evidence of atmospheric sulphur in the martian regolith from sulphur isotopes in meteorites. Nature, 404: 50–52, March 2000. F. Forget and R. T. Pierrehumbert. Warming Early Mars with Carbon Dioxide Clouds That Scatter Infrared Radiation. Science, 278:1273–1276, November 1997. F. Forget, R. Wordsworth, E. Millour, J.-B. Madeleine, L. Kerber, J. Leconte, E. Marcq, and R. M. Haberle. 3D modelling of the early martian climate under a denser CO2 atmosphere: Temperatures and CO2 ice clouds. Icarus, 222:81–99, January 2013. doi: 10.1016/j.icarus.2012.10.019. 21 J. L. Fox and A. Ha´c. Isotope fractionation in the photochemical escape of O from Mars. Icarus, 208:176–191, July 2010. doi: 10.1016/j.icarus.2010.01.019. C. Gillmann, P. Lognonn´e, and M. Moreira. Volatiles in the atmosphere of Mars: The effects of volcanism and escape constrained by isotopic data. Earth Plan. Science Letters, 303:299–309, March 2011. doi: 10.1016/j.epsl.2011.01.009. D. L. Glandorf, A. Colaprete, M. A. Tolbert, and O. B. Toon. CO2 Snow on Mars and Early Earth: Experimental Constraints. Icarus, 160:66–72, November 2002. doi: 10.1006/icar.2002.6953. C. Goldblatt, M. W. Claire, T. M. Lenton, A. J. Matthews, A. J. Watson, and K. J. Zahnle. Nitrogen-enhanced greenhouse warming on early Earth. Nature Geoscience, 2:891–896, December 2009. doi: 10.1038/ngeo692. D. O. Gough. Solar interior structure and luminosity variations. Solar Physics, 74: 21–34, November 1981. doi: 10.1007/BF00151270. J. L. Grenfell, J.-M. Griessmeier, B. Patzer, H. Rauer, A. Segura, A. Stadelmann, B. Stracke, R. Titz, and P. Von Paris. Biomarker Response to Galactic Cosmic Ray-Induced NOx And The Methane Greenhouse Effect in The Atmosphere of An Earth-Like Planet Orbiting An M Dwarf Star. Astrobiology, 7:208–221, February 2007a. doi: 10.1089/ast.2006.0129. J. L. Grenfell, B. Stracke, P. von Paris, B. Patzer, R. Titz, A. Segura, and H. Rauer. The response of atmospheric chemistry on earthlike planets around F, G and K Stars to small variations in orbital distance. Planet. Space Science, 55:661–671, April 2007b. doi: 10.1016/j.pss.2006.09.002. J. L. Grenfell, S. Gebauer, P. von Paris, M. Godolt, P. Hedelt, A. B. C. Patzer, B. Stracke, and H. Rauer. Sensitivity of biomarkers to changes in chemical emis- sions in the Earth's Proterozoic atmosphere. Icarus, 211:81–88, January 2011. doi: 10.1016/j.icarus.2010.09.015. J. L. Grenfell, S. Gebauer, M. Godolt, K. Palczynski, H. Rauer, J. Stock, P. von Paris, R. Lehmann, and F. Selsis. Potential Biosignatures in Super-Earth Atmospheres II. Photochemical Responses. Astrobiology, 13:415–438, May 2013. doi: 10.1089/ast. 2012.0926. J. L. Grenfell, S. Gebauer, P. v. Paris, M. Godolt, and H. Rauer. Sensitivity of biosig- natures on Earth-like planets orbiting in the habitable zone of cool M-dwarf Stars to varying stellar UV radiation and surface biomass emissions. Planet. Space Science, 98:66–76, August 2014. doi: 10.1016/j.pss.2013.10.006. M. Grott, A. Morschhauser, D. Breuer, and E. Hauber. Volcanic outgassing of CO2 and H2O on Mars. Earth Plan. Science Letters, 308:391–400, August 2011. doi: 10.1016/j.epsl.2011.06.014. C. Gueymard. The sun's total and spectral irradiance for solar energy applications and solar radiation models. Solar Energy, 76:423–453, 2004. 22 I. Halevy and J. W. Head, III. Episodic warming of early Mars by punctuated volcan- ism. Nature Geoscience, 7:865–868, December 2014. doi: 10.1038/ngeo2293. I. Halevy, M. T. Zuber, and D. P. Schrag. A Sulfur Dioxide Climate Feedback on Early Mars. Science, 318:1903–, December 2007. doi: 10.1126/science.1147039. I. Halevy, R. T. Pierrehumbert, and D. P. Schrag. Radiative transfer in CO2-rich paleoatmospheres. J. Geophys. Res., 114:18112, September 2009. doi: 10.1029/ 2009JD011915. P. Hedelt, P. von Paris, M. Godolt, S. Gebauer, J. L. Grenfell, H. Rauer, F. Schreier, F. Selsis, and T. Trautmann. Spectral features of Earth-like planets and their de- tectability at different orbital distances around F, G, and K-type stars. Astron. Astro- phys., 553:A9, May 2013. doi: 10.1051/0004-6361/201117723. R. Hu, S. Seager, and W. Bains. Photochemistry in Terrestrial Exoplanet Atmospheres. I. Photochemistry Model and Benchmark Cases. Astrophys. J., 761:166, December 2012. doi: 10.1088/0004-637X/761/2/166. P. G. J. Irwin, J. K. Barstow, N. E. Bowles, L. N. Fletcher, S. Aigrain, and J.-M. Lee. The transit spectra of Earth and Jupiter. Icarus, 242:172–187, November 2014. doi: 10.1016/j.icarus.2014.08.005. C. H. Jackman, J. E. Frederick, and R. S. Stolarski. Production of odd nitrogen in the stratosphere and mesosphere - An intercomparison of source strengths. J. Geophys. Res., 85:7495–7505, December 1980. doi: 10.1029/JC085iC12p07495. B. M. Jakosky and R. J. Phillips. Mars' volatile and climate history. Nature, 412: 237–244, July 2001. S. S. Johnson, M. A. Mischna, T. L. Grove, and M. T. Zuber. Sulfur-induced green- house warming on early Mars. J. Geophys. Res., 113:8005, August 2008. doi: 10.1029/2007JE002962. L. Kaltenegger and W. A. Traub. Transits of Earth-like Planets. Astrophys. J., 698: 519–527, June 2009. doi: 10.1088/0004-637X/698/1/519. L. Kaltenegger, W. A. Traub, and K. W. Jucks. Spectral Evolution of an Earth-like Planet. Astrophys. J., 658:598–616, March 2007. doi: 10.1086/510996. L. Kaltenegger, A. Segura, and S. Mohanty. Model Spectra of the First Potentially Habitable Super-Earth Gl581d. Astrophys. J., 733:35, May 2011. doi: 10.1088/ 0004-637X/733/1/35. J. F. Kasting. Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere. Precambrian Research, 34:205–229, February 1987. J. F. Kasting. Runaway and moist greenhouse atmospheres and the evolution of earth and Venus. Icarus, 74:472–494, June 1988. doi: 10.1016/0019-1035(88)90116-9. 23 J. F. Kasting. CO2 condensation and the climate of early Mars. November 1991. doi: 10.1016/0019-1035(91)90137-I. Icarus, 94:1–13, J. F. Kasting, J. B. Pollack, and T. P. Ackerman. Response of earth's atmosphere to increases in solar flux and implications for loss of water from Venus. Icarus, 57: 335–355, March 1984a. doi: 10.1016/0019-1035(84)90122-2. J. F. Kasting, J. B. Pollack, and D. Crisp. Effects of high CO2 levels on surface tem- perature and atmospheric oxidation state of the early earth. J. Atmospheric Chem., 1:403–428, 1984b. J. F. Kasting, D. P. Whitmire, and R. T. Reynolds. Habitable Zones around Main Sequence Stars. Icarus, 101:108–128, January 1993. doi: 10.1006/icar.1993.1010. H. H. Kieffer, T. Z. Martin, A. R. Peterfreund, B. M. Jakosky, E. D. Miner, and F. D. Palluconi. Thermal and albedo mapping of Mars during the Viking primary mission. J. Geophys. Res., 82:4249–4291, September 1977. doi: 10.1029/JS082i028p04249. E. S. Kite, J.-P. Williams, A. Lucas, and O. Aharonson. Low palaeopressure of the martian atmosphere estimated from the size distribution of ancient craters. Nature Geoscience, 7:335–339, May 2014. doi: 10.1038/ngeo2137. D. Kitzmann, A. B. C. Patzer, P. von Paris, M. Godolt, and H. Rauer. Clouds in the atmospheres of extrasolar planets. III. Impact of low and high-level clouds on the reflection spectra of Earth-like planets. Astron. Astrophys., 534:A63, October 2011. doi: 10.1051/0004-6361/201117375. D. Kitzmann, A. B. C. Patzer, and H. Rauer. Clouds in the atmospheres of extrasolar planets. IV. On the scattering greenhouse effect of CO2 ice particles: Numerical radiative transfer studies. Astron. Astrophys., 557:A6, September 2013. doi: 10. 1051/0004-6361/201220025. Y. N. Kulikov, H. Lammer, H. I. M. Lichtenegger, N. Terada, I. Ribas, C. Kolb, D. Langmayr, R. Lundin, E. F. Guinan, S. Barabash, and H. K. Biernat. Atmospheric and water loss from early Venus. Planet. Space Science, 54:1425–1444, November 2006. doi: 10.1016/j.pss.2006.04.021. M. Kunze, M. Godolt, U. Langematz, J. L. Grenfell, A. Hamann-Reinus, and H. Rauer. Investigating the early Earth faint young Sun problem with a general circulation model. Planet. Space Science, 98:77–92, August 2014. doi: 10.1016/j.pss.2013.09. 011. W. J. Lafferty, A. M. Solodov, A. Weber, W. B. Olson, and J.-M. Hartmann. Infrared collision-induced absorption by N2 near 4.3µm for atmospheric applications: mea- surements and empirical modeling. Applied Optics, 35:5911–5917, October 1996. doi: 10.1364/AO.35.005911. H. Lammer, E. Chassefi`ere, O. Karatekin, A. Morschhauser, P. B. Niles, O. Mousis, P. Odert, U. V. Mostl, D. Breuer, V. Dehant, M. Grott, H. Groller, E. Hauber, and L. B. S. Pham. Outgassing History and Escape of the Martian Atmosphere and 24 Water Inventory. Space Science Rev., 174:113–154, January 2013. doi: 10.1007/ s11214-012-9943-8. J. Laskar, A. C. M. Correia, M. Gastineau, F. Joutel, B. Levrard, and P. Robutel. Long Icarus, term evolution and chaotic diffusion of the insolation quantities of Mars. 170:343–364, August 2004. doi: 10.1016/j.icarus.2004.04.005. J. Lasue, N. Mangold, E. Hauber, S. Clifford, W. Feldman, O. Gasnault, C. Grima, S. Maurice, and O. Mousis. Quantitative Assessments of the Martian Hydrosphere. Space Science Rev., 174:155–212, January 2013. doi: 10.1007/s11214-012-9946-5. S. Lebonnois, E. Qu´emerais, F. Montmessin, F. Lef`evre, S. Perrier, J.-L. Bertaux, and F. Forget. Vertical distribution of ozone on Mars as measured by SPICAM/Mars Express using stellar occultations. J. Geophys. Res., 111:E09S05, September 2006. doi: 10.1029/2005JE002643. A. L´eger, J. M. Mariotti, B. Mennesson, M. Ollivier, J. L. Puget, D. Rouan, and J. Schneider. Could We Search for Primitive Life on Extrasolar Planets in the Near Future? Icarus, 123:249–255, October 1996. doi: 10.1006/icar.1996.0155. P. R. Mahaffy, C. R. Webster, J. C. Stern, A. E. Brunner, S. K. Atreya, P. G. Conrad, S. Domagal-Goldman, J. L. Eigenbrode, G. J. Flesch, L. E. Christensen, H. B. Franz, C. Freissinet, D. P. Glavin, J. P. Grotzinger, J. H. Jones, L. A. Leshin, C. Malespin, A. C. McAdam, D. W. Ming, R. Navarro-Gonzalez, P. B. Niles, T. Owen, A. A. Pavlov, A. Steele, M. G. Trainer, K. H. Williford, J. J. Wray, and aff14. The imprint of atmospheric evolution in the D/H of Hesperian clay minerals on Mars. Science, 347:412–414, January 2015. doi: 10.1126/science.1260291. M. Manga, A. Patel, J. Dufek, and E. S. Kite. Wet surface and dense atmosphere on early Mars suggested by the bomb sag at Home Plate, Mars. Geophys. Res. Letters, 39:L01202, January 2012. doi: 10.1029/2011GL050192. T. R. Marrero and E. A. Mason. Gaseous Diffusion Coefficients. J. Physical and Chemical Reference Data, 1:3–118, January 1972. doi: 10.1063/1.3253094. S. T. Massie and D. M. Hunten. tracer data. JC086iC10p09859. J. Geophys. Res., 86:9859–9868, October 1981. Stratospheric eddy diffusion coefficients from doi: 10.1029/ C. P. McKay and C. R. Stoker. The Early Environment and its Evolution on Mars: Implications for Life. Reviews of Geophysics, 27:189–214, 1989. doi: 10.1029/ RG027i002p00189. M. A. Mischna, J. F. Kasting, A. Pavlov, and R. Freedman. Influence of carbon dioxide clouds on early martian climate. Icarus, 145:546–554, June 2000. doi: 10.1006/icar. 2000.6380. M. A. Mischna, C. Lee, and M. Richardson. Development of a fast, accurate radiative transfer model for the Martian atmosphere, past and present. Journal of Geophysical Research (Planets), 117:E10009, October 2012. doi: 10.1029/2012JE004110. 25 M. A. Mischna, V. Baker, R. Milliken, M. Richardson, and C. Lee. Effects of obliquity and water vapor/trace gas greenhouses in the early martian climate. Journal of Geo- physical Research (Planets), 118:560–576, March 2013. doi: 10.1002/jgre.20054. E. J. Mlawer, S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough. Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102:16663–16682, July 1997. doi: 10.1029/ 97JD00237. F. Montmessin, J.-L. Bertaux, F. Lef`evre, E. Marcq, D. Belyaev, J.-C. G´erard, O. Ko- rablev, A. Fedorova, V. Sarago, and A. C. Vandaele. A layer of ozone detected in the nightside upper atmosphere of Venus. Icarus, 216:82–85, November 2011. doi: 10.1016/j.icarus.2011.08.010. H. Nair, M. Allen, A. D. Anbar, Y. L. Yung, and R. T. Clancy. A photochemical model of the martian atmosphere. Icarus, 111:124–150, September 1994. doi: 10.1006/ icar.1994.1137. R. H. Norton and C. P. Rinsland. ATMOS data processing and science analysis meth- ods. Applied Optics, 30:389–400, February 1991. doi: 10.1364/AO.30.000389. R. J. Phillips, M. T. Zuber, S. C. Solomon, M. P. Golombek, B. M. Jakosky, W. B. Banerdt, D. E. Smith, R. M. E. Williams, B. M. Hynek, O. Aharonson, and S. A. Hauck. Ancient Geodynamics and Global-Scale Hydrology on Mars. Science, 291: 2587–2591, March 2001. doi: 10.1126/science.1058701. R. T. Pierrehumbert and C. Erlick. On the Scattering Greenhouse Effect of CO2 Ice Clouds. J. Atmosph. Sciences, 55:1897–1903, May 1998. S. E. Postawko and W. R. Kuhn. Effect of the greenhouse gases (CO2, H2O, SO2) on Martian paleoclimate. J. Geophys. Res., 91:431, March 1986. doi: 10.1029/ JB091iB04p0D431. R. M. Ramirez, R. Kopparapu, M. E. Zugger, T. D. Robinson, R. Freedman, and J. F. Kasting. Warming early Mars with CO2 and H2. Nature Geoscience, 7:59–63, January 2014. doi: 10.1038/ngeo2000. H. Rauer, S. Gebauer, P. von Paris, J. Cabrera, M. Godolt, J. L. Grenfell, A. Belu, F. Selsis, P. Hedelt, and F. Schreier. Potential Biosignatures in Super-Earths Atmo- spheres I. Spectral appearance of super-Earths around M dwarfs. Astron. Astrophys., 529:A8, May 2011. doi: 10.1051/0004-6361/201014368. I. Ribas, E. F. Guinan, M. Gudel, and M. Audard. Evolution of the Solar Activity over Time and Effects on Planetary Atmospheres. I. High-Energy Irradiances (1-1700 Å). Astrophys. J., 622:680–694, March 2005. doi: 10.1086/427977. E. Roeckner, R. Brokopf, M. Esch, M. Giorgetta, S. Hagemann, L. Kornblueh, E. Manzini, U. Schlese, and U. Schulzweida. Sensitivity of Simulated Climate to Horizontal and Vertical Resolution in the ECHAM5 Atmosphere Model. J. Climate, 19:3771, 2006. doi: 10.1175/JCLI3824.1. 26 L. S. Rothman, I. E. Gordon, A. Barbe, D. C. Benner, P. F. Bernath, M. Birk, V. Boudon, L. R. Brown, A. Campargue, J.-P. Champion, K. Chance, L. H. Coudert, V. Dana, V. M. Devi, S. Fally, J.-M. Flaud, R. R. Gamache, A. Goldman, D. Jacque- mart, I. Kleiner, N. Lacome, W. J. Lafferty, J.-Y. Mandin, S. T. Massie, S. N. Mikhailenko, C. E. Miller, N. Moazzen-Ahmadi, O. V. Naumenko, A. V. Nikitin, J. Orphal, V. I. Perevalov, A. Perrin, A. Predoi-Cross, C. P. Rinsland, M. Rot- ger, M. Simeckov´a, M. A. H. Smith, K. Sung, S. A. Tashkun, J. Tennyson, R. A. Toth, A. C. Vandaele, and J. Vander Auwera. The HITRAN 2008 molecular spec- troscopic database. J. Quant. Spect. Rad. Trans., 110:533–572, June 2009. doi: 10.1016/j.jqsrt.2009.02.013. T. L. Schindler and J. F. Kasting. Synthetic Spectra of Simulated Terrestrial Atmo- Icarus, 145:262–271, May 2000. spheres Containing Possible Biomarker Gases. doi: 10.1006/icar.2000.6340. F. Schreier and U. Bottger. MIRART, a line-by-line code for infrared atmospheric radiation computations including derivatives. Atmospheric and Oceanic Optics, 16: 262–268, 2003. F. Schreier and B. Schimpf. A new efficient line-by-line code for high resolution atmo- spheric radiation computations incl. derivatives. In W.L. Smith and Y. Timofeyev, editors, IRS 2000: Current Problems in Atmospheric Radiation, Current Problems in Atmospheric Radiation. A. Deepak, 2001. A. Segura and R. Navarro-Gonz´alez. Nitrogen fixation on early Mars by volcanic lightning and other sources. Geophys. Res. Letters, 32:L5203, March 2005. doi: 10.1029/2004GL021910. A. Segura, K. Krelove, J. F. Kasting, D. Sommerlatt, V. Meadows, D. Crisp, M. Co- hen, and E. Mlawer. Ozone Concentrations and Ultraviolet Fluxes on Earth-Like Planets Around Other Stars. Astrobiology, 3:689–708, December 2003. doi: 10.1089/153110703322736024. A. Segura, J. F. Kasting, V. Meadows, M. Cohen, J. Scalo, D. Crisp, R. A. H. Butler, and G. Tinetti. Biosignatures from Earth-Like Planets Around M Dwarfs. Astrobi- ology, 5:706–725, December 2005. doi: 10.1089/ast.2005.5.706. A. Segura, V. S. Meadows, J. F. Kasting, D. Crisp, and M. Cohen. Abiotic formation of O2 and O3 in high-CO2 terrestrial atmospheres. Astron. Astrophys., 472:665–679, 2007. T. L. Segura, O. B. Toon, and A. Colaprete. Modeling the environmental effects of moderate-sized impacts on Mars. J. Geophys. Res., 113:E11007, November 2008. doi: 10.1029/2008JE003147. F. Selsis. Review: Physics of Planets I: Darwin and the Atmospheres of Terrestrial Planets. ESA Special Publication, 451:133, 2000. 27 F. Selsis, D. Despois, and J.-P. Parisot. Signature of life on exoplanets: Can Darwin produce false positive detections? Astron. Astrophys., 388:985–1003, June 2002. doi: 10.1051/0004-6361:20020527. B. Stevens, M. Giorgetta, M. Esch, T. Mauritsen, T. Crueger, S. Rast, M. Salzmann, H. Schmidt, J. Bader, K. Block, R. Brokopf, I. Fast, S. Kinne, L. Kornblueh, U. Lohmann, R. Pincus, T. Reichler, and E. Roeckner. Atmospheric component of the MPI-M Earth System Model: ECHAM6. J. of Adv. in Modeling Earth Systems, 5:146–172, June 2013. doi: 10.1002/jame.20015. J. W. Stock, C. S. Boxe, R. Lehmann, J. L. Grenfell, A. B. C. Patzer, H. Rauer, and Y. L. Yung. Chemical pathway analysis of the Martian atmosphere: CO2-formation pathways. Icarus, 219:13–24, May 2012a. doi: 10.1016/j.icarus.2012.02.010. J. W. Stock, J. L. Grenfell, R. Lehmann, A. B. C. Patzer, and H. Rauer. Chemical path- way analysis of the lower Martian atmosphere: The CO2 stability problem. Planet. Space Science, 68:18–24, August 2012b. doi: 10.1016/j.pss.2011.03.002. F. Tian, M. W. Claire, J. D. Haqq-Misra, M. Smith, D. C. Crisp, D. Catling, K. Zahnle, and J. F. Kasting. Photochemical and climate consequences of sulfur outgassing on early Mars. Earth Plan. Science Letters, 295:412–418, July 2010. doi: 10.1016/j. epsl.2010.04.016. F. Tian, K. France, J. L. Linsky, P. J. D. Mauas, and M. C. Vieytes. High stellar FUV/NUV ratio and oxygen contents in the atmospheres of potentially habitable planets. Earth Plan. Science Letters, 385:22–27, January 2014. doi: 10.1016/j.epsl. 2013.10.024. O. B. Toon, C. P. McKay, T. P. Ackerman, and K. Santhanam. Rapid calculation of radiative heating rates and photodissociation rates in inhomogeneous multiple scattering atmospheres. J. Geophys. Res., 94:16287–16301, 1989. P. von Paris, H. Rauer, J. L. Grenfell, B. Patzer, P. Hedelt, B. Stracke, T. Trautmann, and F. Schreier. Warming the early Earth - CO2 reconsidered. Planet. Space Science, 56:1244–1259, October 2008. doi: 10.1016/j.pss.2008.04.008. P. von Paris, S. Gebauer, M. Godolt, J. L. Grenfell, P. Hedelt, D. Kitzmann, A. B. C. Patzer, H. Rauer, and B. Stracke. The extrasolar planet GL 581 d: A potentially habitable planet? Astron. Astrophys., 522:A23, November 2010. doi: 10.1051/ 0004-6361/201015329. P. von Paris, J. L. Grenfell, P. Hedelt, H. Rauer, F. Selsis, and B. Stracke. Atmospheric constraints for the CO2 partial pressure on terrestrial planets near the outer edge of the habitable zone. Astron. Astrophys., 549:A94, January 2013a. doi: 10.1051/ 0004-6361/201016058. P. von Paris, J. L. Grenfell, H. Rauer, and J. W. Stock. N2-associated surface warming on early Mars. Planet. Space Science, 82:149–154, July 2013b. doi: 10.1016/j.pss. 2013.04.009. 28 P. von Paris, P. Hedelt, F. Selsis, F. Schreier, and T. Trautmann. Characterization of potentially habitable planets: Retrieval of atmospheric and planetary proper- ties from emission spectra. Astron. Astrophys., 551:A120, March 2013c. doi: 10.1051/0004-6361/201220009. P. von Paris, F. Selsis, D. Kitzmann, and H. Rauer. The dependence of the ice-albedo feedback on atmospheric properties. Astrobiology, 13:899–909, October 2013d. doi: 10.1089/ast.2013.0993. R. Wordsworth and R. Pierrehumbert. Abiotic Oxygen-dominated Atmospheres on Terrestrial Habitable Zone Planets. Astrophys. J. Letters, 785:L20, April 2014. doi: 10.1088/2041-8205/785/2/L20. R. Wordsworth, F. Forget, and V. Eymet. Infrared collision-induced and far-line ab- Icarus, 210:992–997, December 2010. doi: sorption in dense CO2 atmospheres. 10.1016/j.icarus.2010.06.010. R. Wordsworth, F. Forget, E. Millour, J. Head, J.-B. Madeleine, and B. Charnay. Global modelling of the early Martian climate under a denser CO2 atmosphere: Water cycle and ice evolution. Icarus, 222(1):1–19, January 2013. doi: 10.1016/j.icarus.2012. 09.036. Y. L. Yung and W. B. deMore. Photochemistry of Planetary Atmospheres. Oxford University Press, 1999. Y. L. Yung, H. Nair, and M. F. Gerstell. NOTE: CO2 Greenhouse in the Early Martian Icarus, 130:222–224, November 1997. Atmosphere: SO2 Inhibits Condensation. doi: 10.1006/icar.1997.5808. K. Zahnle. Play it again, SAM. Science, 347:370–371, January 2015. doi: 10.1126/ science.aaa3687. 29
1908.00917
1
1908
2019-07-31T23:00:58
In situ Exploration of the Giant Planets
[ "astro-ph.EP" ]
Remote sensing observations suffer significant limitations when used to study the bulk atmospheric composition of the giant planets of our solar system. This impacts our knowledge of the formation of these planets and the physics of their atmospheres. A remarkable example of the superiority of in situ probe measurements was illustrated by the exploration of Jupiter, where key measurements such as the determination of the noble gases' abundances and the precise measurement of the helium mixing ratio were only made available through in situ measurements by the Galileo probe. Here we describe the main scientific goals to be addressed by the future in situ exploration of Saturn, Uranus, and Neptune, placing the Galileo probe exploration of Jupiter in a broader context. An atmospheric entry probe targeting the 10-bar level would yield insight into two broad themes: i) the formation history of the giant planets and that of the Solar System, and ii) the processes at play in planetary atmospheres. The probe would descend under parachute to measure composition, structure, and dynamics, with data returned to Earth using a Carrier Relay Spacecraft as a relay station. An atmospheric probe could represent a significant ESA contribution to a future NASA New Frontiers or flagship mission to be launched toward Saturn, Uranus, and/or Neptune.
astro-ph.EP
astro-ph
In Situ Exploration of the Giant PlanetsA White Paper Submitted to ESA's Voyage 2050 CallContact Person: Olivier Mousis Aix Marseille Université, CNRS, LAM, Marseille, France ([email protected]) July 31, 2019 WHITE PAPER RESPONSE TO ESA CALL FOR VOYAGE 2050 SCIENCE THEME In Situ Exploration of the Giant Planets Abstract Remote sensing observations suffer significant limitations when used to study the bulk atmospheric composition of the giant planets of our solar system. This impacts our knowledge of the formation of these planets and the physics of their atmospheres. A remarkable example of the superiority of in situ probe measurements was illustrated by the exploration of Jupiter, where key measurements such as the determination of the noble gases' abundances and the precise measurement of the helium mixing ratio were only made available through in situ measurements by the Galileo probe. Here we describe the main scientific goals to be addressed by the future in situ exploration of Saturn, Uranus, and Neptune, placing the Galileo probe exploration of Jupiter in a broader context. An atmospheric entry probe targeting the 10-bar level would yield insight into two broad themes: i) the formation history of the giant planets and that of the Solar System, and ii) the processes at play in planetary atmospheres. The probe would descend under parachute to measure composition, structure, and dynamics, with data returned to Earth using a Carrier Relay Spacecraft as a relay station. An atmospheric probe could represent a significant ESA contribution to a future NASA New Frontiers or flagship mission to be launched toward Saturn, Uranus, and/or Neptune. Keywords Entry Probes -- Giant Planets -- Formation -- Composition -- Atmospheres Contents 1 Context 1 1.1 Why In Situ Measurements in Giant Planets? . . . 1 1.2 Entry Probes in the Voyage 2050 Programme . . . 3 2 Science Themes 3 2.1 Elemental and Isotopic Composition as a Window on the Giant Planets Formation . . . . . . . . . . . . . . . . . 3 2.2 In Situ Studies of Giant Planet Atmospheres . . . . 9 3 Mission Configuration and Profile 13 3.1 Probe Mission Concept . . . . . . . . . . . . . . . . . . . . 13 3.2 Probe Delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.3 Atmospheric Entry Probe System Design . . . . . 15 4 Possible Probe Model Payload 16 19 5 19 6 Education and Public Outreach (EPO) 7 Summary and Perspectives 20 20 21 Acknowledgments References International Collaboration 1. Context 1.1 Why In Situ Measurements in Giant Planets? Giant planets contain most of the mass and the angular momen- tum of our planetary system and must have played a significant role in shaping its large scale architecture and evolution, in- cluding that of the smaller, inner worlds [1]. Furthermore, the formation of the giant planets affected the timing and efficiency of volatile delivery to the Earth and other terrestrial planets [2]. Therefore, understanding giant planet formation is essential for understanding the origin and evolution of the Earth and other potentially habitable environments throughout our solar system. The origin of the giant planets, their influence on planetary system architectures, and the plethora of physical and chemical processes at work within their atmospheres make them crucial destinations for future exploration. Since Jupiter and Saturn have massive envelopes essentially composed of hydrogen and helium and (possibly) a relatively small core, they are called gas giants. Uranus and Neptune also contain hydrogen and helium atmospheres but, unlike Jupiter and Sat- urn, their H2 and He mass fractions are smaller (5 -- 20%). They are called ice giants because their density is consistent with the presence of a significant fraction of ices/rocks in their in- teriors. Despite this apparent grouping into two classes of giant planets, the four giant planets likely exist on a contin- uum, each a product of the particular characteristics of their formation environment. Comparative planetology of the four giants in the solar system is therefore essential to reveal the potential formational, migrational, and evolutionary processes at work during the early evolution of the early solar nebula. As discussed below, in situ exploration of the four giants is the means to address this theme. In Situ Exploration of the Giant Planets -- 2/26 Figure 1. Enrichment factors (with respect to the protosolar value) of noble gases and heavy elements measured in Jupiter, Saturn, Uranus, and Neptune. Error bars, central values and planets share the same color codes (see [4] for references). Much of our understanding of the origin and evolution of the outer planets comes from remote sensing by necessity. However, the efficiency of this technique has limitations when used to study the bulk atmospheric composition that is cru- cial to the understanding of planetary origin, primarily due to degeneracies between the effects of temperatures, clouds and abundances on the emergent spectra, but also due to the limited vertical resolution. In addition, many of the most abun- dant elements are locked away in a condensed phase in the upper troposphere, hiding the main volatile reservoir from the reaches of remote sensing. It is only by penetrating below the "visible" weather layer that we can sample the deeper tropo- sphere where those elements are well mixed. A remarkable example of the superiority of in situ probe measurements is illustrated by the exploration of Jupiter, where key measure- ments such as the determination of the abundances of noble gases and the precise measurement of the helium mixing ratio have only been possible through in situ measurements by the Galileo probe [3]. The Galileo probe measurements provided new insights into the formation of the solar system. For instance, they revealed the unexpected enrichments of Ar, Kr and Xe with respect to their solar abundances (see Figure 1), which sug- gested that the planet accreted icy planetesimals formed at temperatures possibly below ∼50 K to enable the trapping of these noble gases. Another remarkable result was the deter- mination of the Jovian helium abundance using a dedicated instrument aboard the Galileo probe [5] with an accuracy of 2%. Such an accuracy on the He/H2 ratio is impossible to de- rive from remote sensing, irrespective of the giant planet being considered, and yet precise knowledge of this ratio is crucial for the understanding of giant planet interiors and thermal evolution. The Voyager mission has already shown that these ratios are far from being identical in the gas and icy giants, which presumably result from different thermal histories and internal processes at work. Another important result obtained by the mass spectrometer onboard the Galileo probe was the determination of the 14N/15N ratio, which suggested that nitro- gen present in Jupiter today originated from the solar nebula essentially in the form of N2 [6]. The Galileo science payload unfortunately could not probe to pressure levels deeper than 22 bar, precluding the determination of the H2O abundance at lev- els representative of the bulk oxygen enrichment of the planet. Furthermore, the probe descended into a region depleted in volatiles and gases by unusual "hot spot" meteorology [7, 8], and therefore its measurements are unlikely to represent the bulk planetary composition. Nevertheless, the Galileo probe measurements were a giant step forward in our understanding of Jupiter. However, with only a single example of a giant planet measurement, one must wonder to what extent from the measured pattern of elemental and isotopic enrichments, the chemical inventory and formation processes at work in our solar system are truly understood. In situ exploration of giant planets is the only way to firmly characterize their composition. 0.1 1 10 100abundance ratiosHeNeArKrNOSPPlanetary to protosolar elementalNeptuneUranusSaturnJupiterInteriorprocessesHot spotCXe In this context, one or several entry probes sent to the atmo- sphere of any of the other giant planets of our solar system is the next natural step beyond Galileo's in situ exploration of Jupiter, the remote investigation of its interior and gravity field by the Juno mission, and the Cassini spacecraft's orbital reconnaissance of Saturn. In situ exploration of Saturn, Uranus or Neptune's atmo- spheres addresses two broad themes. First, the formation history of our solar system and second, the processes at play in planetary atmospheres. Both of these themes are discussed throughout this White Paper. Both themes have relevance far beyond the leap in understanding gained about an individual giant planet: the stochastic and positional variances produced within the solar nebula, the depth of the zonal winds, the prop- agation of atmospheric waves, the formation of clouds and hazes and disequilibrium processes of photochemistry and ver- tical mixing are common to all planetary atmospheres, from terrestrial planets to gas and ice giants and from brown dwarfs to hot exoplanets. 1.2 Entry Probes in the Voyage 2050 Programme The in situ exploration of Saturn, Uranus, and/or Neptune fits perfectly within the ambitious scope of the ESA Voyage 2050 Programme. A Saturn entry probe proposal has already been submitted to the ESA M4 and M5 calls in 2015 and 2016, re- spectively. Experience from these submissions shows that the development of entry probes match well the envelope allocated to ESA M-class missions provided that the carrier is provided by another space agency. Selection for phase A failed during the M4 and M5 evaluations because of the lack of availability of a NASA carrier at the envisaged launch epoch. An ideal combination would be a partnership between ESA and NASA in which ESA provides an entry probe as an important element of a more encompassing NASA New Frontiers or Flagship mission toward Saturn, Uranus, or Neptune. A joint NASA- ESA Ice Giant Study Science Definition Team (SDT) has been set in 2016-2017 to investigate the best mission scenarios ded- icated to the exploration of Uranus and Neptune in terms of science return [9]. The conclusions of the study outline the high priority of sending an orbiter and atmospheric probe to at least one of the ice giants. The mission architectures assessed by the 2017 NASA SDT showed that 2030 -- 34 were the opti- mal launch windows for Uranus, but it would be even earlier (2029 -- 30) for Neptune, depending on the use of Jupiter for a gravity assist. An internal ESA study led at the end of 2018 (ESA M* Ice Giant CDF study 1) shows that the technology is available in Europe to provide a probe to NASA1 in the framework of a joint mission. Apart the DragonFly mission 1http://sci.esa.int/future-missions-department/61307-cdf-study-report- ice-giants/ In Situ Exploration of the Giant Planets -- 3/26 dedicated to the exploration of Titan and recently selected by NASA for launch in 2026, future New Frontiers proposals could also be devoted to the in situ exploration of Saturn [10]. The selection of such proposals could create an ideal context for ESA to contribute an entry probe to NASA. Under those circumstances, the dropping of one or several probes could be envisaged in the atmosphere of Saturn. 2. Science Themes 2.1 Elemental and Isotopic Composition as a Win- dow on the Giant Planets Formation The giant planets in the solar system formed 4.55 Gyr ago from the same material that engendered the Sun and the entire solar system. Protoplanetary disks, composed of gas and dust, are almost ubiquitous when stars form, but their typical lifetimes do not exceed a few million years. This implies that the gas giants Jupiter and Saturn had to form rapidly to capture their hydrogen and helium envelopes, more rapidly than the tens of millions of years needed for terrestrial planets to reach their present masses [11, 12, 13]. Due to formation at fairly large radial distances from the Sun, where the solid surface density is low, the ice giants Uranus and Neptune had longer formation timescales (slow growth rates) and did not manage to capture large amounts of hydrogen and helium before the disk gas dissipated [14, 15]. As a result, the masses of their gaseous envelopes are small compared to their ice/rock cores. A comparative study of the properties of these giant planets thus gives information on spatial gradients in the physical and chemical properties of the solar nebula as well as on stochastic effects that led to the formation of the solar system. Data on the composition and structure of the giant planets, which hold more than 95% of the mass of the solar system outside of the Sun, remain scarce, despite the importance of such knowledge. The formation of giant planets is now largely thought to have taken place via the core accretion model in which a dense core is first formed by accretion and the hydrogen-helium envelope is captured after a critical mass is reached [16, 11]. When the possibility of planet migration is included [17, 18], such a model may be able to explain the orbital properties of exoplanets, although lots of unresolved issues remain [19, 20]. An alternative giant planets formation scenario is also the gravitational instability model [21, 22], in which the giant planets form from the direct contraction of a gas clump resulting from local gravitational instability in the disk. In the following, we briefly review the interior models, as well as the chemical and isotopic compositions of the four giants of our solar system. We also investigate the enrichment patterns that could be derived from in situ measurements by entry probes in the giant planets atmospheres to derive hints on their formation conditions. We finally summarize the key observables accessible to an atmospheric probe to address the scientific issues to the formation and evolution of the giant planets. Interior Models Interior models for the present state of the planets serve as a link between the formation scenarios out- lined above and observations. Notably, recent interior models of Jupiter that fit the gravity data observed by NASA's cur- rent Juno spacecraft are consistent with a deep interior that is highly enriched in heavy elements up to about 60% of the planet's radius. Comparison of such interior models to mod- els of Jupiter's formation and evolution implies that the deep interior still retains a memory of the infall of planetesimals at the time of formation [23]. In that scenario, accretion of heavy elements into the growing envelope led to persistent compo- sitional gradients that are still inhibiting efficient convection and mixing. However, Jupiter interior models greatly differ in the predicted amount of heavy elements in the atmosphere, which is accessible to observations. Predictions range from less than 1 × solar [24] over 1 -- 2 × solar [25] to ∼6 × solar [23]. These differences are mostly due to uncertainties in the H/He Equation of State (EOS) and can be compared with the atmospheric abundances of elements measured in giant plan- ets atmospheres provided they are representative of the bulk envelope. Such comparisons are highly valuable for constrain- ing formation models and for a better understanding of the interplay between the H/He EOS and the structure of gaseous planets. In the case of Jupiter, at minimum the heavy noble gas abundances measured by the Galileo probe serve that purpose. NASA's Juno mission currently tries to obtain the H2O abun- dance. However, the microwave spectra are highly influenced by the NH3 abundances rendering the quantitative assessment through remote sensing difficult. Bulk heavy element masses in Jupiter are estimated to range from ∼25 M⊕[24] to over ∼32 M⊕[25] up to 40 M⊕[23]. In the case of Saturn, the mass of heavy elements can vary between 0 and ∼7 M⊕ in the envelope, and between 5 and 20 M⊕ in the core [26]. Similar to Jupiter, potential compo- sitional inhomogeneities in Saturn could be the outcome of the formation process [11] and/or the erosion of a primordial core that could mix with the surrounding metallic hydrogen [27, 28]. In addition, it is possible that double diffusive con- vection occurs in the interior of Saturn [29, 30]. If a molecular weight gradient is maintained throughout the planetary enve- lope, double-diffusive convection would take place, and the thermal structure would be very different from the one that is generally assumed using adiabatic models, with much higher center temperatures and a larger fraction of heavy elements. In this case, the planetary composition can vary substantially with depth and therefore, a measured composition of the envelope In Situ Exploration of the Giant Planets -- 4/26 would not represent the overall composition. While standard interior models of Saturn assumed three layers and similar constraints in terms of the helium to hydrogen ratio, they can differ in the assumption on the distribution of heavy elements within the planetary envelope: homogeneous distribution of heavy elements apart from helium, which is depleted in the outer envelope due to helium rain [31, 26] or interior structure models allowing the abundance of heavy elements to be dis- continuous between the molecular and the metallic envelope [32, 33]. At present, it is not clear whether there should be a discontinuity in the composition of heavy elements, and this question remains open. Because of the scarcity of data, the interiors of Uranus and Neptune are even less constrained. Improved gravity field data derived from long-term observations of the planets' satellite motions suggests however that Uranus and Neptune could have different distributions of heavy elements [33]. These authors estimate that the bulk masses of heavy elements are ∼12.5 M⊕ for Uranus and ∼14 M⊕ for Neptune. They also find that Uranus would have an outer envelope with a few times the solar metallicity which transitions to a heavily enriched (∼90% of the mass in heavy elements) inner envelope at 0.9 planet's radius. In the case of Neptune, this transition is found to occur deeper inside at 0.6 planet's radius and accompanied with a more moderate increase in metallicity. Direct access to heavy materials within giant planet cores to constrain these models is impossible, so we must use the composition of the well-mixed troposphere to infer the properties of the deep interiors. It is difficult for remote sounding to provide the necessary information because of a lack of sensitivity to the atmospheric compositions beneath the cloudy, turbulent and chaotic weather layer. These questions must be addressed by in situ exploration, even if the NASA Juno mission is successful in addressing some of them remotely at Jupiter. Giant Planets Composition The abundances of most sig- nificant volatiles measured at Jupiter, Saturn, Uranus, and Neptune are summarized in Tables 1 and 2. The composition of giant planets is diagnostic of their formation and evolution history. Measuring their heavy element, noble gas, and isotope abundances reveals the physico-chemical conditions and pro- cesses that led to formation of the planetesimals that eventually fed the forming planets [3, 34, 35]. Heavy element abundances can be derived through a variety of remote sensing techniques such as spectroscopy. However, the most significant step for- ward regarding our knowledge of giant planet internal composi- tion was achieved with the in situ descent of the Galileo probe into the atmosphere of Jupiter [36, 37, 38, 39, 40, 41, 5]. The various experiments enabled the determination of the He/H2 ra- tio with a relative accuracy of 2% [5], of several heavy element abundances and of noble gases abundances [41, 42, 8]. These measurements have paved the way to a better understanding of Jupiter's formation and evolution. For example, neon in Jupiter's atmospheres has been found to be the most strongly depleted element. Its depletion, in contrast to the measured enrichments in Ar, Kr, Xe, is attributed to the helium rain in Jupiter [43]. It would be very valuable to have measurements of the heavy noble gases in any other giant planet. For Saturn, we would expect a similarly strong depletion in neon as in Jupiter as a result of deep atmospheric helium rain whereas in Uranus and Neptune depletion in He and Ne is not expected. This is because their deep interiors are mostly made of ices, implying that He is rare there and does not rain out. In situ measurements in all of these planets atmospheres would thus allow us to test these assumptions and to offer a diagnostic tool of the behavior of H/He at high pressures in giant planets. The uniform enrichment observed in the Galileo probe data (see Figure 1) tends to favor a core accretion scenario for Jupiter (e.g. [12, 44]), even if the gravitational capture of planetesi- mals by the proto-Jupiter formed via gravitational instability may also explain the observed enrichments [45]. On the other hand, the condensation processes that formed the protoplane- tary ices remain uncertain, because the Galileo probe failed to measure the deep abundance of oxygen by diving into a dry area of Jupiter [46]. Achieving this measurement by means of remote radio observations is one of the key and most chal- lenging goals of the Juno mission [47, 48], currently in orbit around Jupiter. At Saturn, the data on composition are scarcer (see Figure 1) and have mostly resulted from Voyager 2 mea- surements and intense observation campaigns with the Cassini orbiter. The He abundance is highly uncertain [49, 50, 51], and only the abundances of N, C, and P, have been quantified [52, 53, 54, 55, 56]. This scarcity of essential data is the main motivation for sending an atmospheric probe to Saturn and was the core of several mission proposals submitted to ESA and NASA calls over the last decade [57, 58, 59]. Uranus and Neptune are the most distant planets in our Solar System. Their apparent size in the sky is roughly a factor of 10 smaller than Jupiter and Saturn, which makes telescopic observations from Earth much more challenging in terms of detectability. This distance factor is probably also the reason why space agencies have not yet sent any new flyby or orbiter mission to either of these planets since Voyager 2. As a consequence, the knowledge of their bulk composition is dramatically poor (see Figure 1), resulting in a very limited understanding of their formation and evolution. Improving this situation needs ground-truth measurements that can only be carried out in these distant planets by an atmospheric probe, similarly to the Galileo probe at Jupiter. Isotopic Measurements Table 3 represents the isotopic ra- tio measurements realized in the atmospheres of the four giant In Situ Exploration of the Giant Planets -- 5/26 planets of our solar system. The case of D/H is interesting and would deserve further measurements with smaller errors. Because deuterium is destroyed in stellar interiors and trans- formed into 3He, the D/H value presently measured in Jupiter's atmosphere is estimated to be larger by some 5 -- 10% than the protosolar value. This slight enrichment would have resulted from a mixing of nebular gas with deuterium-rich ices dur- ing the planet's formation. For Saturn, the contribution of deuterium-rich ices in the present D/H ratio could be higher (25 -- 40%). The deuterium enrichment as measured by [73] in Uranus and Neptune has been found to be very similar between the two planets, and its supersolar value also sug- gests that significant mixing occurred between the protosolar H2 and the H2O ice accreted by the planets. Assuming that the D/H ratio in H2O ice accreted by Uranus and Neptune is cometary (1.5 -- 3 ×10−4), [73] found that 68 -- 86% of the heavy component consists of rock and 14 -- 32% is made of ice, values suggesting that both planets are more rocky than icy, assuming that the planets have been fully mixed. Alternatively, based on these observations, [74] suggested that, if Uranus and Neptune formed at the carbon monoxide line in the protosolar nebula (PSN), then the heavy elements accreted by the two planets would mostly consists of a mixture of CO and H2O ices, with CO being by far the dominant species. This scenario assumes that the accreted H2O ice presents a cometary D/H and allows the two planets to remain ice-rich and O-rich while providing D/H ratios consistent with the observations. Deeper sounding of Saturn, Uranus, and Neptune's atmospheres with an atmospheric probe, should allow investigating the possi- bility of isotopic fractionation with depth. The measurement of the D/H ratio in Saturn, Uranus and Neptune should be complemented by a precise determination of 3He/4He in their atmospheres to provide further constraints on the protosolar D/H ratio, which remains relatively uncertain. The protoso- lar D/H ratio is derived from 3He/4He measurements in the solar wind corrected for changes that occurred in the solar corona and chromosphere subsequent to the Sun's evolution, and to which the primordial 3He/4He is subtracted [75]. This latter value is currently derived from the ratio observed in meteorites or in Jupiter's atmosphere. The measurement of 3He/4He in Uranus and/or Neptune atmospheres would there- fore complement the Jupiter value and the scientific impact of the protosolar D/H derivation. The 14N/15N ratio presents large variations in the different planetary bodies in which it has been measured and, conse- quently, remains difficult to interpret. The analysis of Genesis solar wind samples [76] suggests a 14N/15N ratio of 441 ± 5, which agrees with the remote sensing [77] and in situ [8] measurements made in Jupiter's atmospheric ammonia, and the lower limit derived from ground-based mid-infrared ob- In Situ Exploration of the Giant Planets -- 6/26 Table 1. Elemental abundances in Jupiter, Saturn, Uranus and Neptune, as derived from upper tropospheric composition Elements He/H (1) Ne/H(2) Ar/H(3) Kr/H(4) Xe/H (5) C/H(6) N/H(7) O/H(8) S/H(9) P/H(10) Jupiter (7.85± 0.16)× 10−2 (1.240± 0.014)× 10−5 (9.10± 1.80)× 10−6 (4.65± 0.85)× 10−9 (4.45± 0.85)× 10−10 (1.19± 0.29)× 10−3 (3.32± 1.27)× 10−4 (2.45± 0.80)× 10−4 (4.45± 1.05)× 10−5 (1.08± 0.06)× 10−6 Saturn (6.75± 1.25)× 10−2 -- -- -- -- (2.65± 0.10)× 10−3 (0.50− 2.85)× 10−4 -- -- (3.64± 0.24)× 10−6 Uranus (8.88± 2.00)× 10−2 -- -- -- -- (0.6− 3.2)× 10−2 -- -- (5− 12.5)× 10−6 -- Neptune (8.96± 1.46)× 10−2 -- -- -- -- (0.6− 3.2)× 10−2 -- -- (2.0− 6.5)× 10−6 -- (1) [5, 41] for Jupiter, [50, 59] for Saturn, [60] for Uranus and [61] for Neptune. We only consider the higher value of the uncertainty on He in the case of Neptune. (2−5) [62] for Jupiter. (6) [8] for Jupiter, [55] for Saturn, [63, 64, 65, 66] for Uranus, [67, 64, 68] for Neptune. (7) [8] for Jupiter, [69] for Saturn (N/H range derived from the observed range of 90 -- 500 ppm of NH3). (8) [8] for Jupiter (probably a lower limit, not representative of the bulk O/H). (9) [8] for Jupiter, lower limits for Uranus [70] and Neptune [71]. (10) [56] for Jupiter and Saturn. Table 2. Ratios to protosolar values in the upper tropospheres of Jupiter, Saturn, Uranus and Neptune Elements He/H Ne/H Ar/H Kr/H Xe/H C/H N/H O/H S/H P/H Jupiter/Protosolar 0.81± 0.05 0.10± 0.03 2.55± 0.83 2.16± 0.59 2.12± 0.59 4.27± 1.13 4.06± 2.02 0.40± 0.15 (hotspot) 2.73± 0.65 3.30± 0.37 Saturn/Protosolar Uranus/Protosolar Neptune/Protosolar 0.70± 0.14 -- -- -- -- 9.61± 0.59 0.61 -- 3.48 -- -- 11.17± 1.31 0.93± 0.16 -- -- -- -- ∼20 -- 120 -- -- 0.13 - 0.42 -- 0.93± 0.21 -- -- -- -- ∼20 -- 120 -- -- 0.32 - 0.80 -- Error is defined as (∆E/E)2 = (∆X/Xplanet)2 + (∆X/XProtosun)2. The ratios only refer to the levels where abundance measurements have been performed, i.e. in the upper tropospheres and are not automatically representative of deep interior enrichments. This is especially true if the deep interior contain a significant fraction of another element (e.g. oxygen in Uranus and Neptune, according to models). Moreover, the helium value was computed for pure H2/He mixtures (i.e. the upper tropospheric CH4 has not been accounted for), because CH4 is condensed at 1 bar where He is measured. Protosolar abundances are taken from [72]. servations of Saturn's ammonia absorption features [78]. The two 14N/15N measurements made in Jupiter and Saturn sug- gest that primordial N2 was probably the main reservoir of the present NH3 in their atmospheres [6, 57, 79]. On the other hand, Uranus and Neptune are mostly made of solids (rocks and ices) [44] that may share the same composition as comets. N2/CO has been found strongly depleted in comet 67P/Churyumov-Gerasimenko [80], i.e. by a factor of ∼25.4 compared to the value derived from protosolar N and C abun- dances. This confirms the fact that N2 is a minor nitrogen reservoir compared to NH3 and HCN in this body [81], and probably also in other comets [82]. In addition, 14N/15N has been measured to be 127 ± 32 and 148 ± 6 in cometary NH3 and HCN respectively [83, 84]. Assuming that Uranus and Neptune have been accreted from the same building blocks as those of comets, then one may expect a 14N/15N ratio in these two planets close to cometary values, and thus quite different from the Jupiter and Saturn values. Measuring 14N/15N in the atmospheres of Uranus and Neptune would provide insights about the origin of the primordial nitrogen reservoir in these planets. Moreover, measuring this ratio in different species would enable us to constrain the relative importance of the chemistry induced by galactic cosmic rays and magnetospheric electrons (see [87] for an example in Titan). The isotopic measurements of carbon, oxygen and noble gas (Ne, Ar, Kr, and Xe) isotopic ratios should be representa- tive of their primordial values. For instance, only little vari- ations are observed for the 12C/13C ratio in the solar system irrespective of the body and molecule in which it has been measured. Table 3 shows that both ratios measured in the atmospheres of Jupiter and Saturn are consistent with the ter- restrial value of 89. A new in situ measurement of this ratio in Uranus and/or Neptune should be useful to confirm whether their carbon isotopic ratio is also telluric. The oxygen isotopic ratios also constitute interesting mea- surements to be made in Uranus' and Neptune's atmospheres. The terrestrial 16O/18O and 16O/17O isotopic ratios are 499 and 2632, respectively [88]. At the high accuracy levels achievable with meteoritic analysis, these ratios present some small vari- ations (expressed in δ units, which are deviations in part per thousand). Measurements performed in comets [89], far less accurate, match the terrestrial 16O/18O value. The 16O/18O ratio has been found to be ∼380 in Titan's atmosphere from Herschel SPIRE observations but this value may be due to some fractionation process [90, 91]. On the other hand, [92] found values consistent with the terrestrial ratios in CO with ALMA. The only 16O/18O measurement made so far in a giant planet was obtained from ground-based infrared observations in Jupiter's atmosphere and had a too large uncertainty to be interpreted in terms of 1 -- 3 times the terrestrial value [93]. Formation Models and Enrichment Patterns in Giant Plan- ets Direct or indirect measurements of the volatile abun- dances in the atmospheres of Saturn, Uranus and Neptune are key for deciphering their formation conditions in the PSN. In what follows, we present the various models and their predic- tions regarding enrichments in the giants. Figure 2 summarizes the predictions of the various models in the cases of Uranus and Neptune. • Gravitational Instability Model. This formation scenario is associated with the photoevaporation of the giant plan- ets envelopes by a nearby OB star and settling of dust grains prior to mass loss [94]. It implies that O, C, N, S, Ar, Kr and Xe elements should all be enriched by a similar factor relative to their protosolar abundances in the envelopes, assuming mixing is efficient. Despite the fact that interior models predict that a metallicity gra- dient may increase the volatile enrichments at growing depth in the planet envelopes [33], there is no identified process that may affect their relative abundances in the ice giant envelopes, if the sampling is made at depths below the condensation layers of the concerned volatiles and if thermochemical equilibrium effects are properly taken into account. The assumption of homogeneous enrichments for O, C, N, S, Ar, Kr and Xe, relative to their protosolar abundances, then remains the natural outcome of the formation scenario proposed by [94]. In Situ Exploration of the Giant Planets -- 7/26 • Core Accretion and Amorphous Ice. In the case of the core accretion model, because the trapping efficiencies of C, N, S, Ar, Kr and Xe volatiles are similar at low temperature in amorphous ice [3, 95], the delivery of such solids to the growing giant planets is also consis- tent with the prediction of homogeneous enrichments in volatiles relative to their protosolar abundances in the envelopes, still under the assumption that there is no process leading to some relative fractionation between the different volatiles. • Core Accretion and Clathrates In the core accretion model, if the volatiles were incorporated in clathrate structures in the PSN, then their propensities for such trapping would strongly vary from a species to another. For instance, Xe, CH4 and CO2 are easier clathrate for- mers than Ar or N2 because their trapping temperatures are higher at PSN conditions, assuming protosolar abun- dances for all elements [96]. This competition for trap- ping is crucial when the budget of available crystalline water is limited and does prevent the full clathration of the volatiles present in the PSN [34, 97, 79]. However, if the O abundance is 2.6 times protosolar or higher at the formation locations of Uranus and Neptune's building blocks and their formation temperature does not exceed ∼45K, then the abundance of crystalline water should be high enough to fully trap all the main C, N, S and P -- bearing molecules, as well as Ar, Kr and Xe [79]. In this case, all elements should present enrichments comparable to the C measurement, except for O and Ar, based on calculations of planetesimals compositions performed under those conditions [79]. The O enrich- ment should be at least ∼4 times higher than the one measured for C in the envelopes of the ice giants due to its overabundance in the PSN. In contrast, the Ar en- richment is decreased by a factor of ∼4.5 compared to C, due to its very poor trapping at 45 K in the PSN (see Figure 2). We refer the reader to [79] for further details about the calculations of these relative abundances. • Photoevaporation Model. An alternative scenario is built upon the ideas that (i) Ar, Kr and Xe were homo- geneously adsorbed at very low temperatures (∼20 -- 30 K) at the surface of amorphous icy grains settling in the cold outer part of the PSN midplane [98] and that (ii) the disk experienced some chemical evolution in the giant planets formation region (loss of H2 and He), due to pho- toevaporation. In this scenario, these icy grains migrated toward the formation region of the giant planet where they subsequently released their trapped noble gases, due to increasing temperature. Due to the disk's pho- In Situ Exploration of the Giant Planets -- 8/26 Table 3. Isotopic ratios measured in Jupiter, Saturn, Uranus and Neptune Isotopic ratio D/H (in H2)(1) 3He/4He(2) 12C/13C (in CH4)(3) 14N/15N (in NH3)(4) 20Ne/22Ne(5) 36Ar/38Ar(6) 136Xe/total Xe(7) 134Xe/total Xe(8) 132Xe/total Xe(9) 131Xe/total Xe(10) 130Xe/total Xe(11) 129Xe/total Xe(12) 128Xe/total Xe(13) Jupiter Saturn (2.60 ± 0.7) × 10−5 (1.66 ± 0.05) × 10−4 92.6+4.5−4.1 434.8+65−50 13 ± 2 5.6 ± 0.25 0.076 ± 0.009 0.091 ± 0.007 0.290 ± 0.020 0.203 ± 0.018 0.038 ± 0.005 0.285 ± 0.021 0.018 ± 0.002 1.70+0.75−0.45 × 10−5 91.8+8.4−7.8 > 357 -- -- -- -- -- -- -- -- -- -- Uranus Neptune (4.4 ± 0.4) × 10−5 (4.1 ± 0.4) × 10−5 -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- -- (1) [85] for Jupiter, [86] for Saturn, [73] for Uranus and Neptune. (2) [85] for Jupiter. (3) [41] for Jupiter, [55] for Saturn. (4) [8] for Jupiter, [78] for Saturn. (5−13) [62] for Jupiter. toevaporation inducing fractionation between H2, He and the other heavier species, these noble gases would have been supplied in supersolar proportions from the PSN gas to the forming giant planets. The other species, whose trapping/condensation temperatures are higher, would have been delivered to the envelopes of the gi- ants in the form of amorphous ice or clathrates. [98] predict that, while supersolar, the noble gas enrichments should be more moderate than those resulting from the accretion of solids containing O, C, N, S by the two giants. • CO Snowline Model Another scenario, proposed by [74], suggests that Uranus and Neptune were both formed at the location of the CO snowline in a stationary disk. Due to the diffusive redistribution of vapors (the so- called cold finger effect; [99, 100]), this location of the PSN intrinsically had enough surface density to form both planets from carbon -- and oxygen -- rich solids but nitrogen-depleted gas. The analysis has not been ex- tended to the other volatiles but this scenario predicts that species whose snowlines are beyond that of CO remain in the gas phase and are significantly depleted in the envelope compared to carbon. Under those circum- stances, one should expect that Ar presents the same depletion pattern as for N in the atmospheres of Uranus and Neptune. In contrast, Kr, Xe, S and P should be found supersolar in the envelopes of the two ice giants, but to a lower extent compared to the C and O abun- dances, which are similarly very high [74]. Summary of Key Measurements Here we list the key mea- surements to be performed by an atmospheric entry probe at Saturn, Uranus and Neptune to better constrain their formation and evolution scenarios: • Temperature -- pressure profile from the stratosphere down to at least 10 bars. This would establish the stability of the atmosphere towards vertical motions and constrain the opacity properties of clouds lying at or above these levels (CH4 and NH3 or H2S clouds). At certain pres- sures convection may be inhibited by the mean molec- ular weight gradient [101] (for instance at ∼2 bar in Neptune) and it is thus important to measure the temper- ature gradient in this region. Probing deeper than ∼40 bars would be needed to assess the bulk abundances of N and S existing in the form of NH4SH but this would re- quire microwave measurements from a Juno-like orbiter, instead of using a shallow probe. • Tropospheric abundances of C, N, S, and P, down to the 10-bar level at least, with accuracies of ±10% (of the or- der of the protosolar abundance accuracies). In the case of the ice giants, N and S could be measured remotely deeper to the 40-bar level at microwave wavelengths by a Juno-like orbiter. • Tropospheric abundances of noble gases He, Ne, Xe, Kr, Ar, and their isotopes to trace materials in the subreser- voirs of the PSN. The accuracy on He should be at least as good that obtained by Galileo at Jupiter (±2%), and the accuracy on isotopic ratios should be ±1% to en- able direct comparison with other known Solar System values. In Situ Exploration of the Giant Planets -- 9/26 Figure 2. Qualitative differences between the enrichments in volatiles predicted in Uranus and Neptune predicted by the different formation scenarios (calibrations based on the carbon determination). The resulting enrichments for the different volatiles are shown in green (gravitational instability model and amorphous ice), orange (clathrates), blue (photoevaporation) and red (CO snowline). • Isotopic ratios in hydrogen (D/H) and nitrogen (15N/14N), with accuracies of ±5%, and in oxygen (17O/16O and 18O/16O) and carbon (13C/12C) with accuracies of ±1%. This will enable us to determine the main reservoirs of these species in the PSN. • Tropospheric abundances of CO and PH3. Having both values brackets the deep H2O abundance [102]. CO alone may not be sufficient to enable the evaluation of the deep H2O because of the uncertainties on the deep thermal profile (convection inhibition possible at the H2O condensation level) as shown in [103]. 2.2 In Situ Studies of Giant Planet Atmospheres The giant planets are natural planetary-scale laboratories for the study of fluid dynamics without the complex effects of topography and ocean -- atmosphere coupling. Remote sensing provides access to a limited range of altitudes, typically from the tropospheric clouds upwards to the lower stratosphere and thermosphere, although microwave radiation can probe deeper below the upper cloud deck. The vertical resolution of "nadir" remote sensing is limited to the width of the contribution func- tion (i.e., the range of altitudes contributing to the upwelling radiance at a given wavelength), which can extend over one or more scale heights and makes it impossible to uniquely iden- tify the temperature and density perturbations associated with cloud formation, wave phenomena, etc. In situ exploration of Saturn, Uranus or Neptune would not only constrain their bulk chemical composition, but it would also provide direct sampling and "ground-truth" for the myriad of physical and chemical processes at work in their atmospheres. In the follow- ing we explore the scientific potential for a probe investigating atmospheric dynamics, meteorology, clouds and hazes, and chemistry. We also provide the key atmospheric observables accessible to an atmospheric probe. Zonal Winds At the cloud tops, Jupiter and Saturn have multi-jet winds with eastward equatorial jets, while Uranus and Neptune have a broad retrograde equatorial jet and nearly symmetric prograde jets at high latitudes [104] (Fig. 3). The question of the origin of the jets and the differences between the gas giants and the icy giants is the subject of intensive research. Numerical attempts to study this question are based In Situ Exploration of the Giant Planets -- 10/26 Figure 3. Zonal winds in Jupiter, Saturn, Uranus and Neptune [104] from different space missions. The equator, which is home to strong vertical wind-shear, is highlighted in each panel. In Jupiter the Galileo Probe measured strong vertical wind shears confined to the first 5 bar of the atmosphere [110, 111]. In Saturn [112] and Neptune there are strong evidences of vertical wind shears at the equator [113, 114]. either on external forcing by the solar irradiation with a shallow circulation, or in deep forcing from the internal heat source of the planets producing internal columnar convection [105, 104]. However none of these models has been able to reproduce the characteristics of the wind systems of the planets without fine- tuning their multiple parameters. It is possible to explore the depth of the winds through measurements of the gravity field of the planet combined with interior models. Recent results from Juno [106, 107] and Cassini [108], and a reanalysis of Uranus and Neptune Voyager data [109] show that the winds are neither shallow, nor deep in any of these planets and may extend 3,000 km in Jupiter, 9,000 km in Saturn and 1,000 km in Uranus and Neptune. Vertical wind-shears are determined by measuring the horizontal distribution of temperature. Remote sensing can provide maps of temperature above the clouds but do not permit the determination of the deeper winds. In addition, in Uranus and Neptune, the horizontal distribution of volatiles causes humidity winds [115], an effect that occurs in hydrogen-helium atmospheres with highly enriched volatiles. In situ measurements of how the wind changes in the top few tens of bars (e.g., like Galileo) would provide insights into how the winds are being generated. The vertical wind shear measured by Galileo defied previous ideas of the expected structure of the winds. Theoretical models of atmospheric jets driven by solar heat flux and shallow atmospheric processes include a crucial role of moist convection in the troposphere [116] and only through knowledge of the vertical distribution of condensables and winds we will be able to understand the generated wind systems of these planets. Temperature Structure Vertical profiles of temperature in the upper atmospheres are retrieved from mid-infrared and sub-millimetre remote sounding. The determination of these vertical profiles from occultation measurements depends on the knowledge of the mean molecular weight, and therefore, requires simultaneous sensing of infrared radiance to con- strain the bulk composition. However, measuring the verti- cal (and horizontal) distribution of volatile gases and their condensed phases from orbit is a fundamentally degenerate problem. Hence entry probes are the only way to determine these quantities with accuracy and provide a ground-truth to the study of the temperature distribution. This is true for Sat- urn even if the very successful Cassini mission has provided unprecedented observations of the temperature structure of the planet [117]. Models of globally-averaged temperatures for Uranus [118] and Neptune [119] present differences with the radio occultation results [120, 121] and an in situ determi- nation of a thermal profile and vertical distribution of mean molecular weight is a vital measurement for the interpreta- tion of thermal data. Furthermore, available data is limited to pressures smaller than 1 bar or is intrinsically degenerate and model-dependent. A considerable uncertainty in Uranus and Neptune is due to the molecular weight gradient caused by methane condensation and the resulting inhibition of moist convection in the atmosphere [101, 122, 123], with a resulting temperature profile that may be sub-adiabatic, dry adiabatic or superadiabatic. This has consequences for interior and evo- lution models, atmospheric dynamics and the interpretation of abundances measurements in particular for disequilibrium In Situ Exploration of the Giant Planets -- 11/26 Figure 4. Multi-wavelength images of Jupiter (upper row), Saturn (middle row) and Uranus and Neptune (bottom row). Images in the near-infrared in methane absorption bands (a, g, i) sample complex layers of hazes. Visible images (b, d, f, h) correspond to the top of the main upper cloud (NH3 in Jupiter and Saturn and CH4 in Uranus and Neptune). Infrared images at 4-5 µm(c, e) sample the opacity of a secondary cloud layer, most probably NH4SH in Jupiter and Saturn. species. In situ measurements will provide ground truth. Be- cause in these planets the methane condensation region is at pressures smaller than 2 bars, this is well within reach of the probe that we consider. Also, solar irradiation alone cannot explain the high temperatures found in the stratospheres and thermospheres of Uranus and Neptune [124, 125], a problem known as the energy crisis that cannot be solved from remote sensing. Measurements of temperatures in the stratosphere would result in a detailed characterization of gravity waves propagation that could help us to resolve energy transfer pro- cesses in planetary atmospheres in general. Clouds Images of the gas and ice giants in the visible and near-infrared show a plethora of clouds that organize in zonal bands, vortices, planetary waves and turbulent regions (Fig. 4). The vertical structure of clouds from multi-wavelength observations can be interpreted via radiative-transfer models, but these models offer multiple possibilities to fit individual observations and require a good knowledge of the vertical dis- tribution of absorbing species like methane or volatile gases. The observable clouds in Jupiter and Saturn are separated in three layers (hazes close to the tropopause at 60 -- 100 mbar, high-opacity clouds with their tops at 400 -- 700 mbar and deep clouds with opacity sources at around 1.5 -- 2.0 bar). The ac- Figure 5. Vertical cloud structure in Saturn (left) and Uranus (right) from Voyager thermal profiles extended following a moist adiabat (dashed-line) and assuming 5 times solar abundance of condensable for Saturn and 30 times solar abundances for Uranus except for NH3, which is assumed to have a lower abundance than H2S. The upper atmosphere is home to several photochemical layers. The vertical distribution of molecular weight is also shown (dotted-line). Simple ECC models do not take into account precipitation of condensates and the actual cloud structure could be different. Tropospheric hazes required by radiative-transfer models to fit the observations are also shown. cessible clouds in Uranus and Neptune are different with an extended haze layer topping at 50 -- 100 mbar located above a thin methane cloud of ice condensates with its base at ∼ 1.3 bar. This cloud is above another cloud of H2S ice that is optically thick, located between 2 and 4 bar of pressure and whose structure can not be discerned from the observations. These basic vertical cloud structures come from multiple independent studies ([126, 127, 128] for Jupiter, [129, 130, 131] for Saturn, [132, 133, 134, 135] for Uranus, and [136, 137, 138] for Nep- tune), and generally assume specific properties of the clouds in different regions of the planet. However, radiative transfer models produce highly degenerate solutions where multiple possibilities for the cloud particle optical properties and verti- cal structure can be found that can fit the observations. Under those circumstances, in situ measurements provide a ground- truth to remote sensing observations. They give us information about clouds much deeper than what can be observed from remote sensing. The relation between the bands and colors in the giant planets is not well understood. The pattern of bands in Jupiter observed in the visible follows the structure of the zonal jets [139]. The same holds partially in Saturn [140], but the bands in Uranus and Neptune have a much richer structure than the wind field [141, 142, 143, 144]. In all planets changes in the bands do not seem to imply changes in the more stable wind system [104]. Questions about how the belt and circulation pattern can be established [145] may require information from atmospheric layers below the visible pattern of clouds, which are not accessible to remote sensing. Exploring deeper into the atmosphere requires the use thermochemical Equilibrium Cloud Condensation (ECC) models which predict the location of clouds based on hypothesis of the relative abundances of condensables and thermal extrapolations of the upper tem- peratures [146, 147]. Depending on the planet and relative abundances of the condensables several cloud layers are pre- dicted to form: NH3, NH4SH and H2O in Jupiter and Saturn, and CH4, H2S, NH4SH and H2O in Uranus and Neptune (Fig. 5). An additional intermediate cloud of NH3 could form at pressures around 10 bar depending of the sequestration of NH3 molecules in the lower NH4SH cloud and the amount of NH3 dissolved in the deep and massive liquid water cloud. This am- monia cloud is not expected currently in Uranus and Neptune due to the detection of tropospheric H2S gas [148, 149] that seems to indicate that H2S is more abundant than NH3 in these atmospheres. A shallow probe to 10 bar in Saturn would descend below the NH4SH cloud but may not probe the water cloud base and its deep abundance. A similar probe in Uranus and Neptune would descend below the H2S cloud, while a deep probe would be needed to reach the NH4SH cloud layer and the top of the H2O cloud, which could extend to hundreds of bars. However, the descent profile would depend on the properties of the meteorological environment of the descent [150], a question we now examine. Convection and Meteorological Features Moist convec- tion develops through the release of latent heat when gases condense and mix vertically impacting the vertical distribu- tion of volatiles, molecular weight and temperature. In the giant planets volatiles are heavier than the dry air reducing the buoyancy of convective storms and potentially inhibiting moist convection in Jupiter's deep water cloud layer for water abundances higher than 5 and in Uranus and Neptune methane and deeper clouds [101, 122]. However, convective storms are relatively common in Jupiter and group in cyclonic regions [139, 105]. In Saturn, they occur seasonally in the tropics over extended periods of time [151] and develop into Great Storms once per Saturn year [152]. Discrete cloud systems form and dissipate episodically in Uranus and Neptune including bright cloud systems that could be intense storms [153, 114]. How- ever, there is no consensus whether or not these features are events of energetic moist convection as their vertical cloud In Situ Exploration of the Giant Planets -- 12/26 structure does not result in the elevated cloud tops [154] ex- pected from comparison with Jupiter and Saturn and basic models of moist convection [155]. Large and small vortices, waves and turbulent regions are common in the atmospheres of Jupiter and Saturn [139, 156]. Neptune is famous for its dark vortices surrounded by bright companion clouds [142, 157] and Uranus has rare dark vortices [158] and bright cloud sys- tems [159]. Many of these meteorological systems last for years to decades but we ignore how deep they extend into the lower troposphere. Large-scale waves can also affect the properties of the atmosphere well below the upper cloud layer [150]. The interpretation of vertical profiles of pressure, tempera- ture, wind speed, and composition obtained by a probe would hugely benefit from an observational characterization of the descending region and its meteorology at cloud level [160]. Chemistry In the upper atmospheres of Jupiter, Saturn, Uranus and Neptune, methane is photolysed into hydrocarbons that diffuse down and condense to form haze layers in the cold stratospheres (altitudes ∼0.1 to 30 mbar) as the temperature decreases down to ∼60 K in the tropopause in Uranus and Neptune. Photochemical models suggest hazes made of hy- drocarbons that become progressively more important from Jupiter to Uranus and Neptune with C2H2, C6H6, C4H2, C4H10, CO2, C3H8, C2H2, add C2H6 [161, 162, 163], where the oxy- gen species derive from external sources such as interplane- tary dust or comets. These species are radiatively active at mid-infrared wavelengths and affect the aerosol structure and energy balance of the atmospheres and, thus, their overall dy- namics. Tropospheric CO is particularly important because it is related with other oxygen bearing molecules including water. Thermochemical models have been used to relate the observed CO abundance with the deep water abundance [103] but results of these models depend on precise measurements of tropospheric CH4 and knowledge of vertical mixing that can only be determined precisely in situ. Summary of Key Measurements Below are indicated the key in situ measurements needed to characterize the atmo- spheres of Saturn, Uranus or Neptune. • Temperature-pressure profile. This basic but essential measurement will be key to check widespread but model- dependent measurements obtained from remote observa- tions. Testing for the presence of sub- or super-adiabatic lapse rates will be key to understand how internal heat is transported in these active atmospheres. • Cloud and haze properties. A descent probe would be able to measure the atmospheric aerosols scattering prop- erties at a range of phase angles, the particles number density, the aerosol shape and opacity properties. Each of these measurements would help constrain the aerosol composition, size, shape, and density. • Winds. Doppler wind measurements provide the wind profile in the lower troposphere, well below the region where most of the cloud tracking wind measurements are obtained. Static and dynamic pressures would provide an estimate of the vertical winds, waves, and convection. The comparison with vertical profiles of condensable abundances and thermal data would quantify the relative importance of thermal and humidity winds. • Conductivity. A vertical profile of atmospheric conduc- tivity would indicate what type of clouds support charge separation to generate lightning. Conductivity measure- ments combined with meteorological and chemical data (particularly measurements of the physical properties of the aerosols themselves) would also permit extraction of the charge distribution on aerosol particles, and im- prove understanding of the role of electrical processes in cloud formation, lightning generation, and aerosol microphysics. • Determine the influence of cloud condensation or photo- chemical haze formation on the temperature lapse rate and deduce the amount of energy relinquished by this phase change in key species (CH4, NH3, H2S). • Ortho-to-para hydrogen ratio. This would constrain the degree of vertical convection through the atmosphere and the convective capability at different cloud condens- ing layers. It would also be essential to understand the vertical profile of atmospheric stability and is especially important in the cold atmospheres of Uranus and Nep- tune. 3. Mission Configuration and Profile 3.1 Probe Mission Concept The three giant planets considered in this White Paper can be targeted with a similar probe payload and architecture. Science Mission Profile To measure the atmospheric com- position, thermal and energy structure, clouds and dynamics requires in situ measurements by a probe carrying a mass spectrometer (atmospheric and cloud compositions), helium abundance detector, atmospheric structure instrument (thermal structure and atmospheric stability), nephelometer (cloud lo- cations and aerosol properties), net flux radiometer (energy structure), physical properties instrument (temperature, pres- sure and density structure, ortho-para ratio), and Doppler-wind In Situ Exploration of the Giant Planets -- 13/26 experiment (dynamics). The atmospheric probe descent tar- gets the 10-bar level located about 5 scale heights beneath the tropopause. The speed of probe descent will be affected by requirements imposed by the needed sampling periods of the instruments, particularly the mass spectrometer, as well as the effect speed has on the measurements. This is potentially an issue for composition instruments, and will affect the altitude resolution of the Doppler wind measurement. Although it is expected that the probe batteries, structure, thermal control, and telecomm will allow operations to levels well below 10 bars, a delicate balance must be found between the total sci- ence data volume requirements to achieve the high-priority mission goals, the capability of the telecomm system to trans- mit the entire science, engineering, and housekeeping data set (including entry accelerometry and pre-entry/entry calibra- tion, which must be transmitted interleaved with descent data) within the descent telecomm/operational time window, and the probe descent architecture which allows the probe to reach 10 bars, i.e. the depth at which most of the science goals can be achieved. Probe Mission Profile to Achieve Science Goals A giant planet probe designed for parachute descent to make atmo- spheric measurements of composition, structure, and dynam- ics, with data returned to Earth using an orbiting or flyby Car- rier Relay Spacecraft (CRSC) could be carried as an element of a dedicated giant planet system exploration mission. The CRSC would receive and store probe science data in real-time, then re-transmit the science and engineering data to Earth. While recording entry and descent science and engineering data returned by the probe, the CRSC would additionally make measurements of probe relay link signal strength and Doppler for descent probe radio science. Carried by the CRSC into the vicinity of the giant planet system, the probe would be configured for release, coast, entry, and atmospheric descent. For proper probe delivery to the entry interface point, the CRSC with probe attached is placed on a planetary entry tra- jectory, and is reoriented for probe targeting and release. The probe coast timer and pre-programmed probe descent science sequence are loaded prior to release from the CRSC, and fol- lowing spin-up, the probe is released for a ballistic coast to the entry point. Following probe release, a deflect maneuver is per- formed to place the CRSC on the proper overflight trajectory to receive the probe descent telemetry. Prior to arrival at the entry interface point, the probe coast timer awakens the probe for sequential power-on, warm-up, and health checks. The only instrumentation collecting data during entry would be the entry accelerometers and possibly heat shield instrumentation including ablation sensors. The end of entry is determined by the accelerometers, initiating parachute deployment, aeroshell release, and the probe atmo- spheric descent. Parachute sequence would be initiated above the tropopause by deploying a pilot parachute which pulls off the probe aft cover, thereby extracting the main descent parachute, followed by release of the probe heatshield and initiation of a transmit-only telecommunications link from the probe to the CRSC. Under the parachute, the altitude of any required descent science operation mode changes would be guided by input from the Atmospheric Structure Instrument sensors, thereby providing the opportunity to optimize the data collection for changing science objectives at different atmo- spheric depths. The probe science data collection and relay transmission strategy would be designed to ensure the entire probe science data set is successfully transmitted prior to probe reaching the targeted depth. Probe descent mission would likely end when the telecomm geometry becomes so poor that the link can no longer be main- tained due to increasing overhead atmospheric opacity, de- pletion of the batteries, or increasing and damaging thermal and/or pressure effects. The probe transmits science and engi- neering data to the CRSC where multiple copies are stored in redundant on-board memory. At the completion of the probe descent mission and once the post-descent context observa- tions have been performed, the CRSC reorients to point the High Gain Antenna towards Earth and all stored copies of the probe science and engineering data are returned to Earth. Fig- ure 6 represents a schematic view of the Galileo entry, descent and deployment sequence which could be the basis for any proposed entry probe mission. 3.2 Probe Delivery Interplanetary Trajectory Four characteristics of interplan- etary transfers from Earth to the giant planets are of primary importance: 1) the launch energy affecting the delivered mass, 2) the flight time which affects required spacecraft reliability engineering and radioisotope power systems whose output power decreases with time, 3) the V∞ of approach (VAP) to the destination planet which influences the ∆V necessary for orbit insertion and the entry speed of an entry probe delivered from approach, and 4) the declination of the approach (DAP) asymptote which influences both the locations available to an entry probe and the probe's atmosphere-relative entry speed which depends on the alignment of the entry velocity vector with the local planetary rotation velocity. Depending on trans- fer design and mass, trajectories to the giant planets can be order of 5 -- 6 years for Jupiter, up to 10 -- 13 years for Uranus and Neptune. When Jupiter and Saturn align to provide gravity assists from both, trajectories with shorter transfer durations are possible. Probe Delivery and Options for Probe Entry Location Gi- ven a transfer trajectory defined by its VAP and DAP, a re- In Situ Exploration of the Giant Planets -- 14/26 maining degree of freedom - the "b" parameter (the offset of the b-plane aim point from the planet's center), determines both the available entry site locations, and the atmosphere- relative entry speed for each of those locations, and the entry flight path angle (EFPA). If the probe is delivered and sup- ported by a flyby spacecraft, designing a trajectory to give data relay window durations of an hour or more is not difficult. However, if the CRSC is an orbiter delivering the probe from hyperbolic approach, the probe mission must compete with the orbit insertion maneuver for best performance. Although orbit insertion maneuvers are most efficiently done near the pMi17lanet thereby saving propellant mass, such trajectories coupled with a moderately shallow probe EFPA that keeps en- try heating rates and inertial loads relatively low would yield impractically short data relay durations. For the ice giants, a different approach to this problem might avoid this situation by delivering the probe to an aim point ∼180 away from the orbiter's aim point. Although this requires a minor increase in the orbiter's total ∆V for targeting and deflection, it allows a moderate EFPA for the probe while providing a data relay window of up to 2 hours. Probe Entry and Enabling Technologies The probe aer- oshell would comprise both a forward aeroshell (heatshield) and an aft cover (backshell). The aeroshell has five primary functions -- 1) to provide an aerodynamically stable config- uration during hypersonic and supersonic entry and descent into the giant planet H2 -- He atmosphere while spin-stabilized along the probe's symmetry (rotation) axis, 2) to protect the de- scent vehicle from the extreme heating and thermomechanical loads of entry, 3) to accommodate the large deceleration loads from the descent vehicle during hypersonic entry, 4) to provide a safe, stable transition from hypersonic/supersonic entry to subsonic descent, and 5) to safely separate the heatshield and backshell from the descent vehicle based on g-switch with timer backup, and transition the descent vehicle to descent science mode beneath the main parachute. The need for a heatshield to withstand the extreme entry conditions encoun- tered at the giant planets is critical and has been successfully addressed by NASA in the past, and is currently addressed by ESA. Because heritage carbon phenolic thermal protection system (TPS) used for the Galileo and Pioneer Venus entry aeroshell heatshields is no longer available, NASA invested in the development of a new heatshield material and system technology called Heatshield for Extreme Entry Environment Technology (HEEET) and also in upgrading arc jet facilities for ablative material testing at extreme conditions. HEEET is an ablative TPS system that uses 3-D weaving to achieve both robustness and mass efficiency at extreme entry conditions, and being tested at conditions that are relevant for Saturn and ice giant entry probe missions [164]. Compared to heritage In Situ Exploration of the Giant Planets -- 15/26 Figure 6. Galileo entry, descent and deployment sequence shown above could be the basis for any proposed giant planet entry probe mission. carbon phenolic system, HEEET is nearly 50% mass efficient [165]. Alternative TPS concepts and materials are currently under evaluation by ESA (ESA M* Ice Giant CDF study 2). 3.3 Atmospheric Entry Probe System Design Overview The probe comprises two major sub-elements: 1) the Descent Vehicle (DV) including parachutes will carry all the science instruments and support subsystems includ- ing telecommunications, power, control, and thermal into the atmosphere, and 2) the aeroshell that protects the DV during cruise, coast, and entry. The probe (DV and aeroshell) is re- leased from the CRSC, and arrives at the entry interface point following a long coast period. Although the probe reaches the entry interface point and the DV with parachutes descends into the atmosphere, elements of the probe system including the probe release and separation mechanism and the probe telemetry receiver remain with the CRSC. Prior to entry, the probe coast timer (loaded prior to probe release) provides a wakeup call to initiate the entry power-on sequence for initial warmup, checks on instrument and subsystem health and sta- tus, and pre-entry calibrations. Entry peak heating, total heat soak, and deceleration pulse depend on the selected mission design including entry location (latitude/longitude), inertial heading, and flight path angle. Following entry, the DV pro- vides a thermally protected environment for the science in- struments and probe subsystems during atmospheric descent, including power, operational command, timing, and control, and reliable telecommunications for returning probe science and engineering data. The probe avionics will collect, buffer, format, process (as necessary), and prepare all science and engineering data to be transmitted to the CRSC. The probe descent subsystem controls the probe descent rate and rotation necessary to achieve the mission science objectives. Entry Probe Power and Thermal Control Following re- lease from the CRSC, the probe has four main functions: 1) to initiate the "wake up" sequence at the proper time prior to arrival at the entry interface point, 2) to safely house, protect, provide command and control authority for, provide power for, and maintain a safe thermal environment for all the subsystems and science instruments, 3) to collect, buffer as needed, and relay to the CRSC all required preentry, entry, and descent housekeeping, engineering, calibration, and science engineer- ing data, and 4) to control the descent speed and spin rate profile of the descent vehicle to satisfy science objectives and operational requirements. Once released from the CRSC, the probe would be entirely self-sufficient for mission operations, thermal control, and power management. During coast, pre- entry, and entry, the batteries support probe coast functions, wake-up and turn-on, system health checks, and entry and descent operations. Autonomous thermal control is provided during coast by batteries, although there may be an option to replace electrical heating with Radioisotope Heater Units to greatly reduce battery requirements. Giant planet missions may include Venus flybys, where temperatures are higher, prior to the long outer solar system cruise. Since the ice giants are much cooler than the gas giants, descent survival at the low ice giant temperatures may dictate a sealed probe. Providing a safe, stable thermal environment for probe subsystems and instruments over this range of heliocentric distances will re- quire careful thermal design. Future technology developments may realize batteries with higher specific energies resulting in potential mass savings, and the development of electronics operating at cryogenic temperatures. Data Relay The transmit-only probe telecommunication sys- tem would comprise two redundant channels that transmit orthogonal polarizations at slightly offset frequencies for iso- lation. Driven by an ultrastable oscillator to ensure a stable link frequency for probe radio science, the frequency of the probe to CRSC relay link is chosen primarily based on the mi- crowave absorption properties of the atmosphere. The actual thermal, compositional, and dynamical structure beneath the cloud tops of the giant planets remains largely unknown. Pos- sible differences in composition and temperature and pressure structure between the atmosphere models and the true atmo- sphere may adversely affect the performance of the probe relay telecomm and must be considered in selection of communi- cation link frequency. In particular, the microwave opacity of the atmosphere depends on the abundance of trace microwave absorbing species such as H2O, NH3, H2S, and PH3. In gen- eral, the microwave opacity of these absorbers increases as the square of the frequency, and this drives the telecomm fre- quency as low as reasonable, often UHF. At Jupiter, the lowest practical frequency is L-band due to the intense low-frequency synchrotron radiation environment. The final decision on fre- quency consequently affects the overall telecomm link budget, including probe transmit antenna design (type, size, gain, and beam pattern, and beam-width), and pointing requirements for the CRSC-mounted receive antenna. Other decisions affect- ing the telecomm link design include probe descent science requirements, the time required to reach the target depth, and the CRSC overflight trajectory, including range, range rate, and angle. Carrier Relay Spacecraft During the long cruise to the outer solar system, the CRSC provides power as well as struc- tural and thermal support for the probe, and supports periodic health checks, communications for probe science instrument software changes and calibrations, and other probe power and thermal control software configuration changes and mission se- quence loading as might be required from launch to encounter. Upon final approach, the CRSC supports a final probe health and configuration check, rotates to the probe release orienta- tion, cuts cables and releases the probe for the probe cruise to the entry interface point. Following probe release, the CRSC may be tracked for a period of time from Earth, preferably In Situ Exploration of the Giant Planets -- 16/26 several days, to characterize the probe release dynamics and improve reconstructions of the probe coast trajectory and en- try interface location. An important release sequence option would be to image the probe following release for optical nav- igation characterization of the release trajectory. Following probe release and once the CRSC tracking period is over, the CRSC is deflected from the planet-impact trajectory required for probe targeting to a trajectory that will properly position the CRSC for receiving the probe descent telecommunications. During coast, the probe will periodically transmit health status reports to the CRSC. Additionally, the CRSC will conduct a planet-imaging campaign to characterize the time evolution of the atmosphere, weather, and clouds at the probe entry site, as well as to provide global context of the entry site. Prior to the initiation of the probe descent sequence, the CRSC will rotate to the attitude required for the probe relay receive antenna to view the probe entry/descent location and subse- quently prepares to receive both channels of the probe science telecommunications. Once the probe science mission ends, the CRSC will return to Earth-point and downlink multiple copies of the stored probe data. 4. Possible Probe Model Payload Table 4 presents a suite of scientific instruments that can ad- dress the scientific requirements discussed in Section 2. This list of instruments should be considered as an example of sci- entific payload that one might wish to see onboard. Ultimately, the payload of a giant planet probe would be defined from a detailed mass, power and design trades, but should seek to address the majority of the scientific goals outlined in Section 2. Atmospheric Structure Instrument The Atmospheric Struc- ture Instrument (ASI) is a multi-sensor package for in situ mea- surements to investigate the atmospheric structure, dynamics and electricity of the outer planets. The scientific objectives of ASI are the determination of the atmospheric vertical pressure and temperature profiles, the evaluation of the density, and the investigation of the atmospheric electrical properties (e.g. conductivity, lightning). The atmospheric profiles along the entry probe trajectory will be measured from the exosphere down deep into the outer planet's atmosphere. During entry, density will be derived from the probe decelerations; pres- sure and temperature will be computed from the density with the assumption of hydrostatic equilibrium. Direct measure- ments of pressure, temperature and electrical properties will be performed under the parachute, after the front shield jet- tisoning, by sensors having access to the atmospheric flow. ASI will measure the atmospheric state (pressure, temperature) as well as constraining atmospheric stability, dynamics and In Situ Exploration of the Giant Planets -- 17/26 Table 4. Measurement requirements Instrument Atmospheric Structure Instrument Mass spectrometer Tunable Laser System Helium Abundance Detector Ortho-Para Instrument Doppler Wind Experiment Nephelometer Net-Flux Radiometer Measurement Pressure, temperature, density, molecular weight profile , atmospheric conductivity, DC electric field Elemental and chemical composition Isotopic composition High molecular mass organics Isotopic composition Helium abundance Temperature, pressure and density vertical structure Measure winds, speed and direction Cloud structure Solid/liquid particles Thermal/solar energy its effect on atmospheric chemistry. The ASI benefits from the strong heritage of the Huygens HASI experiment of the Cassini/Huygens mission [166], and the Galileo and Pioneer Venus ASI instruments [167, 168]. Mass Spectrometer Experiment The Mass Spectrometer Experiment (MSE) of the entry probe makes in situ measure- ments during the descent into the giant planets atmospheres to determine the chemical and isotopic composition of Uranus and Neptune. The scientific objective of MSE is to measure the chemical composition of the major atmospheric species such as H, C, N, S, P, Ge, and As, all the noble gases He, Ne, Ar, Kr, and Xe, and key isotope ratios of major elements D/H, 13C/12C, 15N/14N, 17O/16O, 18O/16O, of the lighter noble gases 3He/4He, 20Ne/22Ne, 38Ar/36Ar, 36Ar/40Ar, and those of Kr and Xe. Given the constrained resources on the entry probe and the short duration of the descent through the atmosphere, time-of-flight instruments are the preferred choice, with strong heritage from the ROSINA experiment on the Rosetta mission [169] (see Fig. 7). The mass spectrometer itself will be com- plemented by a complex gas introduction system handling the range of atmospheric pressures during descent, a reference gas calibration system, and enrichment cells for improving the detection of noble gases and hydro carbons. Tunable Laser Spectrometer A Tunable Laser Spectrome- ter (TLS) [170] will complement the mass spectrometric mea- surements by providing a few isotopic measurements with high accuracy, e.g. D/H, 13C/12C, 18O/16O, and 17O/16O, de- pending on the selected laser system. TLS employs ultra-high spectral resolution (0.0005 cm−1) tunable laser absorption spectroscopy in the near infra-red (IR) to mid-IR spectral re- gion. A TLS is part of the SAM instrument on the NASA Curiosity Rover [171], which was used to measure the isotopic ratios of D/H and of 18O/16O in water and 13C/12C, 18O/16O, 17O/16O, and 13C18O/12C16O in carbon dioxide in the Martian atmosphere [172]. Helium Abundance Detector The Helium Abundance De- tector (HAD), as it was used on the Galileo mission [173, 5], measures the refractive index of the atmosphere in the pressure range of 2 -- 10 bar. The refractive index is a function of the composition of the sampled gas, and since the jovian atmo- sphere consists of mostly of H2 and He, to more than 99.5%, the refractive index is a direct measure of the He/H2 ratio. The refractive index can be measured by any two-beam interfer- ometer, where one beam passes through a reference gas and the other beam through atmospheric gas. The difference in the optical path gives the difference in refractive index between the reference and atmospheric gas. For the Galileo mission, a Jamin-Mascart interferometer was used, because of its sim- ple and compact design, with a high accuracy of the He/H2 measurement. Doppler-Wind Experiment The Doppler Wind Experiment (DWE) will use the probe-CRSC radio subsystem (with ele- ments mounted on both the probe and the Carrier) to measure the altitude profile of zonal winds along the probe descent path under the assumption that the probe in terminal descent beneath the parachute will move with the winds. The DWE will also reflect probe motions due to atmospheric turbulence, aerodynamic buffeting, and atmospheric convection and waves that disrupt the probe descent speed. Key to the Doppler wind measurement is an accurate knowledge of the reconstructed probe location at the beginning of descent, the probe descent speed with respect to time/altitude, and the CRSC position and velocity throughout the period of the relay link. The initial probe descent location depends upon the probe entry trajectory from the entry point to the location of parachute deployment and is reconstructed from measured accelerations during en- try. The descent profile is reconstructed from Atmospheric Structure Instrument measurements of pressure and tempera- In Situ Exploration of the Giant Planets -- 18/26 Figure 8. A schematics of the laboratory model of the TLS spectrometer for the Martian Phobos Grunt mission [170]. With the TLS, four near-infrared laser diodes are injected in a single-path tube filled up with the gases to analyse. The laser beam are partially absorbed by the ambient molecules. The gas concentrations for the various isotopologues are then retrieved from the achieved absorption spectra. up to 1,000 particles per cm3. Such an instrument performs counting measurements at a small scattering angle, to retrieve the size distribution based on the work of [176]. It applies the principle of the Light Optical Aerosol Counter (LOAC) optical aerosols counter used since 2013 under all kinds of atmospheric balloons [178, 177]. These measurements allow one to retrieve the size distribution of the particles typically for 20 size-classes in the 0.2-50 µm range. Also, simultaneous measurements can be conducted at up to 10 scattering angles in the 20 -- 170◦ range, to retrieve the scattering function for each size range. The retrieval of the nature of the aerosols can be conducted by comparing these observed scattering func- tions to theoretical ones computed for scattering theories, and to reference measurements obtained in laboratory for solid particles [179, 180]. Ortho-Para Instrument Vertical mixing in giant planet tro- pospheres carrying significant heat from the deeper atmo- spheres to upper levels where it can be radiated to space is modulated by the atmospheric stability and can be dramatically changed by the condensation and evaporation of CH4, H2S, NH3, and H2O. Thermal profiles and stabilities in the colder outer solar system can be further affected by the atmospheric hydrogen para-fraction [181]. Hydrogen molecules come in two types -- with proton spins aligned (ortho-hydrogen) or op- posite (para-hydrogen), each with significantly different ther- modynamic properties at low temperatures. To interpret the thermal profile and stability, density structure, aerosol layering, net fluxes and vertical motions of giant planet atmospheres, the hydrogen para-fraction must be known, with increasing im- portance for the colder ice giants. The ortho- to para-hydrogen Figure 7. Flight model of DFMS/ROSINA instrument without thermal hardware [169]. ture during descent. From the reconstructed probe and CRSC positions and velocities, a profile of the expected relay link frequencies is found that can be differenced with the mea- sured frequencies to generate a set of frequency residuals. The winds are retrieved utilizing an inversion algorithm similar to the Galileo probe Doppler Wind measurement [174, 39]. To generate the stable probe relay signal, the probe must carry an ultrastable oscillator (USO) with an identical USO in the relay receiver on the Carrier spacecraft. Nephelometer Measurement of scattered visible light within the atmosphere is a powerful tool to retrieve number density and size distribution of liquid and solid particles, related to their formation process, and to understand the overall charac- ter of the atmospheric aerosols based on their refractive index (liquid particles, iced particles, solid particles from transparent to strongly absorbing). In particular, measurements of light scattered by a cloud of particles at several scattering angles was already tested on balloon flights to characterize the atmo- spheric aerosols and condensates [175], using a priori hypoth- esis on the size distribution. A new concept of nephelometer has been proposed to retrieve the full scattering function, this time for individual particles crossing a light source. Dedi- cated fast electronics are necessary to enable the detection of In Situ Exploration of the Giant Planets -- 19/26 radiative flux and upward radiation flux within their respective atmospheres as the probe descended by parachute. A future Net Flux Radiometer could build on the lessons learned from the Galileo probe NFR experiment and is designed to deter- mine the net radiation flux within all giant planets atmospheres. The nominal measurement regime for the NFR extends from ∼0.1 bar to at least 10 bars. These measurements will help us to define sources and sinks of planetary radiation, regions of solar energy deposition, and provide constraints on atmo- spheric composition and cloud layers. The primary objective of the NFR is to measure upward and downward radiative fluxes to determine the radiative heating (cooling) component of the atmospheric energy budget, determine total atmospheric opacity, identify the location of cloud layers and opacities, and identify key atmospheric absorbers such as methane, ammonia, and water vapor. The NFR can measure upward and downward flux densities in multiple spectral channels. 5. International Collaboration Only ESA/Europe and NASA/USA collaborations are consid- ered here. However collaborations with other international partners may be envisaged. For several reasons, the partici- pation of and contributions from NASA are essential for an ESA-led entry probe. NASA has proven its ability to send spacecrafts beyond 5 AU thanks to the use of radioisotope power systems. Although solar panel technologies likely en- able the sending of spacecraft up to the distance of Saturn over the next decade, radioisotope power systems are required to reach the heliocentric distances of the Ice Giants. Also, because of their (relatively) small sizes, probes are ideal com- panion spacecraft to be included in ambitious missions similar to Cassini-Huygens or Galileo. An ESA giant planet probe mission could begin its flight phase as an element of a NASA Saturn, Uranus or Neptune mission (likely a NASA Flagship or New Frontiers mission). The launch would place both the NASA spacecraft, which functions also as the probe's CRSC, and the probe on a transfer trajectory to the giant planets. One of the key probe technologies for an entry probe that is critical for European industry is the heat shield material. If the Euro- pean TPS is too heavy, then alternative material such as the NASA HEEET could be utilized in the context of a partnership between ESA and NASA. 6. Education and Public Outreach (EPO) The interest of the public in the giant planets continues to be significant, with much of the credit for the high interest in Saturn and Jupiter, due to the extraordinary success of the Cassini -- Huygens mission and the currently ongoing Juno mis- sion. Images from the Saturnian system and Jupiter are regu- larly featured as the NASA "Astronomy Picture of the Day", Figure 9. The LOAC instrument used at present for short and long duration balloon flights. This version performs measurements at two scattering angles, while more angles are expected for the space version LONSCAPE ratio can be measured by exploiting the thermodynamic differ- ences between these two forms of hydrogen, which affects the speed of sound. Assuming atmospheric temperature and mean molecular weight are known, the ortho- to para-hydrogen ratio can be found from speed of sound measurements using a pair of ultrasonic capacitive transducers and sophisticated signal processing techniques. Acoustic travel times can be measured to ∼10 ns for travel times in the 0.5ms range (one part in 5e- 4) using a high TRL, compact, energy-efficient and low data volume ultrasonic anemometer originally developed for Mars [182]. Net Energy Flux Radiometer Giant planet meteorology re- gimes depend on internal heat flux levels. Downwelling solar insolation and upwelling thermal energy from the planetary interior can have altitude and location dependent variations. Such radiative-energy differences cause atmospheric heating and cooling, and result in buoyancy differences that are the primary driving force for giant planets atmospheric motions. Three notable Net Flux Radiometer (NFR) instruments have flown in the past namely, the Large probe Infrared Radiometer (LIR) [183] on the Venus Probe, the NFR on the Galileo Probe [40], and the DISR on the Huygens Probe [184] for in situ measurements within Venus, Jupiter and Titan's atmospheres, respectively. All instruments were designed to measure the net and continue to attract the interest of the international media. The interest and excitement of students and the general public can only be amplified by a return to Saturn or an unprecedented mission toward Uranus and/or Neptune. An entry probe mis- sion will hold appeal for students at all levels. Education and Public Outreach activities will be an important part of the mission planning. An EPO team will be created to develop programs and activities for the general public and students of all ages. Additionally, results and interpretation of the science will be widely distributed to the public through internet sites, leaflets, public lectures, TV and radio programmes, numerical supports, museum and planetarium exhibitions, and in popular science magazines and in newspapers. 7. Summary and Perspectives The next great planetary exploration mission may well be a flagship mission to Saturn, or one of the ice giant planets. This could be possibly a mission to Uranus with its unique obliquity and correspondingly extreme planetary seasons, its unusual dearth of cloud features and radiated internal energy, a tenuous ring system and multitude of small moons, or to the Neptune system, with its enormous winds, system of ring arcs, sporadic atmospheric features, and large retrograde moon Triton, likely a captured dwarf planet. The ice giant planets represent the last unexplored class of planets in the solar system, yet the most frequently observed type of exoplanets. Extended studies of Saturn, or one or both ice giants, including in situ measure- ments with an entry probe, are necessary to further constrain models of solar system formation and chemical, thermal, and dynamical evolution, the atmospheric formation, evolution, and processes, and to provide additional ground-truth for im- proved understanding of extrasolar planetary systems. The giant planets, gas and ice giants together, additionally offer a laboratory for studying the dynamics, chemistry, and processes of the terrestrial planets, including Earth's atmosphere. Only in situ exploration by a descent probe (or probes) can unlock the secrets of the deep, well-mixed atmospheres where pristine materials from the epoch of solar system formation can be found. Particularly important are the noble gases, undetectable by any means other than direct sampling, that carry many of the secrets of giant planet origin and evolution. Both absolute as well as relative abundances of the noble gases are needed to understand the properties of the interplanetary medium at the location and epoch of solar system formation, the delivery of heavy elements to the giant planet atmospheres, and to help decipher evidence of possible giant planet migration. A key result from a Saturn, Uranus or Neptune entry probe would be the indication as to whether the enhancement of the heavier noble gases found by the Galileo probe at Jupiter (and hope- fully confirmed by a future Saturn probe) is a feature common In Situ Exploration of the Giant Planets -- 20/26 to all the giant planets, or is limited only to the largest gas giant. This could have broad implications for the properties of known exoplanets of both giant and ice types, specially in planetary systems sharing both types of exoplanets. The primary goal of a giant planet entry probe mission is to measure the well-mixed abundances of the noble gases He, Ne, Ar, Kr, Xe and their isotopes, the heavier elements C, N, S, and P, key isotope ratios 15N/14N, 13C/12C, 17O/16O and 18O/16O, and D/H, and disequilibrium species CO and PH3, which act as tracers of internal processes, and can be achieved by a probe reaching 10 bars. In addition to measurements of the noble gases, chemical, and isotopic abundances in the atmosphere, a probe would measure many of the chemical and dynamical processes within the upper atmosphere, providing an improved context for understanding the chemistries, pro- cesses, origin, and evolution of all the atmospheres in the solar system. Moreover, the choice of an ice giant (Uranus or Nep- tune) entry probe would allow understanding the formation conditions of the entire family of all giant planets, and to pro- vide ground-truth measurement to improve understanding of extrasolar planets. A descent probe would sample atmospheric regions far below those accessible to remote sensing, well into the cloud forming regions of the troposphere to depths where many cosmogenically important and abundant species are ex- pected to be well-mixed. Along the descent, the probe would provide direct tracking of the planet's atmospheric dynam- ics including zonal winds, waves, convection and turbulence, measurements of the thermal profile and stability of the at- mosphere, and the location, density, and composition of the upper cloud layers. Results obtained from a giant planet en- try probe, and more importantly from an ice giant probe, are necessary to improve our understanding of the processes by which all the giants formed, including the composition and properties of the local solar nebula at the time and location of ice giant formation. By extending the legacy of the Galileo probe mission, Saturn, Uranus and/or Neptune probe(s) will further discriminate competing theories addressing the forma- tion, and chemical, dynamical, and thermal evolution of the giant planets, the entire solar system including Earth and the other terrestrial planets, and the formation of other planetary systems. Acknowledgments O.M. acknowledges support from CNES. L.N.F. was supported by a Royal Society Research Fellowship. R.H. and A.S.L were supported by the Spanish MINECO project AYA2015-65041-P (MINECO/FEDER, UE) and Grupos Gobierno Vasco IT-1366- 19 from Gobierno Vasco. References [1] Gomes, R., Levison, H. F., Tsiganis, K., et al. 2005, Nature, 435, 466 [2] Chambers, J. E., & Wetherill, G. W. 2001, Meteoritics and Planetary Science, 36, 381 [3] Owen, T., Mahaffy, P., Niemann, H. B., et al. 1999, Nature, 402, 269 [4] Mousis, O., Atkinson, D. H., Cavali´e, T., et al. 2018, Plan- etary and Space Science, 155, 12 [5] von Zahn, U., Hunten, D. M., & Lehmacher, G. 1998, Journal of Geophysical Research, 103, 22815 [6] Owen, T., Mahaffy, P. R., Niemann, H. B., et al. 2001, The Astrophysical Journal Letters, 553, L77 [7] Orton, G. S., Fisher, B. M., Baines, K. H., et al. 1998, Journal of Geophysical Research, 103, 22791 [8] Wong, M. H., Mahaffy, P. R., Atreya, S. K., et al. 2004, Icarus, 171, 153 [9] Hofstadter, M., Simon, A., Atreya, S., et al. 2019, Planetary and Space Science, in press [10] Banfield, D., Simon, A., Danner, R., et al. 2018, 2018 IEEE Aerospace Conference, Big Sky, MT, 2018, pp. 1-15, 10.1109/AERO.2018.8396829 [11] Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 124, 62 [12] Alibert, Y., Mousis, O., & Benz, W. 2005, The Astrophys- ical Journal Letters, 622, L145 [13] Alibert, Y., Mousis, O., Mordasini, C., et al. 2005, The Astrophysical Journal Letters, 626, L57 [14] Dodson-Robinson, S. E., & Bodenheimer, P. 2010, Icarus, 207, 491 [15] Helled, R., Bodenheimer, P., Podolak, M., et al. 2014, Protostars and Planets VI, 643 [16] Mizuno, H., Nakazawa, K., & Hayashi, C. 1978, Progress of Theoretical Physics, 60, 699 [17] Lin, D. N. C., & Papaloizou, J. 1986, The Astrophysical Journal, 307, 395 [18] Ward, W. R. 1997, Icarus, 126, 261 [19] Ida, S., & Lin, D. N. C. 2004, The Astrophysical Journal, 616, 567 [20] Mordasini, C., Alibert, Y., Klahr, H., et al. 2012, Astron- omy & Astrophysics, 547, A111 [21] Boss, A. P. 1997, Science, 276, 1836 [22] Boss, A. P. 2001, The Astrophysical Journal, 563, 367 In Situ Exploration of the Giant Planets -- 21/26 [23] Vazan, A., Helled, R., & Guillot, T. 2018, Astronomy & Astrophysics, 610, L14 [24] Wahl, S. M., Hubbard, W. B., Militzer, B., et al. 2017, Geophysical research Letters, 44, 4649 [25] Nettelmann, N. 2017, Astronomy & Astrophysics, 606, A139 [26] Helled, R., & Guillot, T. 2013, The Astrophysical Journal, 767, 113 [27] Wilson, H. F., & Militzer, B. 2012, Physical Review Let- ters, 108, 111101 [28] Wilson, H. F., & Militzer, B. 2012, The Astrophysical Journal, 745, 54 [29] Leconte, J., & Chabrier, G. 2012, Astronomy & Astro- physics, 540, A20 [30] Leconte, J., & Chabrier, G. 2013, Nature Geoscience, 6, 347 [31] Saumon, D., & Guillot, T. 2004, The Astrophysical Jour- nal, 609, 1170 [32] Fortney, J. J., & Nettelmann, N. 2010, Space Science Reviews, 152, 423 [33] Nettelmann, N., Helled, R., Fortney, J. J., et al. 2013, Planetary and Space Science, 77, 143 [34] Gautier, D., Hersant, F., Mousis, O., et al. 2001, The Astrophysical Journal Letters, 550, L227 [35] Hersant, F., Gautier, D., & Hur´e, J.-M. 2001, The Astro- physical Journal, 554, 391 [36] Young, R. E. 1998, Journal of Geophysical Research, 103, 22775 [37] Folkner, W. M., Woo, R., & Nandi, S. 1998, Journal of Geophysical Research, 103, 22847 [38] Ragent, B., Colburn, D. S., Rages, K. A., et al. 1998, Journal of Geophysical Research, 103, 22891 [39] Atkinson, D. H., Pollack, J. B., & Seiff, A. 1998, Journal of Geophysical Research, 103, 22911 [40] Sromovsky, L. A., Collard, A. D., Fry, P. M., et al. 1998, Journal of Geophysical Research, 103, 2929 [41] Niemann, H. B., Atreya, S. K., Carignan, G. R., et al. 1998, Journal of Geophysical Research, 103, 22831 [42] Atreya, S. K., Wong, M. H., Owen, T. C., et al. 1999, Planetary and Space Science, 47, 1243 [43] Wilson, H. F., & Militzer, B. 2010, Physical Review Let- ters, 104, 121101 [44] Guillot, T. 2005, Annual Review of Earth and Planetary Sciences, 33, 493 In Situ Exploration of the Giant Planets -- 22/26 [45] Helled, R., Podolak, M., & Kovetz, A. 2006, Icarus, 185, [67] Lindal, G. F., Lyons, J. R., Sweetnam, D. N., et al. 1990, 64 [46] Atreya, S. K., Mahaffy, P. R., Niemann, H. B., et al. 2003, Planetary and Space Science, 51, 105 [47] Matousek, S. 2007, Acta Astronautica, 61, 932 [48] Helled, R., & Lunine, J. 2014, Monthly Notices of the Royal Astronomical Society, 441, 2273 [49] Conrath, B. J., Gautier, D., Hanel, R. A., et al. 1984, The Astrophysical Journal, 282, 807 [50] Conrath, B. J., & Gautier, D. 2000, Icarus, 144, 124 [51] Achterberg, R. K., Schinder, P. J., & Flasar, F. M. 2016, AAS/Division for Planetary Sciences Meeting Abstracts #48, 508.01 [52] Courtin, R., Gautier, D., Marten, A., et al. 1984, The Astrophysical Journal, 287, 899 [53] Davis, G. R., Griffin, M. J., Naylor, D. A., et al. 1996, Astronomy & Astrophysics, 315, L393 [54] Fletcher, L. N., Irwin, P. G. J., Teanby, N. A., et al. 2007, Icarus, 188, 72 [55] Fletcher, L. N., Orton, G. S., Teanby, N. A., et al. 2009, Icarus, 202, 543 [56] Fletcher, L. N., Orton, G. S., Teanby, N. A., et al. 2009, Icarus, 199, 351 [57] Mousis, O., Fletcher, L. N., Lebreton, J.-P., et al. 2014, Planetary and Space Science, 104, 29 [58] Mousis, O., Atkinson, D. H., Spilker, T., et al. 2016, Plan- etary and Space Science, 130, 80 [59] Atkinson, D. H., Simon, A. A., Banfield, D., et al. 2016, AAS/Division for Planetary Sciences Meeting Abstracts #48, 123.29 [60] Conrath, B., Gautier, D., Hanel, R., et al. 1987, Journal of Geophysical Research, 92, 15003 [61] Burgdorf, M., Orton, G. S., Davis, G. R., et al. 2003, Icarus, 164, 244 [62] Mahaffy, P. R., Niemann, H. B., Alpert, A., et al. 2000, Journal of Geophysical Research, 105, 15061 [63] Lindal, G. F., Lyons, J. R., Sweetnam, D. N., et al. 1987, Journal of Geophysical Research, 92, 14987 [64] Baines, K. H., Mickelson, M. E., Larson, L. E., et al. 1995, Icarus, 114, 328 [65] Karkoschka, E., & Tomasko, M. 2009, Icarus, 202, 287 [66] Sromovsky, L. A., Karkoschka, E., Fry, P. M., et al. 2014, Icarus, 238, 137 Geophysical Research Letters, 17, 1733 [68] Karkoschka, E., & Tomasko, M. G. 2011, Icarus, 211, 780 [69] Fletcher, L. N., Baines, K. H., Momary, T. W., et al. 2011, Icarus, 214, 510 [70] Irwin, P. G. J., Toledo, D., Garland, R., et al. 2018, Nature Astronomy, 2, 420 [71] Irwin, P. G. J., Toledo, D., Garland, R., et al. 2019, Icarus, 321, 550 [72] Lodders, K., Palme, H., & Gail, H.-P. 2009, Landolt Born- stein, 4B, 712 [73] Feuchtgruber, H., Lellouch, E., Orton, G., et al. 2013, Astronomy & Astrophysics, 551, A126 [74] Ali-Dib, M., Mousis, O., Petit, J.-M., et al. 2014, The Astrophysical Journal, 793, 9 [75] Geiss, J., & Gloeckler, G. 1998, Space Science Reviews, 84, 239 [76] Marty, B., Chaussidon, M., Wiens, R. C., et al. 2011, Science, 332, 1533 [77] Fouchet, T., Lellouch, E., B´ezard, B., et al. 2000, Icarus, 143, 223 [78] Fletcher, L. N., Greathouse, T. K., Orton, G. S., et al. 2014, Icarus, 238, 170 [79] Mousis, O., Lunine, J. I., Fletcher, L. N., et al. 2014, The Astrophysical Journal, 796, L28 [80] Rubin, M., Altwegg, K., Balsiger, H., et al. 2015, Science, 348, 232 [81] Le Roy, L., Altwegg, K., Balsiger, H., et al. 2015, Astron- omy & Astrophysics, 583, A1 [82] Bockel´ee-Morvan, D., Crovisier, J., Mumma, M. J., et al. 2004, Comets II, 391 [83] Rousselot, P., Pirali, O., Jehin, E., et al. 2014, The Astro- physical Journal Letters, 780, L17 [84] Manfroid, J., Jehin, E., Hutsem´ekers, D., et al. 2009, As- tronomy & Astrophysics, 503, 613 [85] Mahaffy, P. R., Donahue, T. M., Atreya, S. K., et al. 1998, Space Science Reviews, 84, 251 [86] Lellouch, E., B´ezard, B., Fouchet, T., et al. 2001, Astron- omy & Astrophysics, 370, 610 [87] Dobrijevic, M., & Loison, J. C. 2018, Icarus, 307, 371 [88] Asplund, M., Grevesse, N., Sauval, A. J., et al. 2009, Annual Reviews of Astronomy & Astrophysics, 47, 481 [89] Bockel´ee-Morvan, D., Biver, N., Swinyard, B., et al. 2012, Astronomy & Astrophysics, 544, L15 In Situ Exploration of the Giant Planets -- 23/26 [90] Courtin, R., Swinyard, B. M., Moreno, R., et al. 2011, [112] Garcia-Melendo, E., P´erez-Hoyos, S., S´anchez-Lavega, Astronomy & Astrophysics, 536, L2 A. et al. Icarus, 215, 62 [91] Loison, J. C., Dobrijevic, M., Hickson, K. M., et al. 2017, [113] Tollefson, J., de Pater, I., Marcus, P. S. et al. 2018, Icarus, Icarus, 291, 17 311, 317 [92] Serigano, J., Nixon, C. A., Cordiner, M. A., et al. 2016, The Astrophysical Journal Letters, 821, L8 [93] Noll, K. S., Geballe, T. R., & Knacke, R. F. 1995, The Astrophysical Journal Letters, 453, L49 [94] Boss, A. P., Wetherill, G. W., & Haghighipour, N. 2002, Icarus, 156, 291 [95] Bar-Nun, A., Notesco, G., & Owen, T. 2007, Icarus, 190, 655 [96] Mousis, O., Lunine, J. I., Picaud, S., et al. 2010, Faraday Discussions, 147, 509 [97] Mousis, O., Lunine, J. I., Madhusudhan, N., et al. 2012, The Astrophysical Journal Letters, 751, L7 [98] Guillot, T., & Hueso, R. 2006, Monthly Notices of the Royal Astronomical Society, 367, L47 [99] Stevenson, D. J., & Lunine, J. I. 1988, Icarus, 75, 146 [100] Cyr, K. E., Sears, W. D., & Lunine, J. I. 1998, Icarus, 135, 537 [101] Guillot, T. 1995, Science, 269, 1697 [102] Visscher, C., & Fegley, B. 2005, The Astrophysical Jour- nal, 623, 1221 [103] Cavali´e, T., Venot, O., Selsis, F., et al. 2017, Icarus, 291, 1 [104] S´anchez-Lavega, A. et al. 2019, in Zonal Jets: Phe- nomenology, Genesis and Physics, Cambrige University PRess, eds: B. Galperin and P. L. Read. [105] Vasavada, A. R., Showman, A. P. 2005, Reports on Progress in Physics, 68, 1935 [106] Kaspi, Y., Galanti, E., Hubbard, W. B. et al. 2018, Nature, 555, 223 [107] Guillot, T., Miguel, Y., Millitzer, B. et al. 2018, Nature, 555, 227 [108] Galanti, E., Kaspi, Y., Miguel, Y. et al. 2019, Geophysical Research Letters, 46, 616 [109] Kaspi, Y., Showman, A. P., Hubbard, W. B. et al. 2013, Nature, 497, 344 [110] Atkinson, D. H., Pollack, J. B., Seiff, A. 1996, Science, 272, 842 [111] Atkinson, D. H., Ingersoll, A. P., Seiff, A. 1997, Nature, 388, 649 [114] Molter, E., de Pater, I., Luszcz-Cook, S. et al. 2019, Icarus, 321, 324 [115] Sun, Z. P., Schubert, G., Stoker, C. R. 1991, Icarus, 91, 154 [116] Lian Y. and A. P. Showman 2010, Icarus, 207, 373 [117] Fletcher, L. N., Irwin, P.G.J., Teanby, N. A., et al. 2007, Icarus, 189, 457 [118] Orton, G. S. Fletcher, L. N., Moses, J. I. et al. 2014, Icarus, 243, 494 [119] Fletcher, L. N., de Pater, I., Orton, G. S. et al. 2014, Icarus, 231, 146 [120] Lindal, G. F., Lyons, J.R., Sweetnam, D. N. et al. 1987, Journal of Geophys. Res. 92, 14987 [121] Lindal, G. F. 1992, The Astronomical Journal, 103, 967 [122] Leconte, K., Selsis, F., Hersant, F. et al. 2017, Astronomy & Astrophysics, 598, A98 [123] Friedson, A. J., & Gonzales, E. J. 2017, Icarus, 297, 160 [124] Herbert, F., Sandel, B. R., Yelle, R. V. et al. 1987, Journal of Geophys. Res., 92, 15093 [125] Li, C., Le, T., Zhang X., et al. 2018, Journal of Quantita- tive Spectroscopy and Radiative Transfer, 217, 353. [126] Banfield, D., Gierasch, P. J., Bell, M. et al. 1998, Icarus, 135, 230 [127] West, R. A., Baines, K. H., Friedson, J.A. et al. 2004, in Jupiter the Planet, Satellites and Magnetosphere, Cam- bridge University Press, eds: Bagenal, F., Dowling, T. E., McKinnon, W. B. [128] P´erez-Hoyos, S., Sanz-Requena, J.F., Barrado-Izagirre, N. et al. 2012, Icarus, 217, 256 [129] West, R. A., Baines, K. H., Karkoschka, E. et al. 2009, in Saturn from Cassini-Huygens, Cambridge University Press, eds: Dougherty, M. K., Esposito, L. W., Krimigis, S.M. [130] Fletcher, L. N., Baines, K. H., Momary, T. W. et al. 2011, Icarus, 214, 510 [131] P´erez-Hoyos, S., S´anchez-Lavega, A., Irwin, P. G. J. et al. 2016, Icarus, 277, 1 [132] West, R. A., Baines, K. H., Pollack, J. B., 1991, in Uranus, University of Arizona Press, eds: Bergstralh, J.T., Miner, E. D. Matthews, M.S. In Situ Exploration of the Giant Planets -- 24/26 [133] Irwin, P. G. J., Teanby, N. A., Davis, G. R. 2009, Icarus, [155] Stoker, C. R., Toon, O. B. 1989, Geophys. Res. Lett. 16, 203, 287 929 [134] de Kleer, K., Luszcz-Cook, S., de Pater, I. et al. 2015, Icarus, 256, 120 [135] Irwin, P. G. J., Wong, M. H. Simon, A. A. et al. 2017, Icarus, 288, 99 [156] Showman, A. P., Ingersoll, A. P., Achterberg, R. et al. 2018, in Saturn in the 21st Century, Cambridge University Press, eds: Baines, K. H., Flasar, F. M., Krupp, N. et al. [157] Wong, M. H., Tollefson, J., Hsu, A. I. et al. 2018, The [136] Hammel, H. B., Baines, K. H., Bergstralh, J. T., 1989, Astronomical Journal, 155, 117 Icarus, 80, 416 [137] Baines, K. H., Hammel, H. B., 1994, Icarus, 109, 20 [138] Irwin, P. G. J., Fletcher, L. N., Tice, D. et al. 2016, Icarus, 271, 418 [158] Hammel, H. B., Sromovsky, L. A., Fry, P. M. et al. 2009 ,Icarus, 201, 257 [159] Sromovsky, L. A., de Pater, I., Fry, P. M. et al. 2015, Icarus, 258, 192 [139] Ingersoll, A. P., Dowling, T. E., Gierasch, P. J., 2004, in Jupiter the Planet, Satellites and Magnetosphere, Cam- bridge University Press, eds: Bagenal, F., Dowling, T. E., McKinnon, W. B. [140] Vasavada, A. R., Horst, S. M., Kennedy, M. R., et al. 2006, Journal of Geophysical Research (Planets) 111, 5004 [141] Smith, B. A., Soderblom, L. A., Beebe, R. et al. 1986, Science, 233, 43 [142] Smith, B. A., Soderblom, L. A., Banfield, D. et al. 1989, Science, 246, 1422 [160] Orton, G. S., Fisher, B. M., Baines, K. H. et al. 1998, Journal of Geophysical Research, 103, 22791 [161] Taylor, F. W., Atreya, S. K., Encrenaz, Th. et al. 2004, in Jupiter the Planet, Satellites and Magnetosphere, Cam- bridge University Press, eds: Bagenal, F., Dowling, T. E., McKinnon, W. B. [162] Fouchet, T., Moses, J. I., Conrath, B. J. et al. 2009, in Saturn from Cassini-Huygens, Cambridge University Press, eds: Dougherty, M. K., Esposito, L. W., Krimigis, S.M. [163] Moses, K. I. Fletcher, L. N., Greathouse, T. K. et al. 2018, [143] Sromovsky, L. A., Fry, P. M., Dowling, T. E. 2001, Icarus, Icarus 307, 124 149, 459 [144] Fry, P. M., Sromovsky, L. A., de Pater, I., et al. 2012, The Astronomical Journal, 143, 150 [145] Fletcher, L. N., Kaspi, Y., Guillot, T. et al. 2019, arXiv:1907:01822 [146] Weidenschilling, S. J., Lewis, J. S. 1973, Icarus, 20, 465 [147] Atreya, S. K., Wong, A. S. 2005, Space Sci. Rev., 116, 121 [148] Irwin, P. G. J., Toledo, D., Garland, R. 2018, Nature Astronomy, 2, 420 [149] Irwin, P. G. J., Toledo, D., Garland, R. 2019, Icarus, 321, 550 [150] Showman, A. P., Ingersoll, A. P., Icarus, 132, 205 [151] Fischer, G., Pagaran, J. A., Zarka, P. et al. 2019, Astron- omy & Astrophysics, 621, A113 [152] S´anchez-Lavega, A., Fischer, G., Fletcher, L. N. et al. 2018, in Saturn in the 21st Century, Cambridge University Press, eds: Baines, K. H., Flasar, F. M., Krupp, N. et al. [153] de Pater, I., Sromovsky, L. A., Fry, P. M. et al. 2015, Icarus, 252, 121 [164] Venkatapathy, E., Ellerby, D., Gage, P. 2019. In: Work- shop on In Situ Exploration of Ice Giants, Marseille, France [165] Milos, F. S., Chen, Y.-K., & Mahzari, M. 2017, Journal of Spacecraft and Rockets, 47th AIAA Thermophysics Conference, AIAA AVIATION Forum, https://doi.org/10.2514/6.2017-3353 In: [166] Fulchignoni, M., Ferri, F., Angrilli, F., et al. 2002, Space Science Reviews, 104, 395 [167] Seiff, A., & Knight, T. C. D. 1992, Space Science Re- views, 60, 203 [168] Seiff, A., Juergens, D. W., & Lepetich, J. E. 1980, IEEE Transactions on Geoscience and Remote Sensing, 18, 105 [169] Balsiger, H., Altwegg, K., Bochsler, P., et al. 2007, Space Science Reviews, 128, 745 [170] Durry, G., Li, J. S., Vinogradov, I., et al. 2010, Applied Physics B: Lasers and Optics, 99, 339 [171] Webster, C. R., & Mahaffy, P. R. 2011, Planetary and Space Science, 59, 271 [172] Webster, C. R., Mahaffy, P. R., Flesch, G. J., et al. 2013, Science, 341, 260 [154] Irwin, P. G. J., Wong, M. H., Simon, A. A. et al. 2017, [173] von Zahn, U., & Hunten, D. M. 1992, Space Science Icarus, 288, 99 Reviews, 60, 263 In Situ Exploration of the Giant Planets -- 25/26 [174] Atkinson, D. H., Ingersoll, A. P., & Seiff, A. 1997, Na- ture, 388, 649 [175] Gayet, J. F., Cr´epel, O., Fournol, J. F., et al. 1997, An- nales Geophysicae, 15, 451 [176] Lurton, T., Renard, J.-B., Vignelles, D., et al. 2014, At- mospheric Measurement Techniques, 7, 931 [177] Renard, J.-B., Dulac, F., Berthet, G., et al. 2016, Atmo- spheric Measurement Techniques, 9, 3673 [178] Renard, J.-B., Dulac, F., Berthet, G., et al. 2016, Atmo- spheric Measurement Techniques, 9, 1721 [179] Renard, J.-B., Berthet, G., Robert, C., et al. 2002, Ap- plied Optics, 41, 7540 [180] Volten, H., Munoz, O., Hovenier, J. W., et al. 2006, J. Quant. Spectrosc. Radiat. Transf., 100, 437 [181] Smith, M. D., & Gierasch, P. J. 1995, Icarus, 116, 159 [182] Banfield, D., Schindel, D. W., Tarr, S., et al. 2016, Acous- tical Society of America Journal, 140, 1420 [183] Boese, R. W., Twarowski, R. J., Gilland, J., et al. 1980, IEEE Transactions on Geoscience and Remote Sensing, 18, 97 [184] Tomasko, M. G., Buchhauser, D., Bushroe, M., et al. 2002, Space Science Reviews, 104, 469 Core Proposing Team Olivier Mousis Aix Marseille Univ, CNRS, CNES, LAM, Marseille, France David H. Atkinson Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr., Pasadena, CA, 91109, USA Richard Ambrosi Department of Physics and Astronomy, University of Leicester, Leicester, UK Sushil Atreya Department of Atmospheric, Oceanic, and Space Sciences, University of Michigan, Ann Arbor, MI, 48109-2143, USA Don Banfield Cornell Center for Astrophysics and Planetary Science, Cornell University, 420 Space Sciences, Ithaca, NY 14853, USA Stas Barabash Swedish Institute of Space Physics, Box 812, 98128, Kiruna, Sweden Michel Blanc Institut de Recherche en Astrophysique and Plan´etologie (IRAP), CNRS/Universit´e Paul Sabatier, 31028 Toulouse, France Thibault Cavali´e Laboratoire d'astrophysique de Bordeaux, University Bordeaux, CNRS, B18N, all´ee Geoffroy Saint- Hilaire, 33615 Pessac, France Athena Coustenis LESIA, Observatoire de Paris, PSL Re- search University, CNRS, Sorbonne Universit´es, UPMC Univ. Paris 06, Univ. Paris Diderot, Sorbonne Paris Cit´e, 5 place Jules Janssen, 92195 Meudon, France Magali Deleuil Aix Marseille Univ, CNRS, CNES, LAM, Marseille, France Georges Durry Groupe de Spectrom´etrie Mol´eculaire et At- mosph´erique, UMR 7331, CNRS, Universit´e de Reims, Cham- pagne Ardenne, Campus Sciences Exactes et Naturelles, BP 1039, Reims 51687, France Francesca Ferri Universit`a degli Studi di Padova, Centro di Ateneo di Studi e Attivit`a Spaziali "Giuseppe Colombo" (CISAS), via Venezia 15, 35131 Padova, Italy Leigh Fletcher Department of Physics & Astronomy, Uni- versity of Leicester, University Road, Leicester, LE1 7RH, UK Thierry Fouchet LESIA, Observatoire de Paris, PSL Re- search University, CNRS, Sorbonne Universit´es, UPMC Univ. Paris 06, Univ. Paris Diderot, Sorbonne Paris Cit´e, 5 place Jules Janssen, 92195 Meudon, France In Situ Exploration of the Giant Planets -- 26/26 Tristan Guillot Observatoire de la Cote d'Azur, Laboratoire Lagrange, BP 4229, 06304 Nice cedex 4, France Paul Hartogh Max-Planck-Institut fur Sonnensystemforschung, Justus von Liebig Weg 3, 37077 Gottingen, Germany Ricardo Hueso Escuela de Ingenier´ıa de Bilbao, UPV/EHU, 48013 Bilbao, Spain Jet Propulsion Laboratory, California In- Mark Hofstadter stitute of Technology, 4800 Oak Grove Dr., Pasadena, CA, 91109, USA Jean-Pierre Lebreton CNRS-Universit´e d'Orl´eans, 3a Av- enue de la Recherche Scientifique, 45071 Orl´eans Cedex 2, France Kathleen E. Mandt Applied Physics Laboratory, Johns Hop- kins University, 11100 Johns Hopkins Rd., Laurel, MD 20723, USA Institut fur Planetenforschung, DLR, Berlin; Heike Rauer Zentrum fur Astronomie und Astrophysik, TU Berlin, Berlin Pascal Rannou Groupe de Spectrom´etrie Mol´eculaire et At- mosph´erique, UMR 7331, CNRS, Universit´e de Reims Cham- pagne Ardenne, Campus Sciences Exactes et Naturelles, BP 1039, Reims 51687, France Jean-Baptiste Renard CNRS-Universit´e d'Orl´eans, 3a Av- enue de la Recherche Scientifique, 45071 Orl´eans Cedex 2, France Agustin S´anchez-Lavega Escuela de Ingenier´ıa de Bilbao, UPV/EHU, 48013 Bilbao, Spain Kunio Sayanagi Department of Atmospheric and Planetary Sciences, Hampton University, 23 Tyler Street, Hampton, VA 23668, USA Amy Simon NASA Goddard Space flight Center, Greenbelt, MD, 20771, USA Thomas Spilker Solar System Science & Exploration, Mon- rovia, USA Ethiraj Venkatapathy NASA Ames Research Center, Mof- fett field, CA, USA J. Hunter Waite Southwest Research Institute, San Antonio, TX, 78228, USA Peter Wurz Space Science & Planetology, Physics Institute, University of Bern, Sidlerstrasse 5, 3012 Bern, Switzerland Cover illustration by courtesy of Tibor Balint
1802.05823
1
1802
2018-02-16T02:35:15
Single Transits and Eclipses Observed by K2
[ "astro-ph.EP", "astro-ph.SR" ]
Photometric survey data from the Kepler mission have been used to discover and characterize thousands of transiting exoplanet and eclipsing binary (EB) systems. These discoveries have enabled empirical studies of occurrence rates which reveal that exoplanets are ubiquitous and found in a wide variety of system architectures and physical compositions. Because the detection strategy of these missions is most sensitive to short orbital periods, the vast majority of these objects reside within 1 AU of their host star. Although other detection techniques have successfully identified exoplanets at wider orbits beyond the snow lines of their respective host stars (e.g., radial velocity, microlensing, direct imaging), occurrence rates within this population remain poorly constrained. As such, identifying long period objects (LPOs) from archival Kepler and K2 data is valuable from both a statistical and theoretical standpoint, particularly for massive gas giants which are thought to heavily influence the formation and evolution dynamics of their respective systems. Here we present a catalog of 164 single transit and eclipse candidates detected during a comprehensive survey of all currently available K2 data.
astro-ph.EP
astro-ph
Accepted to RNAAS Feb. 6, 2018 Single Transits and Eclipses Observed by K2 Daryll M. LaCourse1, Thomas L. Jacobs2 Subject headings: Planets and satellites: detection – surveys – catalogs – binaries: eclipsing Photometric survey data from the Kepler (Koch, et al. 2010) mission have been used to discover and characterize thousands of transiting exoplanet and eclipsing binary (EB) systems. These discoveries have enabled empirical studies of occurrence rates1 which reveal that exoplanets are ubiquitous and found in a wide variety of system architectures and physical compositions. Because the detection strategy of Kepler is most sensitive to short orbital periods, the vast majority of these objects reside within 1 AU of their host star. Although other detection techniques have successfully identified exoplanets at wider orbits beyond the snow lines of their respective host stars (e.g., radial velocity, microlensing, direct imaging), occurrence rates within this population remain poorly constrained. As such, identifying long period objects (LPOs) from archival Kepler and K2 (Howell, et al. 2014) data is valuable from both a statistical and theoretical standpoint, particularly for massive gas giants which are thought to heavily influence the formation and evolution dynamics of their respective systems. Here we present a catalog of 164 single transit and eclipse candidates detected during a comprehensive survey of all currently available K2 data. Identification of LPO candidates from Kepler data has been accomplished in several previous studies, which we briefly summarize here. Wang, et al. (2013) and Wang et al. (2015) leveraged a crowd sourcing strategy based on manual inspections performed by citizen-scientists of the Planet Hunters project (Fischer, et al. 2012), identifying 15 double-transit candidates and 17 single- transit candidates. Similarly, Uehara, et al. (2016) performed a visual inspection of 7557 Kepler Objects of Interest which recovered 28 single transits; including 7 objects consistent with Neptune- sized and Jupiter-sized exoplanets. Kipping, et al. (2014) and Kipping et al. (2015) utilized an automated pipeline to identify and validate a Uranus-sized exoplanet orbiting Kepler-421 and a Jupiter-sized exoplanet orbiting Kepler-167. Finally, Foreman-Mackey, et al. (2016) utilized a pipeline equipped with a probabilistic model comparison function to systematically search the light curves of 39,036 G and K dwarfs observed by Kepler, recovering 3 double-transit candidates and 13 single-transit candidates. An important result from the Foreman-Mackey study was to show an estimated occurrence rate for G/K dwarfs of 2.00(+/-0.7) exoplanets smaller than Jupiter in the 2 to 25 year orbital period range. 17507 52nd Place NE Marysville, WA, 98270; [email protected], USA 212812 SE 69th Place Bellevue, WA 98006; [email protected], USA 1For a review of occurence rate studies see: https://exoplanetarchive.ipac.caltech.edu/docs/occurrence_rate_papers.html – 2 – The search for LPOs has continued successfully into the K2 mission. Indeed, the first confirmed exoplanet of the mission, K2-2b, was identified from a single transit recovered during the mission's engineering test phase (Vanderburg, et al. 2015). Subsequently, an automated search was performed by Osborn, et al. (2016) for selected stars from Campaigns 1,2 and 3, which recovered 7 additional LPOs. Motivated by these results, we undertook a visual survey of 288,399 unique stellar light curves derived from K2 Campaigns 0 through 14. For Campaigns 0, 1 and 2 we downloaded Target Pixel Files from the MAST2 and extracted light curves in an automated fashion with Guest Observer software PyKE3(Still & Barclay 2012; Vincius, et al. 2017). For Campaigns 3 through 14, both the K2 Ames and K2SFF light curves were made available at the MAST and both sets of data were surveyed independenty by each author. For all data sets, candidates for which the LPO signal contaminated multiple EPIC targets or were noted to be associated with a spacecraft resaturation event, were discarded as false positives. Surviving targets are organized in Table 1 by Campaign and in order of decreasing brightness. Measurements for midpoint, depth, duration were taken with the light curve examination software LcTools4. Included stellar parameters are derived from the EPIC catalog (Huber, et al. 2016), except for Campaign 0 where we include Teff estimations from the K2-TESS catalog5(Stassun, et al. 2014). Based purely on an assessment of transit depth, it is tempting to conclude that many of the signals are caused by long period EBs. However we have opted to present the sample as an ensemble, without conclusively discriminating between stellar and sub-stellar companions. Although the radii of the LPOs can be inferred from a single transit, such estimations are only as accurate as our knowledge of the host star's parameters; for the vast majority of our sample such characterization has not been performed via asteroseismology or spectral fitting. An additional consideration is that large numbers of both late-type and evolved stars have been observed in K2, which can allow scenarios where an exoplanet transit mimics an EB, or vice versa. As such, we do not provide putative radius and period estimations here, save to note that if the EPIC parameters are assumed to be wholly accurate, 38 of the LPOs seem consistent with a planetary origin. Additional photometric or RV follow up is encouraged to properly elucidate the nature of these candidates. Note Added in Review: After submission of this research letter, we became aware of a large number of additional unpublished single transits in Hugh Osborns PhD thesis6. We are gratified to see a significant overlap between the two samples. 2https://archive.stsci.edu/kepler 3http://pyke.keplerscience.org 4https://sites.google.com/a/lctools.net/lctools/home 5https://filtergraph.com/tess k2campaigns 6https://warwick.ac.uk/fac/sci/physics/research/astro/publications/phd msc/hughosborn.phd.pdf – 3 – We are grateful to Andrew Vanderburg, Saul Rappaport, Joeseph Schmitt, Fei Dai and Ben Montet for helpful discussions and comments during the preparation of this catalog. We thank LcTools author Allan R. Schmitt for facilitating this research by making his software available for our survey. This letter includes data from the K2 mission, which is funded by the NASA Science Mission directorate. Much of the data presented was obtained from the Mikulski Archive for Space Telescopes (MAST), which is governed by STScI and operated by the Association of Universities for Research in Astronomy Inc., under NASA contract NAS526555. We thank all past and present members of the Kepler & K2 teams for their efforts to ensure the fantastic success of both missions. Facilities: Kepler/K2 References Auvergne, M., Bodin, P., Boisnard, L., et al. 2009, A&A, 506, 411 Fischer, D. A., Schwamb, M. E., Schawinski, K., et al. 2012, MNRAS, 419, 2900 Foreman-Mackey, D., Morton, T. D., Hogg, D. W., Agol, E. & Scholkopf, B., 2016, arXiv:1607.08237 Huber D., et al., 2016, ApJS, 224, 2 Koch D. G., et al., 2010, ApJ, 713, L79 Kipping, D. M., Torres, G., Buchhave, L. A., et al. 2014b, ApJ, 795, 25 Kipping, D. M., Torres, G., Henze, C., et al. 2016, ApJ, 820, 112 Osborn, H. P., Armstrong, D. J., Brown, D. J. A., et al. 2016, MNRAS, 457, 2273 Stassun, K. G., Pepper, J. A., Paegert, M., De Lee, N., & Sanchis-Ojeda, R. 2014, arXiv:1410.6379 Still, M., & Barclay, T. 2012, Astrophysics Source Code Library, ascl:1208.004 Vincius, Z., Barentsen, G., Hedges, et al. 2018, Zenodo, doi:10.5281/zenodo.835583 Uehara, S., Kawahara, H., Masuda, K., Yamada, S., Aizawa, M. 2016, ApJ, 822, 2 Vanderburg A., Johnson J. A., 2014, PASP, 126, 948 Wang, J., Fischer, D. A., Barclay, T., et al. 2013, ApJ, 776, 10 Wang, J., Fischer, D. A., Barclay, T., et al. 2015, ApJ, 815, 127 Table 1. Single Transits and Eclipses Observed by K2 EPIC Campaign KP (mag) Teff (K) 248811085 248555345 248854690 248847494 201854636 248749087 248657359 248607265 201792207 248912804 248873506 247692298 246732631 247756662 247762843 246849982 246696483 247865067 247795097 247594337 247967714 247450113 246906371 247814074 247015294 246194159 246359551 245929407 245964933 246394998 246089124 246361128 246163364 246301164 14 14 14 14 14 14 14 14 14 14 14 13 13 13 13 13 13 13 13 13 13 13 13 13 13 12 12 12 12 12 12 12 12 12 11.25 11.60 11.98 12.17 12.19 13.80 13.92 15.14 15.35 16.59 17.30 9.71 10.77 11.80 11.96 12.52 12.72 12.94 12.99 13.63 13.75 14.04 14.55 14.64 15.86 9.48 11.14 11.64 12.18 12.52 13.05 14.38 14.89 14.97 6327 5192 5462 4931 4936 4812 4790 4383 4101 3334 3917 10398 8224 4202 6527 9715 6149 5013 3606 4908 5076 5886 4912 6245 6267 6340 4696 - 5480 5877 5865 4717 4681 5063 R* R⊙ 1.53 6.07 2.58 0.79 3.98 0.65 0.70 0.56 0.48 0.17 0.35 3.20 2.06 35.34 1.39 1.85 1.72 8.33 166.45 6.90 2.01 1.86 5.74 1.76 1.26 1.81 5.33 - 0.88 1.12 1.10 0.68 0.73 0.78 BJD0 (BJD-2454833) Depth (ppm) Duration (hours) 3115.3795 3106.6452 3117.3517 3134.1158 3118.2002 3094.5298 3093.0281 3091.0972 3105.8900 3126.7501 3108.1574 3006.2056 3003.3757 3049.91875 3002.60955 3061.15675 3022.59145 3042.9614 3001.9245 2997.6037 3012.0791 3003.5390 2994.9265 3050.1637 2997.8078 2914.3878 2914.3570 2927.1473 2925.2779 2913.9893 2941.4699 2911.2000 2972.8118 2925.7785 91337 125978 84973 2236 30324 1833 19297 19639 465848 94658 114537 188904 137259 28528 1300 235239 32984 2266 192036 175064 258337 3363 24957 443535 174621 378934 204068 22554 40201 438858 138529 45500 69921 131436 6.3746 15.6912 25.4987 56.3914 16.1819 5.8824 14.2224 5.8844 3.9229 9.8064 11.7696 10.2960 21.5760 310.8888 5.3952 26.4792 17.1624 8.3352 12.7488 12.2592 4.9032 14.7120 7.3536 11.7696 9.3168 8.82665 21.0858 6.374717 10.7879 5.3940 9.3167 6.8651 34.8155 4.4132 Comments Deep Deep; secondary eclipse Deep Deep; secondary eclipse Flat-bottom Deep Deep Deep Deep Deep Deep; secondary eclipse Deep; secondary eclipse Deep; secondary eclipse Deep; secondary eclipse Deep; secondary eclipse Deep; secondary eclipse; flat-bottom Deep – 4 – EPIC Campaign KP (mag) 246425172 246302531 246020606 246331715 246279882 230887315 224937601 223332021 226600397 232334247 231674968 236157481 231312005 235099239 240410914 240323152 242209485 224703312 229021605 201132684 201092629 201479221 228786343 228804202 229022237 201093731 228891397 201510813 201496916 220315458 220186865 220562610 220606084 220152847 12 12 12 12 12 11 11 11 11 11 11 11 11 11 11 11 11 11 10 10 10 10 10 10 10 10 10 10 10 8 8 8 8 8 15.16 15.56 15.69 16.95 17.52 8.47 10.47 10.74 10.85 11.01 11.28 11.87 11.94 13.21 13.44 14.61 14.67 15.15 10.47 11.67 11.85 11.91 12.72 13.45 13.60 13.74 13.76 14.19 17.72 11.59 12.00 12.51 13.00 13.19 Teff (K) 4866 4146 5274 5429 5158 5262 9696 - 6744 6566 6073 7525 5854 4046 4799 8545 6229 5509 4895 5549 5259 5697 6140 5719 5168 6134 5088 5739 3833 5832 5158 5915 5761 5129 R* R⊙ 4.64 0.49 0.81 0.81 6.79 2.74 5.30 - 3.58 1.59 3.37 3.09 1.60 0.47 40.26 3.39 2.98 2.98 6.00 0.87 0.76 0.98 1.19 0.94 1.84 1.27 6.36 0.93 0.33 1.05 0.87 1.40 1.50 0.80 Table 1-Continued BJD0 (BJD-2454833) Depth (ppm) Duration (hours) Comments Deep; secondary eclipse; flat-bottom Deep; secondary eclipse; flat-bottom Deep Deep; secondary eclipse Deep; secondary eclipse Deep Deep Deep; secondary eclipse Apparent multi-planet system Deep Deep Deep Deep Deep 2939.1922 2914.9906 2939.7230 2927.6786 2939.5191 2827.3093 2869.8795 2887.6860 2862.2176 2831.1303 2839.8854 2880.6772 2852.5122 2856.5988 2837.1685 2841.3162 2861.1554 2882.4858 2785.7140 2797.8904 2778.0310 2787.3679 2814.7579 2790.4029 2751.6236 2801.8443 2780.2586 2802.0174 2799.5966 2611.4965 2571.6956 2563.6864 2618.2492 2568.8657 62953 63915 34977 57229 65549 30968 202073 72754 53438 1998 10914 116832 4908 2200 589868 12643 39727 146172 17416 2196 1126 159819 14966 731 21467 35729 246193 149321 221704 985 1973 2813 4928 9191 67.6696 4.4133 5.8842 6.3747 53.9395 8.3352 44.1312 17.1624 43.6440 18.1435 46.0941 22.0662 8.8272 14.7096 10.7880 5.8848 11.7672 11.7672 14.2204 12.7512 5.8843 10.7879 4.4132 3.9240 9.3170 6.3747 44.6227 5.8843 10.2975 4.9056 5.8848 11.7672 8.3352 12.2592 – 5 – Table 1-Continued EPIC Campaign KP (mag) 220605820 220515668 220208795 215067200 216831785 213832800 218751675 215307988 213332545 218187050 214783208 215894766 213867148 219240689 212549089 212820423 212542155 212685467 212694013 212555615 212805678 212325089 212343520 212477236 212554009 212332380 212775521 212732378 212715204 212152316 211633458 211598816 211924561 212026444 8 8 8 7 7 7 7 7 7 7 7 7 7 7 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 5 5 5 5 5 13.76 14.38 14.25 5.92 11.08 11.24 11.44 11.55 11.79 12.29 12.86 13.06 13.14 14.12 11.50 11.95 12.52 13.26 13.36 13.42 13.69 13.69 13.80 13.96 14.56 14.56 14.65 14.66 16.67 10.31 10.72 10.75 10.75 10.81 Teff (K) 5089 5385 5064 9210 6251 5326 - 7426 5892 5379 4206 5267 5150 4953 5986 4820 5800 5714 4913 5606 4986 4983 5100 5023 5834 5213 5252 5272 4193 5631 4830 7437 6976 5938 R* R⊙ 3.47 0.82 0.75 4.32 1.34 2.77 - - 1.16 8.74 38.41 2.36 16.29 8.97 1.33 5.10 0.98 0.91 5.03 2.11 4.27 4.85 0.88 3.25 0.96 0.86 2.28 0.65 0.47 2.29 10.55 2.19 1.17 1.18 BJD0 (BJD-2454833) Depth (ppm) Duration (hours) Comments 2590.8710 2599.4419 2575.9658 2493.8114 2505.0904 2488.4585 2502.2907 2489.3981 2505.5805 2512.0469 2493.7603 2501.4119 2533.8786 2532.6731 2412.0360 2433.3465 2389.6836 2409.6665 2409.1754 2428.6783 2460.5206 2441.2127 2451.0103 2449.5901 2462.2474 2449.8354 2424.7963 2396.7732 2399.3379 2374.2193 2368.6513 2363.3709 2346.8923 2343.3680 13884 152383 225648 151215 406 81263 3592 410977 108103 2216 7104 26419 8750 25884 53688 6805 1132 9319 7113 2358 60463 115890 80532 4389 141582 227278 12666 58214 185073 66344 1197 9718 2144 644 6.3744 4.9032 6.8664 41.6806 23.0472 8.3361 11.2782 17.1626 4.4132 14.7107 25.9896 6.3746 98.5620 42.6611 14.2224 22.0656 8.3352 7.3560 40.6992 3.9240 23.0472 53.9376 6.8664 15.2016 25.4976 14.2201 6.8640 6.3768 4.9032 30.4032 9.3168 3.9240 3.4320 25.4976 Deep Deep Flat-bottom Deep Deep; secondary eclipse Deep; secondary eclipse Flat-bottom – 6 – Secondary eclipse Flat-bottom Secondary eclipse Deep Deep; secondary eclipse Flat-bottom Table 1-Continued EPIC Campaign KP (mag) 211953574 211351543 211939692 211939692 211892898 211498244 212070574 211351097 212011230 211821192 211840710 211894612 211490542 211995462 212012030 211503363 211390677 211411112 211489484 210725198 211087003 210857749 211064647 210823406 210760314 210825751 210843533 211075893 205966706 206253908 205936222 206008070 203914123 204546592 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 4 4 4 4 4 4 4 4 4 3 3 3 3 2 2 11.31 11.38 11.75 11.75 11.77 11.88 12.10 12.32 12.37 12.58 12.72 12.76 12.90 12.96 13.07 13.22 13.31 13.40 15.99 10.39 11.64 12.70 14.91 15.74 16.20 16.43 16.94 17.72 9.86 11.18 12.09 13.82 9.03 9.93 Teff (K) 6027 6251 6404 6404 4787 3731 6039 6214 4856 5769 5162 6063 6134 5771 5127 4988 4886 6026 3767 7899 6072 6392 5613 4102 4470 3865 3566 4431 6200 6864 5581 4983 4801 4848 R* R⊙ 1.54 1.66 1.44 1.44 8.74 0.31 0.91 1.42 4.15 0.95 6.88 1.31 1.37 0.97 0.81 4.34 4.86 1.25 0.342 2.31 1.25 1.46 0.74 0.49 0.58 0.39 0.24 0.57 2.73 1.75 2.00 4.88 11.11 10.92 BJD0 (BJD-2454833) Depth (ppm) Duration (hours) Comments 2360.5608 2373.7598 2337.5035 2359.9068 2356.3519 2321.5460 2376.5789 2328.8308 2332.4882 2344.1743 2353.1026 2349.9258 2343.1733 2370.2761 2341.4668 2364.5142 2318.7983 2345.9526 2359.6206 2258.1411 2251.0917 2251.9701 2273.9654 2257.3639 2248.1500 2275.9263 2256.1794 2253.7886 2184.7830 2150.6618 2152.7766 2172.0842 2113.2841 2066.965 590 491 1204 710 8984 94154 3843 905 5003 1103 3890 44324 80800 256342 202281 4616 6780 26062 109099 103537 8089 184683 13344 174979 274496 10907 64606 163591 119208 100778 325665 8626 26917 10212 6.3744 6.3744 11.7696 9.3168 62.2756 6.8664 5.3928 7.3560 6.3744 7.3553 7.3560 8.8272 10.2960 11.7696 4.9032 23.5368 39.7195 4.4112 4.4112 18.6336 7.8456 13.7304 6.3744 5.8848 8.8272 15.2016 9.8064 11.7696 7.8457 6.8651 6.3748 97.5815 75.5344 20.5968 Apparent multi-planet system Apparent multi-planet system Deep; secondary eclipse Deep Deep Higher SnR in Ames aperture Deep; secondary eclipse Deep; secondary eclipse Apparent multi-planet system – 7 – Deep Deep Deep Deep Flat-bottom Noted in Osborn, et al. (2016) EPIC Campaign 203865172 203746451 204918110 203311200 204533587 204634789 204086428 203011840 204952800 205272592 204100531 204776782 201207683 201775904 201631267 201176672 201720401 201663371 201635132 202071902 202126877 202072917 202060921 202137580 202067195 202135247 202073476 202085278 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 0 0 0 0 0 0 0 0 0 KP (mag) 10.67 11.50 11.80 11.89 12.24 12.29 12.31 12.48 12.79 14.51 14.58 14.77 10.64 11.57 12.78 13.98 14.74 15.05 15.14 10.40 11.00 11.90 12.00 13.10 14.30 14.40 15.00 15.50 Teff (K) 6199 5627 6442 6787 5717 6029 6149 5063 6328 4544 6384 7247 5293 6103 5012 4542 5024 4176 3983 6194 7142 9221 7226 3755 6980 3919 4066 3676 R* R⊙ 1.90 0.93 2.03 1.94 0.91 2.02 1.43 1.91 1.22 12.57 1.80 1.95 9.15 1.58 2.64 0.50 3.79 0.52 0.44 - - - - - - - - - Table 1-Continued BJD0 (BJD-2454833) Depth (ppm) Duration (hours) Comments 2082.5250 2114.0612 2111.8748 2121.0070 2093.4343 2088.0105 2066.6689 2080.2971 2110.6789 2102.3539 2098.7673 2127.0038 2002.3225 2002.5464 1996.6831 2044.7690 1983.3515 1979.7246 1993.9755 1958.5707 1955.4341 1953.6261 1971.3294 1962.6458 1955.7918 1963.8721 1957.7729 1948.2010 6043 5550 112761 3656 1350 7814 2160 475216 428538 338711 1436692 267132 236704 14162 6207 31203 23941 48800 31550 91224 61834 248395 105768 66511 15372 89100 47716 8248 2.4504 11.7672 22.5552 13.7304 9.3168 2.9424 6.3744 19.1256 6.3744 12.7488 16.1808 14.2201 7.3554 5.3939 7.3554 16.6723 36.2876 6.3748 5.8843 5.3952 5.8848 7.3536 17.8456 50.9976 8.3376 40.2096 56.3904 19.6128 Deep; flat-bottom Noted in Osborn, et al. (2016) Noted in Osborn, et al. (2016) Deep Deep Deep Deep Deep Deep; Noted in Schmitt, et al. (2016) Noted in Osborn, et al. (2016) Noted in Osborn, et al. (2016) Noted in Osborn, et al. (2016) Deep Deep Deep; secondary eclipse Flat-bottom – 8 –
1604.03007
1
1604
2016-03-30T16:52:31
Keplerian integrals, elimination theory and identification of very short arcs in a large database of optical observations
[ "astro-ph.EP" ]
The modern optical telescopes produce a huge number of asteroid observations, that are grouped into very short arcs (VSAs), each containing a few observations of the same object in one single night. To decide whether two VSAs, collected in different nights, refer to the same observed object we can attempt to compute an orbit with the observations of both arcs: this is called the linkage problem. Since the number of orbit computations to be performed is very large, we need efficient methods of orbit determination. Using the first integrals of Kepler's motion we can write algebraic equations for the linkage problem, which can be put in polynomial form. The equations introduced in (Gronchi et al. 2015) can be reduced to a univariate polynomial of degree 9: the unknown is the topocentric distance $\rho$ of the observed body at the mean epoch of one of the VSAs. Using elimination theory we show an optimal property of this polynomial: it has the least degree among the univariate polynomials in the same variable that are consequence of the algebraic conservation laws and are obtained without squaring operations, that can be used to bring these algebraic equations in polynomial form. In this paper we also introduce a procedure to join three VSAs belonging to different nights: from the conservation of angular momentum at the three mean epochs of the VSAs, we obtain a univariate polynomial equation of degree 8 in the topocentric distance $\rho_2$ at the intermediate epoch. This algorithm has the same computational complexity as the classical method by Gauss, but uses more information, therefore we expect that it can produce more accurate results. For both methods, linking two and three VSAs, we also discuss how to select the solutions, making use of the full two-body dynamics, and show some numerical tests comparing the results with the ones obtained by Gauss' method.
astro-ph.EP
astro-ph
Keplerian integrals, elimination theory and identification of very short arcs in a large database of optical observations Giovanni F. Gronchi ∗, Giulio Ba`u †, Andrea Milani ‡ September 18, 2018 Abstract The modern optical telescopes produce a huge number of asteroid observations, that are grouped into very short arcs (VSAs), each containing a few observations of the same object in one single night. To decide whether two VSAs, collected in different nights, refer to the same observed object we can attempt to compute an orbit with the observations of both arcs: this is called the linkage problem. Since the number of orbit computations to be performed is very large, we need efficient methods of orbit determination. Using the first integrals of Kepler's motion we can write algebraic equations for the linkage problem, which can be put in polynomial form, see [3], [4], [5]. The equations introduced in [5] can be reduced to a univariate polynomial of degree 9: the unknown is the topocentric distance ρ of the observed body at the mean epoch of one of the VSAs. Using elimination theory we show an optimal property of this polynomial: it has the least degree among the univariate polynomials in the same variable that are consequence of the algebraic conservation laws and are obtained without squaring operations, that can be used to bring these algebraic In this paper we also introduce a procedure to join three equations in polynomial form. VSAs belonging to different nights: from the conservation of angular momentum at the three mean epochs of the VSAs, we obtain a univariate polynomial equation of degree 8 in the topocentric distance ρ2 at the intermediate epoch. This algorithm has the same computational complexity as the classical method by Gauss, but uses more information, therefore we expect that it can produce more accurate results. These results can be used as better preliminary orbits to compute a least squares orbital solution with three VSAs. For both methods, linking two and three VSAs, we also discuss how to select the solutions, making use of the full two-body dynamics, and show some numerical tests comparing the results with the ones obtained by Gauss' method. 1 Introduction We consider very short arcs (VSAs) of optical observations of a solar system body whose motion is dominated by the gravitational attraction of the Sun. These small sets of observations are called tracklets and the corresponding arc described in the sky is usually too short to compute a least squares orbit. In each observing night we can detect thousands of these data, so that it is difficult to decide whether two such arcs, collected in different nights, refer to the same observed ∗Dip. di Matematica, Univ. di Pisa, Italy [email protected] †Dip. di Matematica, Univ. di Pisa, Italy [email protected] ‡Dip. di Matematica, Univ. di Pisa, Italy [email protected] 1 body. This gives rise to an identification problem, that can be solved by computing an orbit with the information contained in two or more tracklets. Using the classical methods of initial orbit determination, those by Laplace [6] or Gauss [2], we usually cannot compute a preliminary orbit with three observations belonging to the same VSA because they are too close in time and the arc is usually too short. Even using observations taken from two different VSAs it may be difficult to compute an orbit. Laplace's or Gauss' methods in most cases work well if we use three different observations from three VSAs. In this case to compute a preliminary orbit we have to find the roots of a univariate polynomial of degree 8 (see [9]), that correspond to the possible values of the radial distance (geocentric for Laplace, topocentric for Gauss) of the observed body at a given epoch (the mean epoch of the h=1 th/3 for Laplace, the central epoch t2 for Gauss). observations P3 Assume for simplicity that we deal with this identification problem using the observations made by a single telescope performing an asteroid survey, like Pan-STARRS [10], or the next generation telescope LSST [7]. The average number of observations per night is N ≈ 104 for Pan-STARRS, and presumably we shall have N ≈ 105 for LSST. To perform systematically the identification by Gauss' method using the data of three observing nights we should test compat- ibility for O(N 3) triples of observations. This is clearly a cumbersome task. The identification of two VSAs is usually called the linkage problem, and it has been recently studied in [3], [4], [5] using the first integrals of the two-body motion. In [5] the authors introduced a univariate polynomial equation of degree 9 for the linkage problem, which is comparable with the equation of Gauss' method. This equation is derived in a concise way in Section 3. Moreover, we discuss an optimal property of such polynomial. Using algebraic elimination theory, we show that it has the least degree among the univariate polynomials that are consequence of the algebraic conservation laws of Kepler's problem, provided we drop the dependence between the inverse of the heliocentric distance 1/r appearing in the Keplerian potential and the topocentric distance ρ. This approach avoids the squaring operations needed in [3], [4] to bring the selected equations1 into a polynomial form. In Section 3.4 we sketch a method to check the validity of the identification and select solutions according to some compatibility conditions, similar to the ones in [3], that use the full two-body dynamics. An orbit computed with two VSAs is usually not as reliable as one computed with three observations, each picked up in a different VSAs, because the latter usually represents a longer arc. To obtain more reliable results we have to join together at least three VSAs. In Section 4 we introduce a univariate polynomial equation of degree 8 to link three VSAs of optical observations by means of the conservation of angular momentum only. Then the other laws of Kepler's motion can be used to set up restrictive compatibility conditions, allowing us to test the identification and select solutions. Assume we set up an identification procedure with a large database of asteroid observations. For simplicity, we can consider three observing nights, in which we collect O(N ) VSAs of ob- servations per night. We can try to identify pairs of VSAs belonging to the first two nights by applying O(N 2) times the linkage algorithm introduced in [5] and reviewed in Section 3. The output is composed by preliminary orbits obtained with pairs of VSAs. If the thresholds in the controls for acceptance (see Section 3.4) are well selected, we do not obtain more than O(N ) pairs of VSAs, in fact the number of different objects observed in the two nights is O(N ). Then we can apply the method to link three VSAs introduced in Section 4 to the O(N ) selected pairs and the O(N ) VSAs of the third observing night. We conclude that this identification problem can be faced by O(N 2) computations of roots of a polynomial of degree 9 or 8, instead of O(N 3) computations of roots of Gauss' polynomial. 1In these papers not all the algebraic conservation laws are used. 2 2 Keplerian integrals We consider the Keplerian motion of a celestial body around a center of force, set at the origin of a given reference system, that in the asteroid case corresponds to the center of the Sun. Optical observations of the body are made by a telescope whose heliocentric position is a known function of time. Then the heliocentric position and velocity of the body are given by r = ρeρ + q, r = ρeρ + ρη + q, (1) where q, q are the heliocentric position and velocity of the observer, ρ, ρ are the topocentric radial distance and velocity, eρ is the line of sight unit vector, which can be written in terms of the topocentric right ascension α and declination δ as Moreover in (1) we set where eρ = (cos δ cos α, cos δ sin α, sin δ). η = α cos δeα + δeδ, eα = (cos δ)−1 ∂eρ ∂α , eδ = ∂eρ ∂δ , and α, δ are the angular rates. The Keplerian integrals, represented by the angular momentum vector c, the Laplace-Lenz vector L and the energy E, are defined by 1 2r2 − E = c = r × r, µL = (cid:16)r2 − (2) µ r(cid:17)r − (r · r)r, , µ r as functions of r, r. Given the values of α, δ, α, δ, they can be written as algebraic functions of ρ, ρ using relations (1). 3 Linking two VSAs Given a very short arc of optical observations (αi, δi), i = 1 . . . m, made by the same station at ti different times, it is often possible to compute the attributable vector (see [11]) A = (α, δ, α, δ) m Pm at the mean epoch ¯t = 1 i=1 ti. The missing quantities to obtain a preliminary orbit are the ρ at t = ¯t. When the second derivatives (α, δ) are either topocentric distance and velocity ρ, not available (if m = 2), or not accurate enough due to the errors in the observations, then the attributable summarizes essentially all the information contained in the VSA. In this case a preliminary orbit can be obtained by linking together two different VSAs. The key idea of the linkage method is to use the conservation of the Keplerian integrals c, L, E at the two mean epochs ¯t1, ¯t2 of two attributables A1,A2: c1 = c2, L1 = L2, E1 = E2, (3) where the indexes 1, 2 refer to the epoch. Below we derive in a concise way the polynomial equations for the linkage problem introduced in [5], and we review the procedure to obtain the univariate polynomial of degree 9 giving the possible values for the topocentric distance ρ2. Moreover, we show here an optimal property of this polynomial. 3 3.1 Conservation of angular momentum The angular momentum as function of ρ, ρ can be written as where c(ρ, ρ) = r × r = D ρ + Eρ2 + Fρ + G, D = q × eρ, E = α cos δeρ × eα + δeρ × eδ = α cos δeδ − δeα, F = α cos δq × eα + δq × eδ + eρ × q, G = q × q. Then the equation c1 = c2, representing the conservation of the angular momentum are written as where D1 ρ1 − D2 ρ2 = J(ρ1, ρ2), (4) (5) We can eliminate the radial velocities ρ1, ρ2 by making the scalar product with D1 × D2, that gives the quadratic polynomial 1 + F2ρ2 − F1ρ1 + G2 − G1. J(ρ1, ρ2) = E2ρ2 2 − E1ρ2 q(ρ1, ρ2) := D1 × D2 · J(ρ1, ρ2) = 0 (6) in the variables ρ1, ρ2. The radial velocities are given by ρ1(ρ1, ρ2) = (J × D2) · (D1 × D2) D1 × D22 , ρ2(ρ1, ρ2) = (J × D1) · (D1 × D2) D1 × D22 . (7) These expressions are obtained by projecting (4) onto the vectors D1 × (D1 × D2) and D2 × (D1 × D2), generating the plane orthogonal to D1 × D2. Therefore using such expressions of ρ1, ρ2 we have (c1 − c2) × (D1 × D2) = 0 whatever the values of ρ1, ρ2. 3.2 The univariate polynomial u By relations (7) we can eliminate the dependence on ρ1, ρ2 in the Laplace-Lenz and energy conservation laws (8) These are algebraic equations in ρ1, ρ2 that are not polynomial because of the terms 1/r1, 1/r2. However, in the equation E1 = E2. L1 = L2, ξ := (cid:2)µ(L1 − L2) − (E1r1 − E2r2)(cid:3) × (r1 − r2) = 0, which is a consequence of (8), the terms 1/r1, 1/r2 cancel out. The monomials of ξ with the highest total degree, i.e. 6, are all parallel to eρ 2, so that we consider the bivariate polynomials (9) (10) 1 × eρ p2 = ξ · eρ 2 p1 = ξ · eρ 1, 4 having total degree 5. system In [5] the authors show that the overdetermined bivariate polynomial q = 0, ξ = 0 is consistent, i.e. its set of solutions in C2 is not empty, and is equivalent to Moreover, if we consider the resultants (see [1]) q = p1 = p2 = 0. u1 = Res(p1, q, ρ1), u2 = Res(p2, q, ρ1), then the greatest common divisor of u1 and u2, u = gcd(u1, u2), (11) is a univariate polynomial in the variable ρ2 of degree 9. Remark 1 Since in this problem the role of ρ1 and ρ2 is symmetric, for a generic choice of the data Aj, qj, qj , j = 1, 2, we obtain an analogous result by eliminating the variable ρ2, instead of ρ1, from p1, p2. We also recall the construction used in [5] to compute uj, j = 1, 2. We can write q(ρ1, ρ2) = 2 Xh=0 bh(ρ2)ρh 1 , where b0(ρ2) = q0,2ρ2 2 + q0,1ρ2 + q0,0, b1 = q1,0, b2 = q2,0, with the coefficients qh,k depending only on the data Aj, qj , qj, j = 1, 2. Moreover, we have p1(ρ1, ρ2) = 4 Xh=0 a1,h(ρ2)ρh 1 , p2(ρ1, ρ2) = 5 Xh=0 a2,h(ρ2)ρh 1 , (12) for some univariate polynomials ak,h whose degrees are described by the upper small circles used to construct Newton's polygons of p1, p2 in Figure 1. Assume q2,0, q0,2 6= 0. From q = 0 we obtain ρh 1 = βhρ1 + γh, h = 2, 3, 4, 5, (13) where and β2 = − b1 b2 , γ2 = − b0 b2 , βh+1 = βhβ2 + γh, γh+1 = βhγ2, h = 2, 3, 4. Inserting (13) into (12) we obtain pj(ρ1, ρ2) = aj,1(ρ2)ρ1 + aj,0(ρ2), j = 1, 2, (14) 5 6 5 4 3 2 1 0 2 ρ P 1 P 1 6 5 4 3 2 1 0 2 ρ P 2 P 2 0 1 2 3 ρ 1 4 5 6 0 1 2 3 ρ 1 4 5 6 Figure 1: We draw Newton's polygons Pj, Pj for the polynomials pj, pj, j = 1, 2. In this figure the polygons are overlapping: the nodes with circles correspond to the (multi-index) exponents of the monomials in pj; the nodes with asterisks correspond to the exponents of the monomials in pj. where a1,1 = a1,1 + a2,1 = a2,1 + 4 Xh=2 Xh=2 5 a1,hβh, a1,0 = a1,0 + a2,hβh, a2,0 = a2,0 + 4 Xh=2 Xh=2 5 a1,hγh, a2,hγh. (15) (16) In this case the nodes with asterisks In Figure 1 we also draw Newton's polygons of p1, p2. correspond to the exponents of the monomials in pj and the upper asterisks describe the degrees of the polynomials ak,h. Let us introduce the polynomials v1 = Res(p1, q, ρ1), v2 = Res(p2, q, ρ1). We can show the following result. Lemma 1 By the properties of resultants we find that Proof. We prove the first relation; the proof of the second one is similar. We have u1 = q3 2,0v1, u2 = q4 2,0v2. (17) u1 = Res(p1, q, ρ1) = det a10 0 a11 a10 a12 a11 a13 a12 a14 a13 0 a14   b0 b1 b2 0 0 0 0 b0 b1 b2 0 0 0 0 b0 b1 b2 0 0 0 0 b0 b1 b2 .   6 By performing raw operations and by the properties of determinants we obtain Res(p1, q, ρ1) = det = det     a10 a11 + γ2a13 + β2γ2a14 a12 + β2a13 + (β2 2 + γ2)a14 0 a10 a11 a13 + β2a14 a12 + β2a13 + β3a14 a14 0 a13 + β2a14 a14 b0 b1 b2 0 0 0 0 0 0 b2 0 0 0 0 0 0 b2 0 0 0 0 0 0 b2   = a10 a11 0 0 a10 a11 a13 + β2a14 a12 + β2a13 + β3a14 a14 0 a13 + β2a14 a14 b0 b1 b2 0 0 0 0 0 0 b2 0 0 0 0 0 0 b2 0 0 0 0 0 0 b2   = b3 2Res(p1, q, ρ1). The last matrix is obtained from the previous one by adding to its first column a suitable multiple of the third column. 3.3 An optimal property of the polynomial u If we consider the auxiliary variable u together with the polynomial relation u2r2 = µ2, (cid:3) (18) then the Keplerian integrals introduced in (2) can be viewed as polynomials in the variables ρ, ρ, u. In particular, we obtain L = (r2 − u)r − (r · r)r, E = 1 2r2 − u. We observe that, for all ρ, ρ, u, c · L = 0, µ2L2 = u2r2 + 2Ec2; (19) the second relation generalizes the classical formula relating eccentricity, energy and angular momentum. The full polynomial system c1 = c2, µL1 = µL2, E1 = E2, u2 jrj2 = µ2 (j = 1, 2), (20) with unknowns ρ1, ρ2, ρ1, ρ2, u1, u2, is generically not consistent, see Corollary 2 at the end of this section. Next we show that, if we drop the dependence between uj and ρj given by relation (18), we obtain a consistent polynomial system, and the univariate polynomial u of degree 9 introduced in [5] has the least degree among the polynomials in ρ2 that are a consequence of the polynomials in (3). Therefore u has the least degree among the polynomials in ρ2, consequences of the algebraic Keplerian integrals and obtained without squaring operations, which can be used to bring the algebraic conservation laws in polynomial form. Let I ⊆ R[ρ1, ρ2, ρ1, ρ2, u1, u2] 7 be the ideal of the polynomial ring in the variables ρ1, ρ2, ρ1, ρ2, u1, u2, with real coefficients, generated by the seven polynomials c1 − c2, µL1 − µL2, E1 − E2, where we write uj in place of µ/rj for j = 1, 2. We recall that a set {g1, . . . , gn}, with n ∈ N, is a Groebner basis of a polynomial ideal I for a fixed monomial order ≻ if and only if the leading term (for that order) of any element of I is divisible by the leading term of one gj, see [1]. The main result of this section is the following. Theorem 1 For a generic choice of the data Aj, qj , qj, j = 1, 2, we can find a set of polynomials g1 . . . g6 ∈ R[ρ1, ρ2, ρ1, ρ2, u1, u2], which is a Groebner basis of the ideal I for the lexicographic order ρ1 ≻ ρ2 ≻ u1 ≻ u2 ≻ ρ1 ≻ ρ2, (21) such that where u is the polynomial defined in (11). g6 = u, Proof. Assuming D1 × D2 6= 0, eρ 1 × eρ 2 6= 0, we consider the following set of generators of the ideal I: q1 = (c1 − c2) · D1 × D2, q2 = (c1 − c2) · D1 × (D1 × D2), q3 = (c1 − c2) · D2 × (D1 × D2), q4 = µ(L1 − L2) · eρ q5 = µ(L1 − L2) · D1, q6 = µ(L1 − L2) · D2, q7 = E1 − E2. 1 × eρ 2, The first three polynomials have the form q1 = q, q2 = D1 × D22 ρ1 − J · D1 × (D1 × D2), q3 = D1 × D22 ρ2 − J · D2 × (D1 × D2), with q = q(ρ1, ρ2), J = J(ρ1, ρ2) defined in (6), (5) respectively. The other generators can be written as 2)u1 − (D2 · eρ 1)u2 + f4, q4 = −(D1 · eρ q5 = (D1 · r2)u2 + f5, q6 = −(D2 · r1)u1 + f6, q7 = −u1 + u2 + f7, 8 for some polynomials fj = fj(ρ1, ρ2, ρ1, ρ2), j = 4 . . . 7. Set A = D1 · eρ 2 + D2 · eρ 1 = (q1 − q2) · eρ 1 × eρ 2. Assuming the three terms do not vanish, we can substitute the generators q4 . . . q7 with the polynomials A, D2 · eρ 1, D1 · eρ 2 2)q7 − q4 = Au2 + a1, 1)q7 − q4 = Au1 + a2, p4 = (D1 · eρ p5 = −(D2 · eρ p6 = (D1 · r2)p4 − Aq5, p7 = (D2 · r1)p5 + Aq6, where a1 = (D1 · eρ 2) f7 − f4, a2 = −(D2 · eρ 1) f7 − f4. We observe that, using relations q2 = q3 = 0 to eliminate ρ1, ρ2 from p6, p7, we obtain where p1, p2 are the bivariate polynomials defined in (10). Since we can write p6 = (−D1 · eρ 2)p1, p7 = (D2 · eρ 1)p2, p6 = p6 + b2q2 + b3q3, p7 = p7 + c2q2 + c3q3 for some polynomials bj, cj , j = 2, 3 in the variables ρ1, ρ2, ρ1, ρ2, we have p6, p7 ∈ I. Let us consider the elimination ideal J = hq1, p6, p7i = hq, p1, p2i, in R[ρ1, ρ2]. The ideal with the polynomials pj defined in (14), coincides with J, in fact J = hq, p1, p2i, for some polynomials dj = dj(ρ1, ρ2). In particular, we have pj = pj + djq, j = 1, 2 V (J) = V ( J), where the variety V (K) of a polynomial ideal K ∈ R[ρ1, ρ2] is the set V (K) = {(ρ1, ρ2) ∈ C2 : p(ρ1, ρ2) = 0, ∀p ∈ K}. The ideal fulfills so that Indeed, we shall show that Let us introduce the polynomial J1 = hp1, p2i, J1 ⊆ J, V ( J1) ⊇ V ( J). V ( J1) = V ( J). We need the following results. v := Res(p1, p2, ρ1) = a1,1a2,0 − a1,0a2,1. 9 (22) (23) Lemma 2 For a generic choice of the data Aj, qj , qj, j = 1, 2, the polynomials u, v have 9 distinct solutions in C. Proof. We show this property for u; the proof for v is analogous. Let u(ρ2) = 9 Xj=0 cjρj 2, for some coefficients cj ∈ R depending on the data. First we show that, for a generic choice of the data, the rank of the Jacobian matrix ∂(c0, . . . , c9) ∂(A1,A2, q1, q1, q2, q2) is maximal, that is equal to 10. To check this property it suffices to show that this rank is maximal for a particular choice of the data. In fact, if the rank were < 10 in an open set, then by the analytic dependence of the coefficients cj on the data the rank would not be maximal at any point. We made this check using the symbolic computation software Maple 18 with the following data: A1 = (cid:16)2 arctan q1 = (1, 0, 0), 1 2 , 0, 1, 1(cid:17), A2 = (cid:16)2 arctan 1 2 , 2 arctan 1 2 q1 = (0, 1, 0), q2 = (0, 1, 0), , 1, 1(cid:17), q2 = (−1, 0, 0). Moreover, by a well known property of polynomials, we know that u is square-free (i.e. without multiple roots) for a generic choice of the coefficients cj. This fact, together with the maximal rank property showed above, concludes the proof of the lemma. Lemma 3 For a generic choice of the data Aj, qj , qj, j = 1, 2, we have gcd(a1,1, a2,1) = 1, where a1,1, a2,1 are the univariate polynomials defined in (15), (16). Proof. We give a proof similar to the one of Lemma 2. Let a11(ρ2) = 3 Xj=0 c1,jρj 2, a21(ρ2) = 4 Xj=0 c2,jρj 2, (cid:3) (24) for some coefficients ci,j depending on the data. We can show that the Jacobian matrix ∂(c1,0, . . . , c1,3, c2,0, . . . , c2,4) ∂(A1,A2, q1, q1, q2, q2) has generically maximal rank, i.e. 9, by checking that the rank is maximal for the data of Lemma 2. To conclude we use the fact that for a generic choice of the coefficients ci,j relation (24) holds true. By Lemma 3 we can find two univariate polynomials β, γ in the variable ρ2 such that βa1,1 + γa2,1 = 1. 10 (cid:3) (25) Let us introduce where w := β p1 + γ p2 = ρ1 + z(ρ2), (26) Lemma 4 The polynomial ideal is equal to J1. z = βa1,0 + γa2,0. J2 = hw, vi Proof. From the definition of w and from relation (27) we have J2 ⊆ J1. On the other hand, we can easily invert relations (26), (27) and, using (25), we obtain v = a1,1 p2 − a2,1 p1 p1 = a1,1w + γv, so that the other inclusion J1 ⊆ J2 holds true. p2 = a2,1w − βv, (cid:3) Lemmata 2, 4 imply that V ( J1) has 9 distinct points. In fact, from w = 0, for each root ρ2 of v we find a unique ρ1 such that (ρ1, ρ2) ∈ V ( J1). On the other hand, since J = J we have V ( J) = V (J) and generically V (J) has 9 distinct points too. We can prove it by using Theorem 1 in [5] and Lemma 2 for the polynomial u. Then from (23) we conclude that V ( J1) = V ( J). (28) In particular, the polynomials v and u coincide up to a constant factor. obtain Now we prove that J1 is indeed equal to J. Let us take h ∈ J. Making the division by w we (29) h(ρ1, ρ2) = h1(ρ1, ρ2)(cid:0)ρ1 + z(ρ2)(cid:1) + r(ρ2) for some polynomials h1, r. The remainder r depends only on ρ2 because w is linear in ρ1. From (22) and (29) we have that r ∈ J. Using relation (28) and the fact that u is generically squarefree we obtain that together with (29) implies that h ∈ J1. We conclude that u r, J1 = J. The polynomials g1 . . . g6, with g1 = q2, g2 = q3, g3 = p4, g4 = p5, g5 = w, g6 = u, form a Groebner basis of the ideal I for the lexicographic order (21). To show this we can simply check that the leading monomials of each pair (gi, gj), with 1 ≤ i < j ≤ 6, are relatively prime (see [1], Chapter 2). This concludes the proof of the theorem. (cid:3) From the definition of Groebner basis we immediately obtain the following Corollary 1 The polynomial u has the least degree among the univariate polynomials in the variable ρ2 belonging to the ideal I. 11 As a consequence of the computations in the proof of Theorem 1 we also obtain Corollary 2 The polynomial system (20) is generically not consistent. The same result holds true by removing from (20) only one of the two equations u2 jrj2 = µ2, j = 1, 2. Proof. We show that the system gj = 0, j = 1 . . . 6, u2 2r22 − µ2 = 0 (30) is generically not consistent, where gj are the polynomials in the statement of Theorem 1. By using equations g1 = g2 = g3 = g5 = 0 we can obtain from u2 2r22 = µ2 another univariate polynomial, say u in the variable ρ2. Then u and u have a common root in C (i.e. are compatible) if and only if (31) Assume there is an open set in the space of the data Aj , qj, qj, j = 1, 2 such that equation (31) holds. Since the left-hand side of (31) is an analytic function of the data, then this equation holds on the whole data set. Therefore, to conclude it is enough to check that equations (30) are not compatible for a particular choice of the data, e.g. as in Lemma 2. Res(u, u, ρ2) = 0. In a similar way we can prove that the system is generically not consistent. gj = 0, j = 1 . . . 6, u2 1r12 − µ2 = 0 (cid:3) 3.4 Compatibility conditions and covariance of the solutions In this section we discuss how to discard some of the solutions computed with the method described in Section 3 on the base of the full two-body dynamics. Given a pair of attributables A = (A1,A2) at epochs ¯t1, ¯t2 with covariance matrices ΓA1 , ΓA2 , we call R = (ρ1, ρ1, ρ2, ρ2) one of the solutions of the equation (32) Φ(R; A) = 0, with where Φ(R; A) = (cid:18) c1 − c2 Ξ · eρ 1 (cid:19) , Ξ = 1 2 (r22 − r12)r1 × r2 − (r1 · r1)r1 × (r1 − r2) + (r2 · r2)r2 × (r1 − r2), which corresponds to the vector ξ defined in (9) if we eliminate ρ1, ρ2 by (7). We can repeat what follows for each solution of Φ(R; A) = 0. The notation is similar to [3]. Let us introduce the difference vector where ∆a,ℓ = (∆a, ∆ℓ), ∆a = a1 − a2, where n(a) = √µa−3/2 is the mean motion and ti = ¯ti − ρi/c, i = 1, 2. Note that here we consider the difference of the two mean anomalies at the same epoch t1 in a way that it is a smooth function at each integer multiple of 2π. We introduce the map ∆ℓ = (cid:2)ℓ1 −(cid:0)ℓ2 + n(a2)(t1 − t2)(cid:1) + π(mod 2π)(cid:3) − π, (A1,A2) = A 7→ Ψ(A) = (A1,R1, ∆a,ℓ) , 12 giving the orbit (A1,R1) in attributables coordinates at epoch t1 together with the vector ∆a,ℓ which is not constrained by equation (32). By the covariance propagation rule we have ΓΨ(A) = ∂Ψ ∂A ∂A(cid:21)T ΓA (cid:20) ∂Ψ , (33) where ∂Ψ ∂A =   I ∂R1 ∂A1 ∂∆a,ℓ ∂A1 0 ∂R1 ∂A2 ∂∆a,ℓ ∂A2   and ΓA = (cid:20) ΓA1 0 0 ΓA2 (cid:21) . We can check if there is any solution of (32) fulfilling the compatibility conditions within a threshold defined by the covariance matrix of the attributables ΓA. From (33) we can compute the marginal covariance of the vector ∆a,ℓ: ∆a,ℓ = 0 Γ∆a,ℓ = ∂∆a,ℓ ∂A ∂A (cid:21)T ΓA(cid:20) ∂∆a,ℓ . The inverse matrix C∆a,ℓ = Γ−1 an identification between the attributables A1,A2: we check whether defines a norm k·k⋆ in the (∆a, ∆ℓ) plane, allowing us to test ∆a,ℓ k∆a,ℓk2 ⋆ = ∆a,ℓC∆a,ℓ ∆T a,ℓ ≤ χ2 max, where χmax is a control parameter. If a preliminary orbit (A1,R1) is accepted, from (33) we can also compute its marginal covariance as the 6 × 6 matrix Γ(A1,R1) = (cid:20) ΓA1 ΓR1,A1 ΓA1,R1 ΓR1 (cid:21) , where ∂A1(cid:21)T ΓA1,R1 = ΓA1 (cid:20) ∂R1 , ΓR1 = ∂R1 ∂A ∂A (cid:21)T ΓA(cid:20) ∂R1 , ΓR1,A1 = ΓT A1,R1. 4 Linking three VSAs Here we introduce a method to compute preliminary orbits from three VSAs using the Keplerian integrals (2). In this case the conservation of the angular momentum at the three epochs is enough to obtain a finite number of solutions of the identification problem. In the following the indexes 1, 2, 3 will refer to the mean epochs ¯tj of three VSAs with attributables Aj . We consider the equations: (34) c2 = c3, c1 = c2, c3 = c1, that can be written as D1 ρ1 − D2 ρ2 = J12(ρ1, ρ2), D2 ρ2 − D3 ρ3 = J23(ρ2, ρ3), D3 ρ3 − D1 ρ1 = J31(ρ3, ρ1), 13 where J12(ρ1, ρ2) = E2ρ2 J23(ρ2, ρ3) = E3ρ2 J31(ρ3, ρ1) = E1ρ2 2 − E1ρ2 3 − E2ρ2 1 − E3ρ2 1 + F2ρ2 − F1ρ1 + G2 − G1, 2 + F3ρ3 − F2ρ2 + G3 − G2, 3 + F1ρ1 − F3ρ3 + G1 − G3. Equations (34) are redundant, that is, if two of them hold true then the third equation is also fulfilled. We consider the following projections of equations (34): (c1 − c2) · D1 × D2 = 0, (c1 − c2) · D1 × (D1 × D2) = 0, (c2 − c3) · D2 × D3 = 0, (c2 − c3) · D2 × (D2 × D3) = 0, (c3 − c1) · D3 × D1 = 0, (c3 − c1) · D3 × (D3 × D1) = 0. D1 × D2 · D3 6= 0. Proposition 1 Assume (35) (36) (37) (38) (39) (40) (41) Then the system of equations (35) -- (40) is equivalent to (34). Proof. Assuming that (39), (40) are fulfilled, to prove that c3 = c1 we only need to show that the projection of this equation onto a vector v, such that D3× D1, D3× (D3× D1), v are linearly independent, holds true. We denote by Π12 = hD1 × D2, D1 × (D1 × D2)i, Π23 = hD2 × D3, D2 × (D2 × D3)i the planes passing through the origin generated by the vectors within the brackets. If relation (41) holds, then we have Π12 ∩ Π23 = hD1 × D2i, i.e. the intersection of the two planes is the straight line generated by the vector v = D1 × D2. Moreover, we have (D1 × D2) · (D3 × D1) ×(cid:0)D3 × (D3 × D1)(cid:1) = D3 × D12D1 × D2 · D3, that does not vanish by (41). Therefore, from (35) -- (38) we obtain (c1− c2)· v = (c2− c3)· v = 0, that yield (c3 − c1) · v = 0. In a similar way we can prove that c1 = c2, c2 = c3, provided (35) -- (40) hold. (cid:3) Equations (35), (37), (39) depend only on the radial distances. In fact, they correspond to the system J12 · D1 × D2 = 0, J23 · D2 × D3 = 0, J31 · D3 × D1 = 0, which can be written as q3 = a3ρ2 q1 = a1ρ2 q2 = a2ρ2 2 + b3ρ2 3 + b1ρ2 1 + b2ρ2 1 + c3ρ2 + d3ρ1 + e3 = 0, 2 + c1ρ3 + d1ρ2 + e1 = 0, 3 + c2ρ1 + d2ρ3 + e2 = 0, 14 (42) (43) (44) (45) where a3 = E2 · D1 × D2, c3 = F2 · D1 × D2, b3 = −E1 · D1 × D2, d3 = −F1 · D1 × D2, e3 = (G2 − G1) · D1 × D2, and the other coefficients aj, bj, cj, dj, ej, for j = 1, 2, have similar expressions, obtained by cycling the indexes. To eliminate ρ1, ρ3 from (42) we first compute the resultant which depends only on ρ2, ρ3. Then we compute the resultant r = Res(q3, q2, ρ1), q = Res(r, q1, ρ3), which is a univariate polynomial of degree 8 in the variable ρ2. Therefore, provided (41) holds, to get the solutions of (34) we search for the roots ¯ρ2 of q(ρ2), then we compute the corresponding values ¯ρ3 from system r(ρ3, ¯ρ2) = q1(ρ3, ¯ρ2) = 0, and finally the corresponding values ¯ρ1 from system q3(ρ1, ¯ρ2) = q2(¯ρ3, ρ1) = 0. Since the unknowns ρj represent distances we can discard triples (¯ρ1, ¯ρ2, ¯ρ3) where some ρj is non-positive. From equations (36), (38), (40) we can write the radial velocities ρj as functions of pairs of radial distances: ρ2 = ρ3 = ρ1 = J12(ρ1, ρ2) · D1 × (D1 × D2) J23(ρ2, ρ3) · D2 × (D2 × D3) J31(ρ3, ρ1) · D3 × (D3 × D1) D1 × D22 D2 × D32 D3 × D12 , , . Remark 2 As a simple criterion to discard triples (A1,A2,A3) before making the computation described in this section we can use the intersection criterion introduced in [5] to discard pairs of attributables. More precisely, we can apply this criterion three times, i.e. we check for each j = 1, 2, 3 whether the conic Qj, defined by qj = 0 (see equations (43), (44), (45)), intersects the square R = [ρmin, ρmax] × [ρmin, ρmax] for some fixed ρmax > ρmin > 0. If this criterion fails in one of these cases we discard the selected triple. For more details see the appendix in [5]. 4.1 Solutions with zero angular momentum A particular solution of system (34) can be obtained by searching for values of ρj, ρj such that cj(ρj, ρj) = 0, j = 1, 2, 3. Relation r × r = 0 implies that there exists λ ∈ R such that ρeρ + ρη + q = λ(ρeρ + q), with η = α cos δeα + δeδ. Setting σ = ρ − λρ we can write (46) as We introduce the vector σeρ + ρη − λq = − q. u = q − (q · eρ)eρ − 1 η2 (q · η)η, 15 (46) (47) which is orthogonal to both eρ, η, where η = η is called the proper motion. Thus, we can write (47) as [σ − λ(q · eρ)]eρ +hρ − λ η2 (q · η)iη − λu = − q. Since {eρ, η, u} is generically an orthogonal basis of R3, we find λ = 1 u2 ( q · u), ρ = 1 η2 (λq − q) · η, ρ = λρ + (λq − q) · eρ. In particular we obtain the value ρ = 1 η2(cid:16) 1 u2 ( q · u)(q · η) − q · η(cid:17) for the radial distance, corresponding to a solution with zero angular momentum. 4.2 Compatibility conditions and covariance of the solutions We discuss how to discard solutions of (34) in a way similar to Section 3.4. Given a triple of at- tributables A = (A1,A2,A3) with covariance matrices ΓA1 , ΓA2 , ΓA3, we call R = (ρ1, ρ1, ρ2, ρ2, ρ3, ρ3) one of the solutions of the equation Φ(R; A) = 0, (48) with (c1 − c2) · D1 × D2 (c2 − c3) · D2 × D3 (c3 − c1) · D3 × D1 We can repeat what follows for each solution of Φ(R; A) = 0. (c1 − c2) · D1 × (D1 × D2) (c2 − c3) · D2 × (D2 × D3) (c3 − c1) · D3 × (D3 × D1) Φ(R; A) =   .   Let us introduce the difference vectors ∆12 = (cid:0)a1 − a2, [ω1 − ω2 + π(mod 2π)] − π,(cid:2)ℓ1 −(cid:0)ℓ2 + n(a2)(t1 − t2)(cid:1) + π(mod 2π)(cid:3) − π(cid:1), ∆32 = (cid:0)a3 − a2, [ω3 − ω2 + π(mod 2π)] − π,(cid:2)ℓ3 −(cid:0)ℓ2 + n(a2)(t3 − t2)(cid:1) + π(mod 2π)(cid:3) − π(cid:1), where the third component is the difference of the two mean anomalies referring to epoch ti = ¯ti− ρi/c, and n(a) = √µa−3/2 is the mean motion. Here the difference of two angles is computed in a way that it is a smooth function at each integer multiple of 2π. We introduce the map (A1,A2,A3) = A 7→ Ψ(A) = (A2,R2, ∆12, ∆32) , giving the orbit (A2,R2(A)) in attributable coordinates at epoch t2 together with the vectors ∆12(A), ∆32(A), which are not constrained by the angular momentum integrals. We want to check if there is any solution of (48) fulfilling the compatibility conditions within a threshold defined by the covariance matrix of the attributables ∆12 = ∆32 = 0 ΓA =   ΓA1 0 0 0 ΓA2 0 0 0 ΓA3   . 16 By the covariance propagation rule we have where The matrices ∂R2 ∂Aj ΓΨ(A) = ∂Ψ ∂A ∂A(cid:21)T ΓA (cid:20) ∂Ψ , ∂Ψ ∂A =   0 ∂R2 ∂A1 ∂∆12 ∂A1 ∂∆32 ∂A1 I ∂R2 ∂A2 ∂∆12 ∂A2 ∂∆32 ∂A2 0 ∂R2 ∂A3 ∂∆12 ∂A3 ∂∆32 ∂A3 .   , j = 1, 2, 3, can be computed from the relation ∂R ∂A (A) = −(cid:20) ∂Φ ∂R (R(A), A)(cid:21)−1 ∂Φ ∂A (R(A), A). The marginal covariance matrix for the vector (∆12, ∆32) is given by the block Γ∆ = (cid:20) Γ∆12 Γ∆32,∆12 Γ∆12,∆32 Γ∆32 (cid:21) of ΓΨ(A), where Γ∆12 = Γ∆32 = ∂∆12 ∂A ∂∆32 ∂A ∂A (cid:21)T ΓA(cid:20) ∂∆12 ∂A (cid:21)T ΓA(cid:20) ∂∆32 , , Γ∆12,∆32 = ∂∆12 ∂A ∂A (cid:21)T ΓA(cid:20) ∂∆32 , Γ∆32,∆12 = ΓT ∆12,∆32 . The inverse matrix C∆ = Γ−1 ∆ defines a norm k·k⋆ in the six dimensional space with coordinates ∆ = (∆12, ∆32), allowing us to test an identification between the attributables A1,A2,A3: we check whether (49) max, k∆k2 ⋆ = ∆C∆∆T ≤ χ2 where χmax is a control parameter. For each orbit, solution of (48), fulfilling condition (49) we can also define a covariance matrix Γ2 for the attributable coordinates (A2,R2): Γ2 = (cid:20) ΓA2 ΓR2,A2 ΓA2,R2 ΓR2 (cid:21) , where ΓA2 is given and ∂A2(cid:21)T ΓA2,R2 = ΓA2 (cid:20) ∂R2 , ΓR2 = ∂R2 ∂A ∂A (cid:21)T ΓA(cid:20) ∂R2 , ΓR2,A2 = ΓT A2,R2. 17 tr 1 2 3 obs 1 2 3 4 1 2 3 4 1 2 3 4 α (rad) 3.834760347106644 3.834778963951999 3.834797653519405 3.834816924863230 3.717640827594138 3.717559015285451 3.717476330312138 3.717393936227034 3.369239074975656 3.369201695840843 3.369164534872187 3.369126864849164 δ (rad) -7.983116606074819E-02 -7.982534829657487E-02 -7.981962749513778E-02 -7.981410061917313E-02 4.346887946196211E-03 4.378546279572663E-03 4.410543982525893E-03 4.442250797270456E-03 7.801563578772919E-02 7.800763636199089E-02 7.800021871266992E-02 7.799251017514028E-02 epoch (MJD) 57052.58743759259 57052.59951759259 57052.61160759259 57052.62370759259 57102.52326759259 57102.53596759259 57102.54885759259 57102.56162759259 57163.27290759259 57163.28720759259 57163.30153759259 57163.31588759259 Table 1: The three selected tracklets, each composed by four observations of asteroid (154229). 5 Numerical tests In this section we compare the preliminary orbits obtained by Gauss' method and the methods described in Sections 3, 4 for a test case: the near-Earth asteroid (154229). In Table 1 we list three tracklets, each composed by four observations (right ascension, declination), of this asteroid collected with the Pan-STARRS telescope. In Table 2 we show the approximated values of the components of the three attributables com- puted from the tracklets in Table 1. att α (rad) 1 3.83479 3.71752 2 3 3.36918 δ (rad) -7.98225E-02 4.39460E-03 7.80039E-02 α (rad/s) 1.55849E-03 -6.43398E-03 -2.60900E-03 δ (rad/s) 4.70783E-04 2.48563E-03 -5.36020E-04 epoch (MJD) 57052.60557 57102.54243 57163.29439 Table 2: Attributables computed from the three tracklets in Table 1. To compare the preliminary orbits we use two least squares solutions: one is computed with tracklets 1, 2 only, the other with all the tracklets. In Table 3 we list these solutions together with the preliminary orbits computed by the different methods. The labels G2, G3 refer to the orbits obtained with Gauss' method using different observations from Table 1: for G2 we use observations 1, 4 of tracklet 1 and observation 1 of tracklet 2; for G3 we use observation 1 of each tracklet. The labels L2, L3 refer to the methods described in Sections 3, 4. For L2 we use attributables 1, 2 listed in Table 2; for L3 we use all the attributables in this table. The labels LS2, LS3 refer to the least squares orbits computed from G2, G3 respectively. For LS2 we use the observations of tracklets 1, 2 only, for LS3 we use all the observations in Table 1. Let EG2, EL2, EG3, EL3 be the preliminary orbits computed with the different methods. Moreover, let ELS2, ELS3 be the least squares orbits corresponding to the different sets of data employed, and let ΓLS2, ΓLS3 be the related covariance matrices. All the preliminary orbits are propagated to the mean epoch of the arc of observations used to compute the least squares solution ELS2 or ELS3, according to the index 2 or 3. Then we consider the normal matrices CLS2 = Γ−1 LS3 corresponding to ELS2, ELS3. The norms displayed LS2, CLS3 = Γ−1 18 epoch 57077.57400 57077.57400 57077.57400 57106.14746 57106.14746 57106.14746 57106.14746 a (au) 1.85046 1.85384 1.84903 1.88095 1.84725 1.85112 1.84899 e 0.71629 0.71913 0.71930 0.73082 0.72153 0.71865 0.71930 I 10.00603 10.11799 10.09292 10.02343 10.17272 10.07393 10.09304 Ω 66.70400 67.29283 67.65173 67.97447 67.25235 67.70983 67.65083 ω 343.12690 341.93359 341.39098 341.61797 341.51657 341.48650 341.39054 ℓ 59.78415 61.35804 61.73660 69.37321 73.17327 72.68650 72.93972 norm 4639.4 521.3 // 5882.0 775.9 // 206660.0 G2 L2 LS2 G3 L3 LS3 LS2 Table 3: Preliminary orbits obtained with Gauss' method and with the linkage methods described in Sections 3, 4. The angles are given in degrees. The values of the norms defined in (50), (51) are listed in the last column. in Table 3 are defined as where EG2 = ∆G2 · CLS2∆G2, EL2 = ∆L2 · CLS2∆L2 , EG3 = ∆G3 · CLS3∆G3, EL3 = ∆L3 · CLS3∆L3 , (50) ∆G2 = EG2 − ELS2, ∆L2 = EL2 − ELS2, ∆G3 = EG3 − ELS3, ∆L3 = EL3 − ELS3. The orbit ELS2 in the last raw of Table 3 is the least squares solution obtained with tracklets 1, 2 and propagated at the mean epoch of the three tracklets. The corresponding norm is given by ELS2 = ∆LS2 · CLS3 ∆LS2, (51) where ∆LS2 = ELS2 − ELS3. From the values of the norms in this test case we conclude that EL3 is better than EG3, because it is closer to the least squares orbit ELS3. We also observe that EL2 is better than EG2. However, the value of the norm in the last raw in Table 3 implies that ELS2 is not close to ELS3. Both EL3 and EG3 are much better, as preliminary orbits, than the orbit obtained by propagating ELS2 to the mean epoch of the three tracklets. These results are consistent with the large scale test performed in [8, Fig. 2], showing that least squares solutions with two VSAs are very poor approximations of the true orbit, while least squares solutions with three VSAs are accurate enough. This implies the need to compute from scratch a preliminary orbit when we join a third tracklet to a pair of linked VSAs. The role of the linkage of two VSAs is to test the compatibility of pairs of tracklets and discard a large number of them. The orbit computed with two VSAs should not be used for the attribution (see [11]) of a third tracklet. Acknowledgments We wish to thank M. Caboara, P. Gianni, E. Sbarra, B. Trager, who gave us very useful sugges- tions on the algebraic aspects of this work. This work is partially supported by the Marie Curie Initial Training Network Stardust, FP7-PEOPLE-2012-ITN, Grant Agreement 317185. 19 References [1] Cox, D., Little, J., O'Shea, D.: Ideals, Varieties, and Algorithms, Springer (2005) [2] Gauss, C. F.: Theory of the Motion of the Heavenly Bodies Moving about the Sun in Conic Sections (1809), reprinted by Dover publications (1963) [3] Gronchi, G. F., Dimare, L., Milani, A.: Orbit determination with the two-body integrals, CMDA 107/3, 299-318 (2010) [4] Gronchi, G. F., Farnocchia, D., Dimare, L.: Orbit determination with the two-body integrals. II, CMDA 110/3, 257-270 (2011) [5] Gronchi, G. F., Ba`u, G., Mar`o, S.: Orbit determination with the two-body integrals. III, CMDA 123/2, 105-122 (2015) [6] Laplace, P. S., M´em. Acad. R. Sci. Paris, in Laplace's collected works, 10, 93-146 (1780) [7] Large Synoptic Survey Telescope, http://www.lsst.org/ [8] Milani, A., Gronchi, G. F., Knezevi´c, Z.: New Definition of Discovery for Solar System Objects, EMP 100/1-2, 83-116 (2007) [9] Milani, A., Gronchi, G. F.: The theory of orbit determination, Cambridge University Press (2010) [10] Panoramic Survey Telescope & Rapid Response System, http://pan-starrs.ifa.hawaii.edu/public/ [11] Milani, A., Sansaturio, M. E., Chesley, S. R.: The Asteroid Identification Problem IV: Attributions, Icarus 151, 150-159 (2001) 20
1208.2957
1
1208
2012-08-14T19:58:53
Statistical Study of the Early Solar System's Instability with 4, 5 and 6 Giant Planets
[ "astro-ph.EP" ]
Several properties of the Solar System, including the wide radial spacing and orbital eccentricities of giant planets, can be explained if the early Solar System evolved through a dynamical instability followed by migration of planets in the planetesimal disk. Here we report the results of a statistical study, in which we performed nearly 10^4 numerical simulations of planetary instability starting from hundreds of different initial conditions. We found that the dynamical evolution is typically too violent, if Jupiter and Saturn start in the 3:2 resonance, leading to ejection of at least one ice giant from the Solar System. Planet ejection can be avoided if the mass of the transplanetary disk of planetesimals was large (M_disk>50 M_Earth), but we found that a massive disk would lead to excessive dynamical damping (e.g., final e_55 < 0.01 compared to present e_55=0.044, where e_55 is the amplitude of the fifth eccentric mode in the Jupiter's orbit), and to smooth migration that violates constraints from the survival of the terrestrial planets. Better results were obtained when the Solar System was assumed to have five giant planets initially and one ice giant, with the mass comparable to that of Uranus and Neptune, was ejected into interstellar space by Jupiter. The best results were obtained when the ejected planet was placed into the external 3:2 or 4:3 resonance with Saturn and M_disk ~ 20 M_Earth. The range of possible outcomes is rather broad in this case, indicating that the present Solar System is neither a typical nor expected result for a given initial state, and occurs, in best cases, with only a ~5% probability (as defined by the success criteria described in the main text). The case with six giant planets shows interesting dynamics but does offer significant advantages relative to the five planet case.
astro-ph.EP
astro-ph
STATISTICAL STUDY OF THE EARLY SOLAR SYSTEM’S INSTABILITY WITH 4, 5 AND 6 GIANT PLANETS David Nesvorn´y1,2 and Alessandro Morbidelli2 (1) Department of Space Studies, Southwest Research Institute, Boulder, CO 80302, USA (2) D´epartement Cassiop´ee, University of Nice, CNRS, Observatoire de la Cote d’Azur, Nice, 06304, France ABSTRACT Several properties of the Solar System, including the wide radial spacing and orbital eccentricities of giant planets, can be explained if the early Solar System evolved through a dynamical instability followed by migration of planets in the planetesimal disk. Here we report the results of a statistical study, in which we performed nearly 104 numerical simulations of planetary instability starting from hundreds of different initial conditions. We found that the dynamical evolution is typically too violent, if Jupiter and Saturn start in the 3:2 resonance, leading to ejection of at least one ice giant from the Solar System. Planet ejection can be avoided if the mass of the transplanetary disk of planetesimals was large (Mdisk & 50 MEarth ), but we found that a massive disk would lead to excessive dynamical damping (e.g., final e55 . 0.01 compared to present e55 = 0.044, where e55 is the amplitude of the fifth eccentric mode in the Jupiter’s orbit), and to smooth migration that violates constraints from the survival of the terrestrial planets. Better results were obtained when the Solar System was assumed to have five giant planets initially and one ice giant, with the mass comparable to that of Uranus and Neptune, was ejected into interstellar space by Jupiter. The best results were obtained when the ejected planet was placed into the external 3:2 or 4:3 resonance with Saturn and Mdisk ≃ 20 MEarth . The range of possible outcomes is rather broad in this case, indicating that the present Solar System is neither a typical nor expected result for a given initial state, and occurs, in best cases, with only a ≃5% probability (as defined by the success criteria described in the main text). The case with six giant planets shows interesting dynamics but does offer significant advantages relative to the five planet case. – 2 – 1. Introduction As giant planets radially migrate in the protoplanetary nebula they should commonly be drawn into compact systems, in which the pairs of neighbor planets are locked in the orbital resonances (Kley 2000, Masset & Snellgrove 2001). The resonant planetary systems emerging from the protoplanetary disks can become dynamically unstable after the gas disappears, leading to a phase when planets scatter each other. This model can help to explain the observed resonant exoplanets (e.g., Gliese 876; Marcy et al. 2001), commonly large exoplanet eccentricities (Weidenschilling & Marzari 1996, Rasio & Ford 1996), and microlensing data that show evidence for a large number of planets that are free-floating in interstellar space (Sumi et al. 2011; but see Veras & Raymond 2012). The Solar System, with the widely spaced and nearly circular orbits of the giant planets, bears little resemblance to the bulk of known exoplanets. Yet, if our understanding of physics of planet–gas-disk interaction is correct, it is likely that the young Solar System followed the evolutionary path outlined above. Jupiter and Saturn, for example, were most likely trapped in the 3:2 resonance (Masset & Snellgrove 2001, Morbidelli & Crida 2007, Pierens & Nelson 2008, Pierens & Raymond 2011, Walsh et al. 2011), defined as PSat/PJup = 1.5, where PJup and PSat are the orbital periods of Jupiter and Saturn (this ratio is 2.49 today). To stretch to its current configuration, the outer Solar System most likely underwent a dynamical instability during which Uranus and Neptune were scattered off of Jupiter and Saturn and acquired eccentric orbits (Thommes et al. 1999, Tsiganis et al. 2005, Morbidelli et al. 2007, Levison et al. 2011). The orbits were subsequently stabilized (and circularized) by damping the excess orbital energy into a massive disk of planetesimals located beyond the orbit of the outermost planet (hereafter transplanetary disk), whose remains survived to this time in the Kuiper belt. Finally, as evidenced by dynamical structures observed in the Kuiper belt, planets radially migrated to their current orbits by scattering planetesimals (Malhotra et al. 1995, Gomes et al. 2004, Levison et al. 2008). Several versions of the Solar System instability were proposed. Thommes et al. (1999), motivated by the apparent inability of the existing formation models to accrete Uranus and Neptune at their present locations, proposed that Uranus and Neptune formed in the Jupiter-Saturn zone, and were scattered outwards when Jupiter, and perhaps Saturn, ac- creted nebular gas. This work represented an important paradigm shift in studies of the Solar System formation. Inspired by this work, Levison et al. (2001) suggested that the instability and subsequent dispersal of the planetesimal disk, if appropriately delayed (or due to the late formation of Uranus and Neptune; Wetherill 1975), could explain the Late Heavy Bombardment (LHB) of the Moon. The LHB was a spike in the bombardment rate some 4 Gyr ago suggested by the clustering of ages of several lunar basins. The nature of the – 3 – LHB, however, is still being debated (see Hartmann et al. 2000 and Chapman et al. 2007 for reviews). In the Nice version of the instability (Tsiganis et al. 2005), Uranus and Neptune were assumed to have formed just outside the orbit of Saturn, while Jupiter and Saturn were assumed to have initial orbits (where “initial” means at the time when the protoplanetary nebula dispersed), such that PSat/PJup < 2. The instability was triggered when Jupiter and Saturn, radially drifting by scattering planetesimals, approached and crossed the 2:1 resonance (PSat/PJup = 2). The subsequent evolution was similar to that found in the Thommes et al. model, but led to a better final orbital configuration of the planets (assuming that the disk contained ∼ 20-50 MEarth and was truncated at 30-35 AU). The Nice model is compelling because it explains the separations, eccentricities, and inclinations of the outer planets (Tsiganis et al. 2005), and many properties of the popula- tions of small bodies (Morbidelli et al. 2005, Nesvorn´y et al. 2007, Levison et al. 2008, 2009, Nesvorn´y & Vokrouhlick´y 2009). It is currently the only migration model that is consistent with the current dynamical structure of the terrestrial planets (Brasser et al. 2009) and the main asteroid belt (Morbidelli et al. 2010). Moreover, the Nice model can also reproduce the magnitude and duration of the LHB (Gomes et al. 2005, Bottke et al. 2012). Two potentially problematic issues of the original Nice model (hereafter ONM) were addressed by Morbidelli et al. (2007) and Levison et al. (2011). Morbidelli et al. pointed out that the initial planetary orbits in the ONM were chosen without the proper regard to the previous stage of evolution during which the giant planets interacted with the protoplanetary nebula. As shown by hydrodynamical studies (Masset & Snellgrove 2001, Morbidelli & Crida 2007, Pierens & Nelson 2008, Pierens & Raymond 2011), this interaction most likely led to the convergent migration of planets, and their trapping in orbital resonances. Morbidelli et al. (2007) studied the dynamical instability starting from the resonant planetary systems and showed that the orbit evolution of planets was similar to that of the ONM. To produce the LHB, the instability needs to occur late relative to the dispersal of the protoplanetary nebula. A protoplanetary nebula typically disperses in ∼3-10 Myr after the birth of the star (Haisch et al. 2001), which for the Sun was probably contemporary to the formation of the first Solar System solids, 4.568 Gyr ago (Bouvier et al. 2007, Burkhardt et al. 2008). The onset of the LHB, traditionally considered to be 3.9-4.0 Gyr ago (Ryder 2002), has been recently revised to 4.1-4.2 Gyr ago (Bottke et al. 2012). This means that the instability had to occur with a delay of 350-650 Myr, with the exact value depending on the actual time of the LHB event.1 Such a late instability can be triggered in the ONM if 1Here we opted for including the full range of previous and new estimates of the delay, because it is not – 4 – the inner edge of the planetesimal disk was close, but not too close, to the outer ice giant. If the edge had been too close, the instability would have happened too early to be related to the LHB. If the edge had been too far, the instability would have happened too late or would not have happened at all, because the planetesimals would had stayed radially confined and not evolved onto planet-crossing orbits (where they could influence planets by short-range interactions). To avoid the need for fine tuning of the inner disk’s edge in the ONM, Levison et al. (2011) proposed that the late instability was caused by the long-range interactions between planets and the radially-confined distant planetesimal disk. The long-term interactions arise when gravitational scattering between disk’s planetesimals is taken into account. As a result of these interactions, the planets and planetesimal disk slowly exchanged the energy and angular momentum until, after hundreds of Myr of small changes, the resonant locks between planets were broken and the instability reigned supreme. Note that the instability trigger in Levison et al. (broken resonant locks) is fundamentally different from that of the ONM (ma jor resonance crossing during migration). Here we report the results of a new statistical study of the Solar System instability. Sections 2-4 explain our method and constraints. The results are presented in Sect. 5. We discuss the plausible initial configurations of planets after the gas nebula dispersal, mass of the planetesimal disk, and effect of different trigger mechanisms on the results. The first steps in this direction were taken by Batygin & Brown (2010), who found that some initial resonant configurations may work while others probably do not. Following Nesvorn´y (2011) and Batygin et al. (2012) we also consider cases in which the young Solar System had extra ice giants initially and lost them during the scattering phase (these works and their relation to the present paper will be discussed in more detail below; see, e.g., Sect. 6). We extend the previous studies of the Solar System instability by: (1) exploring a wide range of initial parameters (new resonant chains, six planets, etc.), (2) improving the statistics with up to 100 simulations per case, (3) considering different trigger mechanisms (including the one recently proposed by Levison et al. 2011), and (4) applying strict criteria of success/failure to all studied cases (see Sect. 3 and 4). Our conclusions are given in Sect. 7. 100% guaranteed that the new estimate is correct. Imposing the 350-450 My delay, instead of 350-650 My, would not make much of a difference, because most planetary systems that are stable over 450 My are also stable over 650 My. – 5 – 2. Method We conducted computer simulations of the early evolution of the Solar System. In the first step (Phase 1), we performed hydrodynamic and N -body simulations to identify the resonant configurations that may have occurred among the young Solar System’s giant planets. Our hydrodynamic simulations used Fargo (Masset 2000) and followed the method described in Morbidelli et al. (2007). As the Fargo simulations are CPU expensive, we used these results as a guide, and generated many additional resonant systems with the N -body integrator known as SyMBA (Duncan et al. 1998). Planets with masses corresponding to those of Jupiter, Saturn and ice giants, ordered in increasing orbital distance from the Sun, were placed on initial orbits with the period ratios slightly larger that those of the selected resonances. We tested several cases for the initial radial order of ice giants including the case with Uranus on the inside of Neptune’s orbit, Uranus on the outside of the Neptune’s orbit, and case where both ice giants were given masses intermediate between that of Uranus and Neptune. The results (discussed in Sect. 5) obtained in these cases were statistically equivalent showing that the initial ordering of Uranus and Neptune was not important. The planets were then migrated into resonances with SyMBA modified to include forces that mimic the effects of gas. We considered cases with four, five and six initial planets, where the additional planets were placed onto resonant orbits between Saturn and the inner ice giant, or beyond the orbit of the outer ice giant. Additional planets were given the mass between 1/3 and 3 times the mass of Uranus. Different starting positions of planets, rates of the semima jor axis and eccentricity evo- lution (as implied by different gas disk densities), and timescales for the gas disk’s dispersal produced different results. For Jupiter and Saturn, we confined the scope of this study to the 3:2 and 2:1 resonances, because the former one is strongly preferred from previous hy- drodynamic studies (Masset & Snellgrove 2001, Morbidelli et al. 2007, Pierens & Nelson 2008, Walsh et al. 2011, Pierens & Raymond 2011). The 2:1 resonance was included for comparison. According to our tests with Fargo, trapping of Jupiter and Saturn in the 2:1 resonance may require special conditions (e.g., a low gas density implying slow convergent migration; Thommes et al. 2008). The 2:1 resonance does not generically lead to the out- ward migration of Jupiter and Saturn that is required for the Grand Tack model (Walsh et al. 2011). The planetary eccentricities, inclinations and resonant amplitudes obtained at the end of our Phase-1 simulations were e < 0.1, i < 0.2◦ and < 60◦ . The inner ice giant typically had the most eccentric orbit (0.05 < e < 0.1), while the other planets’ orbits were more circular – 6 – (e.g., Jupiter had e . 0.01 in most cases). We also considered cases with several different damping strengths for the same resonant chain. These cases helped us to understand the effect of the initial excitation on the evolution of orbits during the instability. Changing the damping strength may not be strictly justified, because the semima jor axis migration and eccentricity damping are expected to be coupled (Goldreich & Sari 2003). The exact nature of this coupling, however, depends on several disk parameters that are poorly constrained (e.g., Lee & Peale 2002). The instability of a planetary system can occur after the gas disk’s dispersal when the stabilizing effects of gas are removed. Such an instability can be triggered spontaneously (e.g., Weidenschilling & Marzari 1996), by divergent migration of planets produced by their interaction with planetesimals leaking into the planet-crossing orbits from a transplanetary disk (Tsiganis et al. 2005), or by long-range perturbations from a distant planetesimal disk (Levison et al. 2011). It is often assumed that the instability in the Solar System occurred at the time of the LHB. As discussed in the introduction this requires that the giant planets remained on their initial resonant orbits for ∼350-650 Myr. To allow for this possibility, we examined the resonant configurations identified above and selected those that were stable over hundreds of Myr, if considered in isolation (i.e., no disk, planets only, no external perturbations). Only the stable systems were used for the follow-up simulations, in which we tracked the evolution of planetary orbits through and past the instability (Phase 2). We included the effects of the transplanetary planetesimal disk in the Phase-2 simula- tions. The disk was represented by 1,000 equal-mass bodies2 that were placed into orbits with low orbital eccentricities and inclinations (0.01), and radial distances between rin < r < rout . We also performed a limited number of simulations with 100, 300, 3,000, and 10,000 disk bodies to test the effect of resolution,3 and with excited disks, where the eccentricities and inclinations of disk bodies were set to be ≃0.1 (Sect. 5.2.1). The surface density was fixed at Σ(r) = 1/r in all simulations, because our tests showed that considering different density profiles has only a minimal effect on the results (see also Batygin & Brown 2010). The outer 2Planets fully interact with each other and with the disk bodies. The gravitational interaction between disk bodies was neglected to cut down the CPU cost. This is justified because the behavior of the sys- tem during the instability is mainly driven by the interaction between planets, and between planets and planetesimals. 3Note that even our highest resolution runs did not have enough bodies to properly model the planetesimal disk, which is thought to have contained ∼1,000 Pluto-size and myriads of smaller planetesimals (Morbidelli et al. 2009a). This should not be a ma jor problem, however, because our results with increased resolution showed only a minor dependence on the number of bodies used. – 7 – edge of the disk was placed at rout = 30 AU, so that the planetesimal-driven migration is expected to park Neptune near its present semima jor axis (Gomes et al. 2004).4 We considered cases with rin = an + ∆, where an is the semima jor axis of the outer ice giant, n is the number of planets, and ∆ = 0.5-5 AU. The instability was usually triggered early for ∆ < 1 AU. To trigger the instability for ∆ > 1 AU, we broke the resonant locks by altering the mean anomaly of one of the ice giants. This was done at the start of Phase-2 simulations. This method was inspired by the results of Levison et al. (2011). Different cases are denoted by B in the following text, where B(j ) stands for breaking the resonant lock of ice giant planet j (e.g., B(1) corresponds to the innermost ice giant; B(0) is the ONM trigger). In either case discussed above, the scattering phase between planets started shortly after the beginning of our simulations, which guaranteed low CPU cost. We considered different masses of the planetesimal disk, Mdisk , with Mdisk between 10 and 100 MEarth . Thirty simulations were performed in each case, where different evolution histories were generated by randomly seeding the initial orbit distribution of planetesimals. The number of simulations was increased to 100 in the interesting cases. In total, we completed nearly 104 scattering simulations from over 200 different initial states. The new trials were selected with the knowledge of the results of the previous simulations and were optimized to sample the interesting parts of parameter space. Each system was followed for 100 Myr with the standard SyMBA integrator (Duncan et al. 1998), at which point the planetesimal disk was largely depleted and planetary migration ceased. We used a h = 0.5 yr time step and verified that the results were statistically the same with h = 0.25 yr. 3. Constraints We defined four criteria to measure the overall success of simulations. First of all, the final planetary system must have four giant planets (criterion A) with orbits that resemble the present ones (criterion B). Note that A means that one and two planets must be ejected in the five- and six-planet planet cases, while all four planets need to survive in the four-planet case. As for B, we claim success if the final mean semima jor axis of each planet is within 20% to its present value, and if the final mean eccentricities and mean inclinations are no larger than 0.11 and 2◦ , respectively. These thresholds were obtained by doubling the current mean 4The real planetesimal disk probably continued all the way to ≃47 AU, as implied by the existence of the cold classical Kuiper belt (Batygin et al. 2011), but this extension likely contained too little mass to drive Neptune’s migration. – 8 – eccentricity of Saturn (eSat = 0.054) and mean inclination of Uranus (iUra = 1.02◦ ; Table 1). For the successful runs, as defined above, we also checked on the history of encounters between giant planets, evolution of the secular g5 , g6 and s6 modes, and secular structure of the final planetary systems. To explain the observed populations of the irregular moons that are roughly similar at each planet (when corrected for observational incompleteness; Jewitt & Haghighipour 2007), all planets –including Jupiter– must participate in encounters with other planets (Nesvorn´y et al. 2007). Encounters of Jupiter and/or Saturn with ice giants may also be needed to excite the g5 mode in the Jupiter’s orbit to its current amplitude (e55 = 0.044; Morbidelli et al. 2009b). Tables 2-4 list the secular frequencies and amplitudes of the outer Solar System planets. It turns out that it is generally easy to have encounters of one of the ice giants to Jupiter if the planets start in a compact resonant configuration (such encounters occur in most simulations for most cases tested here). The amplitudes of g6 and s6 modes also do not pose a problem. It is much harder to excite e55 , however. We therefore opt for a restrictive criterion in which we require that e55 > 0.022 in the final systems, i.e., at least half of its current value (criterion C). The e55 amplitude was determined by following the final planetary systems for additional 10 Myr (without planetesimals), and Fourier-analyzing the results (Sidlichovsk´y & Nesvorn´y 1996). The evolution of secular modes, mainly g5 , g6 and s6 , is constrained from their effects on the terrestrial planets and asteroid belt. As giant planets scatter and migrate, these frequencies change. This may become a problem, if g5 slowly swipes over the g1 or g2 modes, because the strong g1 = g5 and g2 = g5 resonances can produce excessive excitation and instabilities in the terrestrial planet system (Brasser et al. 2009, Agnor & Lin 2011). The behavior of the g6 and s6 modes, on the other hand, is important for the asteroid belt (Morbidelli et al. 2010, Minton & Malhotra 2011). As g5 , g6 and s6 are mainly a function of the orbital separation between Jupiter and Saturn, the constraints from the terrestrial planets and asteroid belt can be conveniently defined in terms of PSat/PJup . This ratio needs to evolve from <2.1 to >2.3 in < 1 Myr (criterion D; also final PSat/PJup < 2.8; i.e., close to present PSat/PJup = 2.49 but not too restrictive; see Sect. 4), which can be achieved, for example, if planetary encounters with an ice giant scatter Jupiter inward and Saturn outward. The 1-Myr limit is conservative in that a slower evolution of secular frequencies, which fails to satisfy the criterion D, would clearly violate the constraints. Note also that in most simulations that satisfy criterion D, PSat/PJup evolves from <2.1 to >2.3 in much less than 1 Myr. Most simulations that satisfy D should therefore satisfy the constraints from the terrestrial planets and asteroids. – 9 – In summary, our constraints A and B express the basic requirements on a successful model. Constraint C is more restrictive in that it places a more precise condition on the dynamical structure of the final systems. Constraint D is the least rigorous. Given that it is related to the orbits of terrestrial planets, this constraint would not need to apply if the giant planet instability occurred before the terrestrial planets formed (i.e., early and well before the LHB). Note, however, that evolutions violating the constraint D would be difficult to reconcile with the present dynamical structure of the asteroid belt (Walsh & Morbidelli 2011), regardless the timing of the giant planet instability. We therefore do not see a way around criterion D. We report our results in Sect. 5. In doing so, we consider criteria C and D independently, but also discuss the most interesting cases where both C and D were satisfied simultaneously. Note that the success rates for C, D and C&D were computed over the subset of simulations that satisfied A and B (i.e., it would not make sense, for example, to claim success for C if the system ended with incorrect number of planets). Also, B can be satisfied only if A is satisfied. The reported fractions were normalized by the total number of simulations performed in each case (equal to 30 or 100). We ignored other constraints from the populations of small bodies in this work (e.g., cold classical Kuiper belt survival; Batygin et al. 2012). These constraints will be addressed in the upcoming work. 4. Measure of Success and Failure The instability is a chaotic process and as such it is not expected to produce deterministic results. Two planetary systems starting from the same initial conditions (that is, initial after the dispersal of the protoplanetary nebula) are expected to follow different evolutionary paths during the instability, and end up producing different final systems. It would thus be worthless to base our conclusions on one or a small number of simulations that may not be representative of the full range of possible outcomes. Here we performed 30 to 100 simulations for each initial setup. The number of simulations was chosen as a compromise between the need for reasonable statistics and our CPU limitations. A fundamental difficulty with studying the Solar System instability is that the Solar System, as we know it now, is just one realization of al l possible evolution paths. We do not know if this realization is typical of those arising from the initial state, or the nature played a trick on us, and the Solar System has taken an unusual path. The latter possibility would offer a grim perspective on the possibility to determine the Solar System’s state before the – 10 – instability. Here we assume that the Solar System followed a path that was not “exceptional”. Therefore we assume some cutoff probability for success in matching the criteria described in Sect. 3, and give preference to those initial conditions and evolution paths that exceed the cutoff. Now, the assumed cutoff must depend on how selective our criteria are because it is obvious that the measure of the set of initial conditions that lead to the exact architecture of the Solar System is practically zero. We therefore cannot set the criteria that would be too restrictive because we would not be able to obtain satisfactory statistics with our 30-100 simulations per case. The criteria described in Sect. 3 were defined with this in mind. For example, there is nothing wrong with the initial setup that would lead to 3 or 4 giant planets in the end, as long as the two results occur with a roughly comparable likelihood. Statistical fluctuations are also expected for criteria B and C, because some systems may become more excited and/or radially spread (thus failing on B), while others end up dynamically cold (leaving Jupiter with low e55 and failing C). Criterion D can be even more difficult to satisfy because planetary encounters produce a rather large variety of jumping Jupiter histories, in some of which Jupiter does not jump enough (leading to PSat/PJup < 2.3) or jumps too much (leading to PSat/PJup ≫ 2.5). Both these cases would fail D. If we would optimistically assume that the success rate of 50% for each of our four criteria would be satisfactory, the combined probability to simultaneously satisfy all of them would be 1/16 ≃ 6%. Based on this we set the target cutoff probability to be of the order of a few percent. The number of simulations used in this work (see Sect. 2) was chosen so that we are able to measure such probabilities. Note that in many cases studied here the differences between the model results and present Solar System are systematic. For example, in the runs starting with four planets, e55 systematically ends up too low, because the evolution needs to be relatively tranquil to satisfy A and B, and fails on C because e55 is not excited enough (Sect. 5.1.1). It is clear that the probability to match our criteria is ≪ 1% in these cases (although we do not have enough statistics to say how low the success rate actually is). 5. Results 5.1. Case with 4 Planets We start by discussing the results of simulations with four initial giant planets. All four planets have to survive in this case. We performed 2670 integrations for many different resonant chains, including cases of the 2:1, 3:2, 4:3 and 5:4 resonances between pairs of – 11 – planets. Table 5 summarizes the results of selected simulations. These cases were the most interesting ones, i.e. those where the criteria defined in Sect. 3 were satisfied with the highest percentages (Tables 6 and 7 show the selected simulations for the 5- and 6-planet cases). The second and higher order resonances were not considered. The results can be divided into two broad categories that share common traits: (1) Jupiter and Saturn starting in 3:2, and the inner ice giant in the 3:2 or higher degree resonance with Saturn (e.g., 4:3); and (2) Jupiter and Saturn in 3:2, and the inner ice giant in the 2:1 resonance with Saturn (or any other comparably distant orbit), or Jupiter and Saturn in the 2:1 resonance. We discuss the two categories in sections 5.1.1 and 5.1.2, respectively. 5.1.1. 3:2 and Tight In category (1), the system can easily be destabilized, and when it becomes unstable, the instability tends to be violent: one ice giant gets ejected from the Solar System or both ice giants survive, but their orbits end up being very excited (e > 0.1) and/or scattered to large heliocentric distances (a > 50 AU). This happens particularly for disk masses Mdisk . 35 MEarth . Thus the simulations with low disk masses are clearly unsuccessful in matching the present Solar System. For example, in the case with (3:2,3:2,4:3) very closely matching one of the systems discussed in Morbidelli et al. (2007), 87% of simulations show ejection of one or more ice giants (for Mdisk = 35 MEarth and B(1)). None of the remaining 13% satisfied our criterion B, typically because the final orbits were overly excited (e.g., Uranus ended up with i ∼ 10◦). The simulations with Mdisk < 35 MEarth show even larger percentage of ejected planets and more excitation than the ones with Mdisk = 35 MEarth . Thus, disks with Mdisk . 35 MEarth do not apply. This result is robust in that it does not depend on the initial resonant chain (within the definition of category (1)). A single exception from the above rule was one simulation started with (3:2,4:3,3:2), which was, perhaps incidentally, a resonant chain favored by Batygin & Brown (2010). In this particular run, Neptune was scattered to a ≃ 26 AU, where it was quickly circularized by dynamical friction from the planetesimal disk (Fig. 1). The subsequent migration of Neptune and Uranus was minimal and the two orbits ended too close to the Sun (a = 16 and 26 AU, respectively). On the positive side, Jupiter’s orbit was excited by the encounters with Uranus, such that ¯eJup = 0.042 and ¯iJup = 0.37◦ in the end (bar indicates mean values; initial values were 0.002 and 0.013◦ , respectively), both in a good agreement with the present values – 12 – (¯eJup = 0.046 and ¯iJup = 0.37◦ ; Table 1). Even more encouragingly, final e55 = 0.036, also in a close agreement with present e55 = 0.044 (Table 3). The main problem with this run, however, was that the jump in Jupiter and Saturn orbits produced by scattering encounters with Uranus and Neptune was minimal (PSat/PJup moved to ≃ 1.8 at t < 105 yr) so that the final PSat/PJup ratio was too low (≃2.1). It is difficult to evaluate whether this is a fundamental problem, or whether we would find better cases if we increased the number of simulations. We increased the number of simulations to 100 but did not obtain a case that would be even remotely as good as the one shown in Fig. 1. We were therefore probably just lucky when finding this case with fewer runs. We conclude that it might be possible to match the A, B and C constraints with Mdisk ≈ 35 MEarth , but the likelihood of this occurring is .1%. If, in addition, criterion D is considered, the probability drops to ≪1%. A better success in matching criteria A and B was obtained with more massive disks (Mdisk & 50 MEarth ). The success for A and B increased with Mdisk , because the more massive disks were capable of stabilizing the system of four planets more efficiently. The best success rate for A and B was obtained with Mdisk = 75 MEarth and rout = 26 AU, where 67% of jobs matched A and 40% satisfied B. We set the outer disk edge at 26 AU, because in our initial runs with rout = 30 AU Neptune migrated to ∼35 AU. A massive planetesimal disk is apparently capable of migrating Neptune beyond the initial outer disk’s edge. The overall distribution of final orbits was reasonable in this case (Fig. 2). Uranus ended up slightly inside its present orbit (aUra = 17.5 AU compared to real aUra = 19.2 AU), which may or may not be a problem. On one hand, Uranus ended up too close to the Sun in most of the simulations that we performed with 4 planets and the 3:2 resonance between Jupiter and Saturn. So, this is a systematic effect. On the other hand, one might be able to resolve this problem by modifying the initial setup (e.g., starting Uranus on a larger orbit; but see Sect. 5.1.2). Alternatively, the present orbit of Uranus is an outlier in the semima jor axis distribution about 2σ from the mean. A more fundamental (and also systematic) problem of these simulations was that both Jupiter and Saturn ended up on orbits that were too circular. The mean eccentricities that we obtained in the simulations with Mdisk = 75 MEarth were ¯eJup = 0.01 and ¯eSat = 0.02, while real ¯eJup = 0.046 and ¯eSat = 0.054. This was a consequence of the damping effect of massive planetesimal disk on the gas giant orbits (see Fig. 3 for an illustration). This problem was general for all simulations that we performed with massive disks. To understand this effect better, we analyzed the simulations with massive planetesimal disks in more detail. We found that it is generally not a problem to excite eJup and eSat to – 13 – ∼0.05 by scattering encounters with one of the ice giants. What is problematic, however, is to maintain the excited state when the massive planetesimal disk is present. In fact, in all our simulations with massive disks, eJup and eSat were quickly damped after the excitation event. We believe that the damping effect arises from secular friction (Levison et al. 2011) and occurs when e and/or the longitude of perihelion of a planet suddenly changes (e.g., by being scattered off of another planet or due to orbital resonance crossing). This changes the phase and amplitude of the secular forcing between the planet and the planetesimal disk. When the system relaxes the mean eccentricity of the planetesimals increases, while that of the planet decreases. The secular friction damps planet’s inclination as well.5 The excessive damping of eJup and eSat could potentially be avoided if the planetesimal disk were dynamically excited or depleted prior to the event that excites e55 . This might happen, for example, if Neptune were scattered deep into the planetesimal disk (to 28-30 AU) and disrupted it before the e55 excitation event. Unfortunately, we did not see this happening in our simulations, except for a few cases that were flawed for other reasons (e.g., Fig. 1). Another possibility, discussed in more detail in Sect. 5.2.1, is that the planetesimal disk was stirred by large ob jects that formed in the disk and/or by resonances with ice giants. Whatever the stirring mechanism was, however, we found that a disk that did not exert a strong damping effect on Jupiter and Saturn was not capable of preventing ejection of Uranus or Neptune. These simulations thus failed on A or C. There were two additional problems with the simulations such as the one illustrated in Fig. 3. First, the ones that matched criteria A and B were typically those that avoided any strong interactions between planets. This resulted in a nearly smooth migration histories of planets that violated D. In the specific (and representative) case showed in Fig. 3, the system spent more than 3 Myr with 2.1 < PSat/PJup < 2.3, which is expected to be incompatible with the survival of the terrestrial planets (as discussed in Sect. 3). Second, the smooth crossing of the 2:1 resonance between Jupiter and Saturn excited e56 but not e55 . The simulations therefore failed on C. The excitation by resonant crossing mainly occurs as the eccentricity of Saturn evolves over the resonance with Jupiter (Jupiter’s eccentricity changes less because Jupiter has a larger mass). Thus, the resonant crossing di- rectly affects the amplitudes related to the g6 mode, and not those of the g5 mode (Morbidelli et al. 2009b). The needed excitation of e55 could more easily be achieved by scattering en- counters of an ice giant with Jupiter, but such encounters generally lead to violent orbital 5Secular friction is different from an effect known as the dynamical friction (Binney & Tremaine 1987), which arises from encounters between bodies. – 14 – histories that fail A or B if only four giant planets are considered. Finally, we discuss the most promising four-planet case obtained with (3:2,3:2,4:3) and Mdisk = 50 MEarth . In this case, 27% of runs satisfied A, 11% of runs satisfied B, and we also found one case (out of a hundred) that satisfied criterion C (Fig. 4). This specific simulation ended with g5 = 5.98 arcsec yr−1 and g6 = 32.4 arcsec yr−1 , reflecting the fact that Jupiter and Saturn finished a bit closer to each other than in reality, and e55 = 0.026, e56 = 0.010, e66 = 0.031, and e65 = 0.022. This is a pretty good match to the real values (Table 3), although e55 is only a bit larger than a half of the present value. Unfortunately, although this simulations also showed additional encouraging signs (e.g., a ∼0.5 jump of PSat/PJup ), it violated D, because the jump of PSat/PJup was not large enough. In summary, we were not able to find any initial condition for the four-planet case, where the likelihood of simultaneously matching all our four criteria would exceed 1%. 5.1.2. 3:2 and Loose, or 2:1 Here we first consider cases where Jupiter and Saturn are in the 3:2 resonance, the inner ice giant in the 2:1 resonance with Saturn, and various resonances between the first and second ice giants. These cases showed a different behavior than the ones discussed in the previous section. They were more stable in that breaking the resonant locks of the inner ice giant did not always generate an instability. If it did, the instability was usually mild such that close encounters between planets did not occur, and planets ended up migrating smoothly. This had two conflicting consequences. The smooth migration gave a large success rate for A and B, because the system did not suffer destabilizing perturbations and evolved quietly. In most other aspects, however, these simulations produced incorrect results. First of all, Jupiter and Saturn did not reach the current PSat/PJup = 2.5 unless the disk was heavy. This was because, in absence of encounters with ice giants, the gas giants did not jump, and all the migration work had to be accomplished by planetesimals. With heavy disks, which gave a more reasonable final PSat/PJup ratio, PSat/PJup slowly migrated over 2.1-2.3 and violated our constraint D. The eccentricities of Jupiter and Saturn were excited when these planets crossed their mutual 2:1 resonance. The resonant crossing led to eJup ≃ 0.04 and eSat ≃ 0.08, which was lovely, except that this eccentricity excitation was contained in the secular modes related to g6 . The simulations therefore violated C as e55 was never large enough. We have not found a single case where the constraint C or D were satisfied. We conclude that the resonant chains with the inner ice giant in the 2:1 resonance with Saturn can be ruled out. – 15 – The case with the 2:1 resonance between Jupiter and Saturn showed behavior that was reminiscent to that of the ONM (Tsiganis et al. 2005). The success rate for A and B was large, but Jupiter and Saturn migrated smoothly and criterion D was not satisfied. Also, because of the lack of an adequate excitation agent, Jupiter ended up with orbital eccentricity that was too low (Fig. 5), and C was not satisfied as well. In addition, Uranus typically reached aUra < 18 AU, well inside its present orbit. In the ONM (Tsiganis et al. 2005), the inner ice giant was started far beyond the 2:1 resonance with Saturn, which helped Uranus to reach 19.2 AU. 5.2. Case with 5 Planets Following Nesvorn´y (2011) and Batygin et al. (2012) we considered the case with five giant planets as an interesting solution to some of the problems discussed in the previous section. The main advantage of the five planet case is that the system can afford to lose a planet. This resolves, to a degree, the conflicting requirements for a significant excita- tion of e55 , which requires strong scattering encounters between Jupiter and ice giants, and constraints A and B. We performed 4050 integrations with five planets for 19 different resonant chains. We tested different masses of the fifth planet and different initial orbits. We found that the best results were obtained when the fifth planet had mass comparable to that of Uranus/Neptune, and was placed between the orbits of Saturn and Uranus. We discuss on this case below. The case in which the fifth planet was given a large mass typically resulted in a violent instability and ejection of Uranus and/or Neptune from the system. If the fifth planet was given a lower mass and orbit beyond that of Neptune, the inner ice giant got ejected during the instability thus leaving a final system with incorrect masses. This later case bears similarities to the four-planet case discussed in Sect. 5.1. 5.2.1. (3:2,3:2,4:3,5:4) and Alike Morbidelli et al. (2007) found one system with five planets that remained dynamically stable over 1 Gyr (in absence of perturbations from the planetesimal disk).6 The five planets 6 In the published article, Morbidelli et al. (2007) claimed that the five-planet configuration was unstable. Unfortunately, to test the stability, Morbidelli et al. (2007) used the original version of SyMBA (Duncan et al. 1998), which had a problem with integrating very compact planetary configurations for long periods of time. After fixing this problem, Levison et al. (2011) found that the five-planet configuration identified by – 16 – were in the (3:2,3:2,4:3,5:4) resonant chain. The extra ice giant had mass similar to that of Uranus/Neptune. The orbital eccentricities were 0.004, 0.02, 0.08, 0.035, and 0.01 (from the inner to outer planet). Levison et al. (2011) showed that in such a case, where the inner ice giant had the largest eccentricity, it can be expected that the (late) instability was triggered by the inner ice giant (i.e., B(1) in notation introduced in Sect. 2). We discuss this case first and get to different trigger mechanisms and resonant chains later. We found that the low-mass disks with Mdisk . 20 MEarth did not work, because the system frequently lost two or more planets, ending with three or less. Specifically, only ∼10% of the simulations with Mdisk = 20 MEarth ended up with four planets, but the orbits were all wrong. This is similar to what happened in the four-planet case described in Sect. 5.1.1 for Mdisk . 35 MEarth . The instability is simply too violent in this case to be contained by a low-mass planetesimal disk. Very massive planetesimal disks with Mdisk > 50 MEarth did not work as well mainly because too much mass was processed through the Jupiter/Saturn region, leading to excessive migration of Jupiter and Saturn, and incorrect final PSat/PJup > 3. Intermediate disk masses, Mdisk = 35 MEarth and 50 MEarth , were more promising. With ∆ = 1 AU (recall that ∆ denotes the initial radial separation between the outer ice giant and inner edge of the planetesimal disk; see Sect. 2), these cases showed the 13% and 37% success rates for A, and 3% and 23% for B, respectively. These fractions are pretty much independent of the instability trigger. The larger success rate for Mdisk = 50 MEarth than for Mdisk = 35 MEarth reflects the stabilizing effect of the planetesimal disk, which increases with its mass. The success rate tends to drop with increasing ∆, but this trend is not always clear. For example, the success rate for A drops from 37% to 23% when ∆ is increased to 3.5 AU for Mdisk = 50 MEarth (similarly, B drops from 23% to 10%), while the percentages for ∆ = 1 AU and 3.5 AU are similar for Mdisk = 35 MEarth . Figure 6 shows the final systems for Mdisk = 50 MEarth . This result is encouraging. The distributions show a larger range of values than what we got in the four-planet case, because the interaction of five planets is more complex and leads to a larger variety of results. Neptune’s model orbit tends to be slightly more excited than the real one, but this difference is small and probably not fundamental in that it could be resolved by tuning the parameters. On the positive side, some of the simulations that satisfied A and B are now also showing signs of success for the criterion D (Fig. 7). This has not happened in the four-planet case at all. Still, the success rate for the criterion D is only marginal, mainly because it is intrin- sically difficult to jump from PSat/PJup = 1.5 to >2.3, and end up below 2.5. First, as Morbidelli et al. (2007) was stable. – 17 – the jump must be rather large, it is hard find cases where PSat/PJup is larger than 2.3 and smaller than 2.5 after the jump. This is statistically unlikely (3.3% averaging over the first seven lines of Table 6) given the spread of possible jump amplitudes for different encounter geometries. Second, there is a problem with the residual migration if the planetesimal disk is initially massive. In such cases, even if 2.3 < PSat/PJup < 2.5 after the jump, there is still a significant mass in the planetesimal disk that needs to be processed through the Jupiter/Saturn zone. Consequently, PSat/PJup keeps increasing after the jump and reaches > 2.5 (for Mdisk & 50 MEarth ; e.g., Fig. 7c). The residual migration is difficult to avoid unless the planetesimal disk had a relatively low mass to start with. An additional problem with the case discussed here is that e55 that we obtained in the simulations was nearly always too small. This happened because even if eJup was kicked by the ejected planet, it was quickly damped at later times by secular friction from the planetesimal disk (see, e.g., Fig. 7b). We tested several possible solutions to this problem. Disks with Mdisk . 35 MEarth did not work for the (3:2,3:2,4:3,5:4) resonant chain because two or more planets were lost too often, end even if four planets survived in the end their orbits were too wild (B satisfied in 5-10% of cases only). We tested dynamically excited planetesimal disks. The disk excitation reduces the disk’s ability to damp eJup and may thus lead to better results. The disk could have been excited by large bodies that formed in it, as evidenced by Pluto-sized bodies in the present Kuiper belt. In addition, the disk could have been excited near its inner edge by orbital resonances with the outer ice giants. In both cases, the disk planetesimals were distributed according to the Rayleigh distribution. We used hei = 0.15 and hii = 0.075 (case E1), and hei = 0.1 and hii = 0 (case E2), where brackets denote the mean values. Case E1 mimicked the excitation expected from the large bodies in the disk. Case E2 would be more appropriate for the orbital resonances that do not generally excite inclinations. Recall that for the cold planetesimal disk we used a minimal separation of ∆ ≥ 0.5 AU between the outer planet’s orbit and semima jor axis of disk particles (see discussion in Sect. 2). Now that we deal with excited disks with e up to 0.15, we used ∆ ≥ 3.5 AU for consistency, so that the minimal physical distance between planetesimals and planets was larger than 0.5 AU. With Mdisk = 50 MEarth , case E1 gave A = 30% and B = 10%, which was comparable to what we obtained for this setup with a dynamically cold disk. Interestingly, there was one simulation (out of 30) that matched C as well (e55 = 0.032). Case E2 gave A = 26% and B = 7%, and also one simulation where the criterion C was satisfied (e55 = 0.026). Conversely, the excited disks with Mdisk = 35 MEarth did not show any case where C would be satisfied. We conclude that the disk excitation may help, but the simulations matching – 18 – C are still disturbingly rare. We studied the dependence of the results on the mass of the fifth planet. We found that decreasing the fifth planet’s mass helps to boost the success rate for A and B. For example, with (3:2,3:2,4:3,5:4), Mdisk = 50 MEarth , B(1), and the mass of the inner ice giant MIce = 0.5 MNep , the success rate in matching A and B was 63% and 30%, while it was 37% and 23% with MIce = MNep . In all studied cases with MIce = 0.5 MNep , the small inner planet was removed (83% ejected and 17% collided with Jupiter). On the downside, we found that the jump of PSat/PJup and excitation of eJup with MIce < 1 MNep were smaller than with MIce ≃ MNep , which compromised the success rate for C and D. In fact, eJup was excited in most of the runs with MIce = 0.5 MNep not by the scattering encounter itself, but rather when Jupiter and Saturn crossed the 2:1 resonance. As we discussed Sect. 5.1, the resonant crossing has only a minor effect on e55 . Increasing the mass of the inner ice giant did not help either, because the system became violently chaotic during the instability and typically lost two or more planets. In addition, in many cases where only one planet was ejected, the lost planet was one of the outer ice giants. For example, with Mdisk = 50 MEarth , B(1) and MIce = 2MNep we found that A = 15% and B = 3% (when only the removal of the massive inner planet was considered). We conclude that the problems with the low success rate of C and D cannot be resolved by changing the mass of the fifth planet. The problems discussed above may be related to the fact that the ice giants’ orbits were closely packed together for (3:2,3:2,4:3,5:4), perhaps too closely. We found that the other compact resonant chains investigated here, such as (3:2,3:2,4:3,4:3) or even (3:2,3:2,3:2,4:3), behave in very much the same way as (3:2,3:2,4:3,5:4), giving way to violent instabilities that fail to satisfy the constraints. We therefore decided to focus on more relaxed resonant chains. We discuss the results obtained with (3:2,3:2,3:2,3:2) first. These results were more encouraging. 5.2.2. Relaxed Chains with the 3:2 Jupiter-Saturn Resonance The (3:2,3:2,3:2,3:2) resonant chain leads to a different mode of instability, if ∆ < 2 AU. Figure 8 illustrates a case with Mdisk = 35 MEarth , B(1) and ∆ = 1.5 AU. Unlike in all cases discussed so far, breaking the resonant lock of the inner ice giant did not result in an im- mediate instability. Instead, Neptune migrated deep into the planetesimal disk before any scattering encounters between planets occurred. As Neptune and Uranus scattered planetes- imals into the Jupiter/Saturn zone, Jupiter and Saturn underwent divergent migration and – 19 – left the 3:2 resonance. Their eccentricities were then excited by the the 5:3 resonance cross- ing, and the instability happened shortly after (16.8 Myr after the start of the simulation). The epoch of planetary encounters was brief, lasted from 16.816 Myr to 16.9 Myr at which point the fifth planet was ejected from the Solar System (escape speed V∞ = 0.74 km s−1 ).7 All planets participated in the encounters, as required for capture of the irregular satellites,8 but all these encounters involved the fifth planet only. None of the sur- viving giant planets had an encounter with another surviving giant. In total, 282 encounters happened in this case, where an encounter is defined here by the condition that the Hill spheres of the two planets at least partially overlap. Now, the case illustrated in Fig. 8 simultaneously matched al l our criteria. Amplitude e55 was excited by the ejection of the fifth planet by Jupiter and was not damped too much during the following evolution, because the planetesimal disk had a relatively low initial mass, and was partially disrupted by Neptune prior to the excitation event. The final amplitudes were e55 = 0.024 and e56 = 0.012. Here, the amplitude e55 was lower than real one, but in another simulation (with the same setup) that matched all our criteria as well, we obtained e55 = 0.041 and e56 = 0.013. The scattering encounters with the fifth planet produced a jump of PSat/PJup from 1.7 to 2.5, just as required to avoid the secular resonances with the terrestrial planets (Fig. 8c). There was a small problem with the residual migration in this run, because PSat/PJup continued to evolve past 2.5, while PSat/PJup = 2.49 in the present Solar System, but we obtained better results in another simulation with the same setup (Figure 9; although in this case e55 was damped to 0.013). The overall distribution of planetary orbits obtained in simulations with the (3:2,3:2,3:2,3:2) resonant chain, Mdisk = 35 MEarth , B(1) and ∆ = 1.5 AU is shown in Fig. 10. The overall success rate in matching the criteria A, B, C and D was 33%, 16%, 4% and 8%, respectively (note that matching C or D required that B was matched). The overall success rate for simultaneous matching of C and D was 4%. This is the best result discussed so far. For a comparison, a similar setup with a more massive disk, Mdisk = 50 MEarth , gave to A = 30%, B = 17%, C = 0% and D = 3%. The success for C dropped because e55 was damped more efficiently with the massive planetesimal disk. A light disk with Mdisk = 20 MEarth gave A = 20%, B = 7%, C = 3% and D = 0%. 7Typically, V∞ = 0.5-2 km s−1 . 8We do not define planetary encounters as a separate criterion, because constraints C and D require planetary encounters and are generally much more restrictive. – 20 – The mode of instability discussed for (3:2,3:2,3:2,3:2) above may be problematic because the migration of Uranus and Neptune prior to the instability may require that the inner edge of the planetesimal disk was initially close to the outer ice giant. The edge may then need to be fine-tuned to generate the delay between the formation of the Solar System and the LHB. This is a problem similar to that of the original Nice model, which prompted Levison et al. (2011) to consider distant disks and a different trigger mechanism. Here we do not study the problem of delay in detail. Instead, given that the instability mode with prior migration of Neptune works much better than any other case, we pursued investigations into this instability mode further. We first tested several cases where the inner ice giant was placed the 2:1 resonance with Saturn ((3:2,2:1,2:1,3:2), (3:2,2:1,3:2,2:1), etc.). We found that these resonant chains did not work, because the orbits spread without suffering any ma jor instability, which left five planets behind. Unlocking the resonance of one of the ice giants did not destabilize the system as well, mainly because the inner ice giant was initially far from Saturn, and safe, even if its orbit was not locked deep in the 2:1 resonance. Motivated by these results, we considered several resonant chains with the inner ice giant in the exterior 3:2 (or 4:3) resonance with Saturn, and in the 2:1 resonance with the inner surviving ice giant (e.g., (3:2,3:2,2:1,2:1), (3:2,4:3,2:1,2:1) and (3:2,3:2,2:1,3:2). This initial setup increases the likelihood that the inner ice giant gets ejected, because it starts relatively close to Saturn, and that the outer two ice giants survive, because there is initially a large radial gap between them and the inner ice giant. We found that the extended resonant chains such as (3:2,3:2,2:1,2:1) and (3:2,4:3,2:1,2:1) did not work because Neptune migrated beyond 30 AU, independently of the assumed mass and extension of the planetesimal disk. The explanation for this is clear. As in all simu- lations, Neptune scattered planetesimals outward and inward relative to its orbit. Many of those that were scattered outward were scattered to a > 30 AU, because initial aNep ≃ 25 AU for the extended resonant chains. The ones that were scattered inward must have been scattered relatively strongly to be handed to Uranus, because the orbits of Uranus and Nep- tune were radially spaced (due to the initial 2:1 resonance). As a result, Neptune migrated outward, being propelled by the strong scattering encounters, all the way into the cloud of planetesimals that it scattered beyond 30 AU. This occurred even if the planetesimal disk had low mass (Mdisk = 20 MEarth ) and/or was initially narrow (spanning, e.g., 26-27 AU). Also, even lighter disks with Mdisk . 15 MEarth did not show much success for A. Figure 11 shows the distribution of final orbits for the (3:2,3:2,2:1,3:2) resonant chain and Mdisk = 20 MEarth . These results are more promising. As the distributions obtained with B(0) and B(1) did not differ (assuming ∆ ≃ 1 AU), we combined them together in Fig. 11. – 21 – The combined success rate was A = 32%, B = 12%, C = 7%, D = 5% and C&D = 5%. This was similar to the statistics obtained for (3:2,3:2,3:2,3:2) and Mdisk = 35 MEarth (Fig. 10), but here things were slightly better in detail. For example, the distribution of model aUra nicely matched the present semima jor axis of Uranus, while the model aUra was slightly lower than required in Fig. 10. The model values of eJup obtained with (3:2,3:2,2:1,3:2) were almost perfect in that they closely surrounded the target ¯eJup = 0.046. Figures 12, 13 and 14 show examples of the successful simulations. We illustrate these cases in detail because they were some of the best results obtained in this work. The simulation in Fig. 12 ((3:2,3:2,2:1,3:2) and Mdisk = 20 MEarth ) ended with g5 = 4.95 arcsec yr−1 , g6 = 29.7 arcsec yr−1 , g7 = 4.53 arcsec yr−1 , and g8 = 0.52 arcsec yr−1 . These values were a close match to those listed for the real planets in Table 2. The amplitudes of eccentricity modes were e55 = 0.049, e56 = 0.033, e57 = 0.024, e65 = 0.041, e66 = 0.042, e67 = 0.003, and e77 = 0.039. The frequencies and amplitudes of inclination modes were equally good. Neptune ended with an eccentricity that was about twice as large as its present value, but this was at least partly related to a minor mismatch in the simulation, because Uranus and Neptune crossed the 2:1 resonance, which did not happened in reality. Figure 13 shows another case obtained with the same setup. In this case, g5 = 4.83 arcsec yr−1 , g6 = 28.9 arcsec yr−1 , g7 = 3.55 arcsec yr−1 , and g8 = 0.64 arcsec yr−1 . The amplitudes were e55 = 0.027, e56 = 0.014, e57 = 0.002, e65 = 0.041, e66 = 0.024, e67 = 0.002, e77 = 0.033. Finally, the case illustrated in Fig. 14 had frequencies g5 = 4.66 arcsec yr−1 , g6 = 28.1 arcsec yr−1 , g7 = 3.45 arcsec yr−1 , g8 = 0.61 arcsec yr−1 , and amplitudes e55 = 0.024, e56 = 0.019, e57 = 0.004, e65 = 0.019, e66 = 0.057, e67 = 0.004, e77 = 0.040, e88 = 0.007. This case differs with respect to the previous two in that the innermost ice giant survived and evolved onto Uranus-like orbit, while the middle ice giant was ejected. The Neptune’s model eccentricity was very good in this simulation. Using (3:2,3:2,2:1,3:2) and Mdisk ≃ 20 MEarth we tested how the results were affected by small changes (up to 50%) in the mass of the fifth planet. We found that it was probably not lower than 0.7 MUra because in the simulations with these low masses Jupiter (and Saturn) were not kicked enough by the planet’s ejection. Masses slightly larger than MNep showed lower success rates for A and B, but still worked relatively well for C and D. To summarize, we find that the fifth planet probably had mass in the Uranus/Neptune range, which is encouraging in that we do not need to invoke any special mass regime. – 22 – 5.2.3. 2:1 Jupiter-Saturn Resonance The initial conditions discussed in this section are probably academic, because the 2:1 resonance between Jupiter and Saturn is not favored by the standard model of the migration of planets in the protoplanetary gas disks, and their capture in resonances (see Sect. 3). We therefore discuss this case briefly, concentrating on the main differences with respect to the results obtained with more conservative assumptions. The five-planet case with the 2:1 resonance would require that the (ejected) inner ice giant had lower mass, because when Jupiter and Saturn start in the 2:1 resonance, their period ratio needs to change by ∼0.5 only (from 2 to 2.49), which requires a smaller pertur- bation. The best results were obtained with MIce = 0.5 MNep and 15 . Mdisk . 35 MEarth . For example, with Mdisk = 20 MEarth , we obtained a ∼50-70% success for A and 20-40% success for B. In addition, because the required jump of PSat/PJup is smaller and easier to achieve, the simulations also show a large success for D (reaching 20% in the most favorable cases). Figure 15 illustrates this case. The model semima jor axis of Uranus and Neptune were a bit smaller than what would be ideal, but this should easily be corrected by extending the disk slightly beyond 30 AU (we used rout = 30 AU). All else looked great, except that the Jupiter’s eccentricity was generally too small and violated constraint C. We found only two simulations out of more than 1,000, where the criterion C was satisfied. This problem arises because while a relatively low-mass ice giant is required to produce the needed small jump of PSat/PJup , the same low-mass planet is generally incapable of ex- citing e55 enough during encounters. In addition, the problem with secular friction discussed in the previous sections is also apparent here. To show this clearly, we adjusted the initial system so that Jupiter had a substantial eccentricity initially (Fig. 16). This did not helped at all because the initial eccentricity was quickly damped. The problem with matching C in the case with five planets and 2:1 resonance between Jupiter and Saturn is difficult to avoid (at least we were not able to resolve this problem with 1,000+ simulations). Thus, even if the success rate for A, B and D was promisingly large, we are not overly optimistic about this case. A more thorough testing will be required to reach a more definitive conclusion.9 9The discussion presented here was based on the overall synthesis of our simulation results. Note that it can be misleading to compare two specific simulation sets in Table 6. For example, we have just been lucky to find one case (out of 30) for (2:1,3:2,3:2,3:2), Mdisk = 20 MEarth and B (1) for which the criterion C was satisfied. In contrast, the results for the 3:2 Jupiter-Saturn resonance were consistently good, showing >1% – 23 – 5.3. Case with 6 Planets We performed 1290 simulations in total for six initial planets and nine different resonant chains. Given that the number of parameters in the six-planet case is much larger than the one with four or five planets, we are not confident that we sampled parameter space exhaustively enough to make detailed conclusions. Still, the six-planet case is very interesting because we were able to obtain good results even after a relatively small number of trials. Figure 17 illustrates one of the most promising six-planet results. This result was obtained for the (3:2,4:3,3:2,3:2,3:2) resonant chain and Mdisk = 20 MEarth . The two extra ice giants were placed between Saturn and the inner surviving ice giant, and were given masses MIce1 = MIce2 = 0.5 MNep (the six-planet case with MIce1 = MIce2 = 1 MNep does not work, because the instability is too violent). The results show a large spread, because the six planets have complex interactions during the instability. The distributions of model orbits nicely overlap with the real orbits. Even the details match. For example, most simulations ended with eN ≃ 0.01, just as needed. Given our previous experience with the simulations of instability in the four- and five-planet cases, the agreement in Fig. 17 is remarkable. Because of the larger spread, however, the success rate was lower than that for our best five-planet simulations. In the specific case discussed above, we obtained A = 30%, B = 10%, C = 3%, D = 3% and C&D = 2%. The instability typically occurred in two steps, corresponding to the ejection of the two planets. Sometimes, as in Fig. 18, the ejection of the two planets was nearly simultaneous, but most of the times there was a significant delay between ejections. This was useful because the first planet’s ejection partially disrupted the planetesimal disk and reduced its capability to damp e55 , which was then excited by the second planet’s ejection. While this mode of instability can be important, we would need to increase the statistics (>100 simulations for each initial condition) to be able to properly resolve the small success fractions in the six-planet case. 6. Discussion One of the main results discussed in this paper is that the Solar System instability with five initial planets tends to give better results then the instability with four initial planets. Batygin et al. (2012; hereafter B12) performed a similar analysis and found instead that the success rate for matching all criteria simultaneously. – 24 – four- and five-planet cases show about the same success rate in matching constraints. Here we discuss the possible causes of this disagreement. B12 used an N -body integrator with forces that mimic the effects of gas to generate the initial resonant chains of planets. This method is similar to that described for our Phase-1 integrations in Sect. 2, and should result in similar initial orbits for Phase 2. The resonant chains explicitly discussed in B12 included (3:2,2:1,5:4,5:4) and (3:2,3:2,2:1,4:3). We discussed this type of relaxed resonant chains in Sect. 5.2.2. The initial properties of the planetesimal disk in B12 were also similar to the ones used here. We therefore believe that the initial conditions are not the main cause of the disagreement. Instead, we explain below that different conclusions were reached in B12 probably because B12 gave a different emphasis to different constraints (see Sect. 3 for our constraints). On one hand, B12 found, correcting the results of Batygin & Brown (2010), that ∼ 10% of simulations with four initial planets matched constraints. Their constraints, however, were somewhat different and less restrictive than the ones we used here. First of all, B12 did not apply any upper limit on the excitation of planetary orbits, while we required in the criterion B that e < 0.11 and i < 2◦ (relative to the invariant plane). It is therefore possible that in at least some of the simulations of B12, which gave reasonable e55 , some planetary orbits had e > 0.11 and/or i > 2◦ . B12 did not consider the constraint from the terrestrial planets (our criterion D). It is therefore possible that in some of the simulations that were found to be successful in B12, the evolution of the g5 mode was smooth, which would lead to the secular resonances with the terrestrial planets. Note that some of these models could still be valid, but the fraction should be small, probably <10% (Brasser et al. 2009). While the two issues pointed out above can resolve some of the main differences between B12 and our results, we are still puzzled, given our clearly negative results for the four-planet case, that B12 were able to find positive results for the four-planet case even with very massive disks (Mdisk > 50 MEarth ). We tried to repeat the simulations reported in B12 and found that massive disks efficiently damp e55 and lead to problems with residual migration of Jupiter and Saturn. On the other hand, B12 considered constraints from the Kuiper belt that we ignored here. As B12 pointed out, only about half of their five-planet simulations were compatible with the low eccentricities and low inclinations in the cold classical Kuiper belt,10 while 10The cold classical Kuiper Belt is a population of trans-Neptunian bodies dynamically defined as having orbits with semima jor axis a = 42-48 AU, perihelion distances that are large enough to avoid close encounters – 25 – most of their four-planet cases were fine. We may have been therefore overly optimistic, by a factor of ∼2, in favoring the five-planet case over the four-planet case. The constraints on the planetary instability from the small body reservoirs in the Solar system (Kuiper belt ob jects, asteroids, Tro jans and satellites) are clearly very important. We will consider these constraints in the follow-up work. B12 and this work agree on one central issue. As B12 was not concerned with the delay between the protoplanetary nebula dispersal and LHB, they placed the planetesimal disk’s inner edge just beyond the outer ice giant’s orbit. This triggered, almost immediately, Neptune’s fast migration through the disk and the instability mode akin to that we discussed for our most successful runs in Sect. 5. B12 therefore incidentally explored the setup that we favored here based on a broader sampling of the initial parameters. It remains to be understood whether this setup can lead to the late instability without overly restrictive assumptions on the structure of the planetesimal disk. It would be useful in this context, for example, if the LHB started ∼ 4.2 Gyr ago as sug- gested by Bottke et al. (2012), because a shorter time delay since the protoplanetary nebula dispersal (≃350 Myr) would help to relax constraints on ∆. Alternatively, the planetesimal disk could have been stirred by large bodies that formed in it, and spread over time, so that the effective ∆ decreased. This could lead to the late instability even if the initial ∆ was large. Investigations of these issues are left for future work. 7. Conclusions Recent studies suggest that Jupiter and Saturn formed and migrated in the protoplane- tary gas disk to reach a mutual resonance, most likely the 3:2 resonance, where the Saturn’s orbital period was 3/2 longer than that of Jupiter. After the gas disk’s dispersal, the orbits of Jupiter and Saturn must have evolved in some way to eventually arrive to the current orbits with the orbital period ratio of 2.49. This can most easily be achieved, considering constraints from the terrestrial planets and e55 , if Jupiter and Saturn scattered off of Uranus or Neptune, or a planet with mass similar to that of Uranus or Neptune. We performed N -body integrations of the scattering phase between the Solar System’s giant planets, including cases where one or two extra ice giants were assumed to have formed in the outer Solar System and ejected into interstellar space during instability. We found that the initially compact resonant configurations and low masses of the planetesimal disk to Neptune, and low inclinations (i . 5◦ ). – 26 – (Mdisk < 50 MEarth ) typically lead to violent instabilities and planet ejection. On the other hand, the initial states with orbits that are more radially spread (e.g., Jupiter and Saturn in the 2:1 resonance) and larger Mdisk result in smooth migration of the planetary orbits that leads to incorrectly low e55 and excitation of the terrestrial planet orbits. Finding the sweet spot between these two extremes is difficult. Some of the statistically best results were obtained when assuming that the Solar System initially had five giant planets and one ice giant, with the mass comparable to that of Uranus and Neptune, was ejected to interstellar space by Jupiter. The best results were obtained when the fifth planet was assumed to have the mass similar to Uranus/Neptune, was placed on an orbit just exterior to Saturn’s (3:2 and 4:3 resonances work best), and the orbits of Uranus and Neptune migrated into the planetesimal disk before the onset of planetary scattering. This mode of instability is favored for several reasons, as described below. As planetesimals are scattered by Uranus and Neptune and evolve into the Jupiter/Saturn region, Jupiter, Saturn and the fifth planet undergo divergent migration. This triggers an instability during which the fifth planet suffers close encounters with all planets and is even- tually ejected from the Solar System by Jupiter. Uranus and Neptune generally survive the scattering phase, because their orbits migrated outward during the previous stage and opened a protective gap between them and the gas giants. This mode of instability produces just the right kind of Jupiter’s semima jor axis evolution -known as jumping Jupiter- that is required from the terrestrial planet constraint. Moreover, e55 , excited by the fifth planet ejection, is not damped to incorrectly low values by secular friction from the planetesimal disk, because the planetesimal disk had been disrupted by Uranus and Neptune before the excitation event. The low mass of the planetesimal disk at the time of planet scattering also leads to only a brief migration phase of Jupiter and Saturn after the scattering phase, and prevents PSat/PJup from evolving beyond its current value. The excessive residual migration of Jupiter and Saturn was a problem in most other cases investigated here. The mode of instability with early migration of Uranus and Neptune is problematic, however, because the inner edge of the planetesimal disk may need to be fine-tuned to generate the delay between the formation of the Solar System and the LHB. In addition, the range of possible outcomes is rather broad, indicating that the present Solar System is neither typical nor an expected result, and occurs, in best cases, with only a ≃5% probability (as defined by simultaneous matching of all four criteria defined in Sect. 3). In ≃95% of cases, the simulations ended up failing at least one of our constraints. This may seem unsatisfactory, but given the issues discussed in Sect. 4, the fact that our criteria are relatively strict, and – 27 – because the instability-free model does not work at all,11 these findings should be seen in a positive light. An important follow-up of this work will be to consider additional constraints from the small body populations (e.g., the dynamical structure of the Kuiper belt), and see whether the successful simulations identified here will also match those constraints. The case with six giant planets is also interesting in that the instability occurs in two steps, corresponding to the ejection of the two planets. The best results were obtained in this case when the two ejected planets were given similar masses (about half the mass of Neptune) and were placed between the orbits of Saturn and the inner surviving ice giant. As expected, the six-planet case leads to a larger variety of results than the five-planet case. The probability of ending the six-planet simulation with the present properties of the solar system is therefore lower than in the five-planet case. Still, our six-planet results are fundamentally better than those obtained in the four-planet case, where the differences were systematic (e.g., e55 never large enough), and the success rate was below the resolution limit of our study. This work was supported by the NASA’s National Lunar Science Institute and Outer Planets Research programs. Alessandro Morbidelli thanks Germany’s Helmholtz Alliance for providing support through its Planetary Evolution and Life program. David Nesvorny thanks the Observatoire de la Cote d’Azur for hospitality during his sabbatical in Nice. We also thank an anonymous reviewer for helpful suggestions. REFERENCES Agnor, C. B., & Lin, D. N. C. 2012, ApJ, 745, 143 Batygin, K., & Brown, M. E. 2010, ApJ, 716, 1323 Batygin, K., Brown, M. E., & Fraser, W. C. 2011, ApJ, 738, 13 Batygin, K., Brown, M. E., & Betts, H. 2012, ApJ, 744, L3 Binney, J., & Tremaine, S. 1987, Princeton, NJ, Princeton University Press. Bottke, W. F., Vokrouhlick´y, D., Minton, D., et al. 2012, Nature, 485, 78 11The instability-free model with four planets can be easily tuned to satisfy B, but not C and D, and constraints from the small body populations (e.g., Walsh & Morbidelli 2011, Dawson & Murray-Clay 2012). – 28 – Bouvier, A., Blichert-Toft, J., Moynier, F., Vervoort, J. D., Albarede, F. 2007. Geochim. Cosmochim. Acta, 71, 1583 Burkhardt, C., Kleine, T., Bourdon, B., Palme, H., Zipfel, J., et al. 2008. Geochim. Cos- mochim. Acta, 72, 6177 Brasser, R., Morbidelli, A., Gomes, R., Tsiganis, K., & Levison, H. F. 2009, A&A, 507, 1053 Chapman, C. R., Cohen, B. A., & Grinspoon, D. H. 2007, Icarus, 189, 233 Cuk, M. 2007, Bulletin of the American Astronomical Society, 38, 537 Dawson, R. I., & Murray-Clay, R. 2012, arXiv:1202.6060 Duncan, M. J., Levison, H. F., & Lee, M. H. 1998, AJ, 116, 2067 Goldreich, P., & Sari, R. 2003, ApJ, 585, 1024 Gomes, R. S., Morbidelli, A., & Levison, H. F. 2004, Icarus, 170, 492 Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005, Nature, 435, 466 Haisch, K. E., Jr., Lada, E. A., Lada, C. J. 2001. ApJ, 553, L153 Hartmann, W. K., Ryder, G., Dones, L., & Grinspoon, D. 2000, Origin of the Earth and Moon, 493 Jewitt, D., & Haghighipour, N. 2007, ARA&A, 45, 261 Kley, W. 2000, MNRAS, 313, L47 Lee, M. H., & Peale, S. J. 2002, ApJ, 567, 596 Levison, H. F., Dones, L., Chapman, C. R., et al. 2001, Icarus, 151, 286 Levison, H. F., Morbidelli, A., Vanlaerhoven, C., Gomes, R., & Tsiganis, K. 2008, Icarus, 196, 258 Levison, H. F., Bottke, W. F., Gounelle, M., et al. 2009, Nature, 460, 364 Levison, H. F., Morbidelli, A., Tsiganis, K., Nesvorn´y, D., & Gomes, R. 2011, AJ, 142, 152 Malhotra, R. 1995, AJ, 110, 420 Marcy, G. W., Butler, R. P., Fischer, D., et al. 2001, ApJ, 556, 296 – 29 – Masset, F. 2000, A&AS, 141, 165 Masset, F., & Snellgrove, M. 2001, MNRAS, 320, L55 Minton, D. A., & Malhotra, R. 2011, ApJ, 732, 53 Morbidelli, A., & Crida, A. 2007, Icarus, 191, 158 Morbidelli, A., Levison, H. F., Tsiganis, K., & Gomes, R. 2005, Nature, 435, 462 Morbidelli, A., Tsiganis, K., Crida, A., Levison, H. F., & Gomes, R. 2007, AJ, 134, 1790 Morbidelli, A., Levison, H. F., Bottke, W. F., Dones, L., & Nesvorn´y, D. 2009a, Icarus, 202, 310 Morbidelli, A., Brasser, R., Tsiganis, K., Gomes, R., & Levison, H. F. 2009b, A&A, 507, 1041 Morbidelli, A., Brasser, R., Gomes, R., Levison, H. F., & Tsiganis, K. 2010, AJ, 140, 1391 Nesvorn´y, D., & Vokrouhlick´y, D. 2009, AJ, 137, 5003 Nesvorn´y, D., Vokrouhlick´y, D., & Morbidelli, A. 2007, AJ, 133, 1962 Nesvorn´y, D. 2011, ApJ, 742, L22 Pierens, A., & Raymond, S. N. 2011, A&A, 533, A131 Pierens, A., & Nelson, R. P. 2008, A&A, 482, 333 Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954 Ryder, G., 2002, Journal Geophysical Research-Planets, 107, 6 Sidlichovsk´y, M., & Nesvorn´y, D. 1996, Celestial Mechanics and Dynamical Astronomy, 65, 137 Sumi, T., et al. 2011, Nature, 473, 349 Thommes, E. W., Duncan, M. J., & Levison, H. F. 1999, Nature, 402, 635 Thommes, E. W., Bryden, G., Wu, Y., & Rasio, F. A. 2008, ApJ, 675, 1538 Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Nature, 435, 459 Veras, D., & Raymond, S. N. 2012, MNRAS, 421, L117 – 30 – Walsh, K. J., & Morbidelli, A. 2011, A&A, 526, A126 Walsh, K. J., Morbidelli, A., Raymond, S. N., O’Brien, D. P., & Mandell, A. M. 2011, Nature, 475, 206 Weidenschilling, S. J., & Marzari, F. 1996, Nature, 384, 619 This preprint was prepared with the AAS LATEX macros v5.2. – 31 – ¯a (AU) 5.203 9.555 19.22 30.11 ¯e ¯i (◦) 0.37 0.046 0.90 0.054 0.044 1.02 0.67 0.010 Jupiter Saturn Uranus Neptune Table 1: Mean orbital elements of planets. The mean values reported here were obtained by numerically integrating the orbits of Jupiter, Saturn, Uranus and Neptune for 10 Myr, and computing the average of orbital elements over this interval. The mean inclinations are given with respect to the invariant plane. j 5 6 7 8 gj (arcsec yr−1 ) 4.24 28.22 3.08 0.67 sj (arcsec yr−1 ) - -26.34 -2.99 -0.69 Table 2: Secular frequencies of giant planets in the Solar System. The frequencies were obtained by numerically integrating the orbits for 10 Myr, and Fourier analyzing the results. The s5 frequency vanishes when the inclinations are referred to the invariant plane. 8 7 6 5 - 0.002 Jupiter 0.044 0.015 - 0.002 0.033 0.048 Saturn 0.002 0.029 0.038 0.002 Uranus Neptune 0.004 0.009 0.002 - Table 3: Secular amplitudes ej k , where j denotes individual planets (5 to 8 from Jupiter to Neptune). The proper modes of each planet’s orbit are denoted in bold. The amplitudes were obtained by numerically integrating the orbits for 10 Myr, and Fourier analyzing the results. Values lower than 0.001 are not reported. – 32 – 7 6 8 0.06 Jupiter 0.36 0.07 Saturn 0.90 0.05 0.06 0.04 1.02 Uranus 0.06 0.12 0.66 - Neptune Table 4: Secular amplitudes ij k . The proper modes of each planet’s orbit are denoted in bold. The amplitudes are reported in degrees. They were obtained by numerically integrating the orbits for 10 Myr, and Fourier analyzing the results. Values lower than 0.01◦ are not reported. The s5 frequency vanishes in the invariant plane and gives no contribution here. – 33 – Mdisk (MEarth ) 35 35 50 50 50 50 75 100 35 50 35 50 35 50 35 50 35 50 35 50 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 10 11 3 3 40 27 A B C D rout B(j ) Nsim ∆ (AU) % % % % (AU) (3:2,3:2,4:3), a4 = 11.6 AU 0.5 30 0 30 13 13 30 1 30 1.0 37 30 0 30 0.5 1.0 30 1 100 27 13 30 1 30 3.5 10 30 1 30 5.0 1.0 26 1 30 67 1.0 80 30 1 25 (3:2,3:2,3:2), a4 = 12.3 AU 1.0 30 1 30 40 1.0 39 100 1 30 (3:2,4:3,3:2), a4 = 11.9 AU 1.0 30 1 30 7 1.0 30 1 30 10 (3:2,2:1,2:1), a4 = 18.9 AU 1.0 30 1 100 100 1.0 30 1 100 100 (3:2,2:1,3:2), a4 = 18.9 AU 30 1 30 1.0 1.0 30 1 30 (2:1,3:2,3:2), a4 = 14.8 AU 100 100 1 30 1.0 1.0 30 1 100 93 (2:1,3:2,4:3), a4 = 13.7 AU 63 100 1 30 1.0 1.0 30 1 100 100 0 0 0 0 3 0 0 0 0 4 3 3 0 0 0 0 0 0 33 87 13 50 0 0 0 0 88 89 0 0 0 0 0 0 0 0 11 0 Table 5: The results of selected four-planet models. The columns are: the (1) mass of the planetesimal disk (Mdisk ), (2) ∆ defined as rin − an , where rin is the initial radial distance of the inner edge of the planetesimal disk and an is the semima jor axis of the outer ice giant, (3) radial distance of the outer edge of the planetesimal disk (rout ), (4) B(j ) specifying the instability trigger (see Sect. 2 for a definition), (5) number of simulations performed for each case (Njob ), and (6-9) success rate for our four criteria defined in Sect. 3. – 34 – Mdisk (MEarth ) 35 50 50 35 35 50 50 20 20 35 35 35 35 35 20 35 20 20 0 0 0 0 4 0 3 3 0 3 3 8 3 3 3 16 17 3 23 10 rout B(j ) Nsim A B C D ∆ (AU) % % % % (AU) (3:2,3:2,4:3,5:4), a5 = 13.9 AU 1.0 30 1 30 13 37 30 1 30 1.0 3.5 30 1 30 23 (3:2,3:2,3:2,3:2), a5 = 16.1 AU 23 30 0 30 1.5 33 100 1 30 1.5 1.5 30 1 100 30 (3:2,3:2,4:3,4:3), a5 = 14.5 AU 1.0 30 1 30 47 (3:2,3:2,2:1,3:2), a5 = 22.2 AU 1.0 30 0 30 33 30 30 1 30 1.0 1.0 30 1 30 33 (3:2,3:2,2:1,2:1), a5 = 24.5 AU 43 30 0 30 1.0 23 30 1 30 1.0 2.0 30 1 30 30 3.0 44 100 1 30 (2:1,3:2,3:2,3:2), a5 = 19.3 AU 1.0 30 1 30 53 1.0 53 30 1 30 (2:1,4:3,3:2,3:2), a5 = 17.9 AU 1.0 30 1 30 67 75 30 1 30 3.5 17 13 23 19 13 10 17 3 7 10 7 7 0 7 3 3 3 20 17 17 20 7 3 3 3 23 36 43 42 25 3 0 0 0 Table 6: The results of selected five-planet models. See the caption of Table 5 for a definition of different parameters shown here. – 35 – Mdisk (MEarth ) 20 35 20 35 10 20 20 7 0 rout B(j ) Nsim A B C D ∆ (AU) (AU) % % % % (3:2,3:2,3:2,4:3,3:2), a6 = 20.4 AU 23 30 1 30 1.0 1.0 30 1 30 40 (3:2,4:3,3:2,3:2,3:2), a6 = 20.6 AU 1.0 30 1 100 30 10 1.0 8 42 30 1 30 (2:1,3:2,4:3,3:2,3:2), a6 = 24.2 AU 1.0 30 1 30 69 25 1.0 28 44 30 1 30 (2:1,3:2,4:3,3:2,3:2), a6 = 24.9 AU 1.0 30 1 30 31 12 6 20 6 3 3 3 0 3 0 7 0 3 3 7 Table 7: The results of selected six-planet models. See the caption of Table 5 for a definition of different parameters shown here. – 36 – Fig. 1.— Orbit histories of the giant planets in a simulation with four initial planets. The four planets were started in the (3:2,4:3,3:2) resonant chain, Mdisk = 35 MEarth and B(1) (see Sect. 2 for the definition of B(1)). (a) The semima jor axes (solid lines), and perihelion and aphelion distances (dashed lines) of each planet’s orbit. The red, green, turquoise and blue lines correspond to Jupiter, Saturn, Uranus and Neptune. The black dashed lines show the semima jor axes of planets in the present Solar System. (c) The period ratio PSat/PJup . The dashed line shows PSat/PJup = 2.49, corresponding to the period ratio in the present Solar System. The shaded area approximately denotes the zone where the secular resonances with the terrestrial planets occur. (b) Jupiter’s eccentricity. (d) Jupiter’s inclination. The dashed lines in (b) and (d) show the present mean eccentricity and inclination of Jupiter. – 37 – Fig. 2.— Final orbits obtained in our simulations with four planets started in the (3:2,3:2,4:3) resonant chain, Mdisk = 75 MEarth and B(1) (see Sect. 2 for the definition of B(1)). (a) Mean eccentricity. (b) Mean inclination. The mean orbital elements were obtained by averaging the osculating orbital elements over the last 10 Myr (i.e., from 90 to 100 Myr). Only the systems ending with four planets are plotted here (dots). The bars show the mean and standard deviation of the model distribution of orbital elements. The mean orbits of real planets are shown by triangles. Colors red, green, turquoise and blue correspond to Jupiter, Saturn, Uranus and Neptune. – 38 – Fig. 3.— Orbit histories of the giant planets in a simulation with four initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The four planets were started in the (3:2,3:2,4:3) resonant chain, and Mdisk = 75 MEarth . – 39 – Fig. 4.— Orbit histories of the giant planets in a simulation with four initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The four planets were started in the (3:2,3:2,4:3) resonant chain, Mdisk = 50 MEarth and ∆ = 1 AU. – 40 – Fig. 5.— Final orbits obtained in our simulations with four planets started in the (2:1,3:2,3:2) resonant chain, and Mdisk = 35 MEarth . See the caption of Fig. 2 for the description of orbital parameters shown here. – 41 – Fig. 6.— Final orbits obtained in our simulations with five planets started in the (3:2,3:2,4:3,5:4) resonant chain, Mdisk = 50 MEarth and ∆ = 1 AU. See the caption of Fig. 2 for the description of orbital parameters shown here. – 42 – Fig. 7.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,4:3,5:4) resonant chain, Mdisk = 50 MEarth and ∆ = 1 AU. The fifth planet was ejected at t = 0.8 Myr after the start of the simulation. – 43 – Fig. 8.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,3:2,3:2) resonant chain, Mdisk = 35 MEarth and B(1). The fifth planet was ejected at t = 17 Myr after the start of the simulation. – 44 – Fig. 9.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,3:2,3:2) resonant chain, Mdisk = 35 MEarth and B(0) The fifth planet was ejected at t = 15.2 Myr after the start of the simulation. – 45 – Fig. 10.— Final orbits obtained in our simulations with five planets started in the (3:2,3:2,3:2,3:2) resonant chain, Mdisk = 35 MEarth and B(1). See the caption of Fig. 2 for the description of orbital parameters shown here. – 46 – Fig. 11.— Final orbits obtained in our simulations with five planets started in the (3:2,3:2,2:1,3:2) resonant chain, and Mdisk = 20 MEarth . Results obtained with B(0) and B(1) for ∆ = 1 AU were combined here. See the caption of Fig. 2 for the description of orbital parameters shown here. – 47 – Fig. 12.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,2:1,3:2) resonant chain, Mdisk = 20 MEarth and B(1). The fifth planet was ejected at t = 34.4 Myr after the start of simulation. – 48 – Fig. 13.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,2:1,3:2) resonant chain, Mdisk = 20 MEarth and B(0). The fifth planet was ejected at t = 14.9 Myr after the start of the simulation. – 49 – Fig. 14.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (3:2,3:2,2:1,3:2) resonant chain, Mdisk = 20 MEarth and B(1). The fifth planet was ejected at t = 6.1 Myr after the start of the simulation. – 50 – Fig. 15.— Final orbits obtained in our simulations with five planets started in the (2:1,3:2,3:2,3:2) resonant chain, and Mdisk = 20 MEarth . See the caption of Fig. 2 for the description of orbital parameters shown here. – 51 – Fig. 16.— Orbit histories of the giant planets in a simulation with five initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The five planets were started in the (2:1,3:2,3:2,3:2) resonant chain, and Mdisk = 20 MEarth . The fifth planet was ejected at t = 5.6 Myr after the start of the simulation. – 52 – Fig. 17.— Final orbits obtained in our simulations with six planets started in the (3:2,4:3,3:2,3:2,3:2) resonant chain, and Mdisk = 20 MEarth . See the caption of Fig. 2 for the description of orbital parameters shown here. – 53 – Fig. 18.— Orbit histories of the giant planets in a simulation with six initial planets. See the caption of Fig. 1 for the description of orbital parameters shown here. The six planets were started in the (3:2,4:3,3:2,3:2,3:2) resonant chain, and Mdisk = 20 MEarth . The fifth and sixth planet were ejected at t = 3.18 and 3.65 Myr after the start of simulation (yellow and purple lines in (a)).
1506.02870
2
1506
2015-09-01T14:26:30
Rigorous treatment of the averaging process for co-orbital motions in the planetary problem
[ "astro-ph.EP", "math-ph", "math.DS", "math-ph" ]
We develop a rigorous analytical Hamiltonian formalism adapted to the study of the motion of two planets in co-orbital resonance. By constructing a complex domain of holomorphy for the planetary Hamilto-nian, we estimate the size of the transformation that maps this Hamil-tonian to its first order averaged over one of the fast angles. After having derived an integrable approximation of the averaged problem, we bound the distance between this integrable approximation and the averaged Hamiltonian. This finally allows to prove rigorous theorems on the behavior of co-orbital motions over a finite but large timescale.
astro-ph.EP
astro-ph
Rigorous treatment of the averaging process for co-orbital motions in the planetary problem Philippe Robutel∗ Laurent Niederman† ,∗ Alexandre Pousse∗ October 18, 2018 Abstract We develop a rigorous analytical Hamiltonian formalism adapted to the study of the motion of two planets in co-orbital resonance. By con- structing a complex domain of holomorphy for the planetary Hamilto- nian, we estimate the size of the transformation that maps this Hamil- tonian to its first order averaged over one of the fast angles. After having derived an integrable approximation of the averaged problem, we bound the distance between this integrable approximation and the averaged Hamiltonian. This finally allows to prove rigorous theorems on the behavior of co-orbital motions over a finite but large timescale. 1 Introduction Averaging methods are common techniques to study the dynamics of Hamil- tonian systems in celestial mechanics. The first and most famous example of averaged Hamiltonian system is perhaps the secular planetary problem, where the Hamiltonian of the planetary problem is averaged over the mean longitudes of the planets. The secular equations of the planetary motion appear in Lagrange's work on stability of the solar system (Lagrange, 1778) while the secular Hamiltonian appears in Delaunay's memory about the the- ory of the Moon (Delaunay, 1860). Poincar´e (1892) gave an expression of the secular Hamiltonian of the planetary three body problem while Laskar and Robutel (1995) present an analytical method allowing to compute the ∗IMCCE, Observatoire de Paris, UPMC, CNRS UMR8028, 77 Av. Denfert-Rochereau, 75014 Paris, France †Universit´e Paris XI, LMO, ´equipe Topologie et Dynamique, Batiment 425, 91405 Orsay, France 1 expansion of the planetary Hamiltonian, and in particular, to get a concise expression of its secular part. The secular Hamiltonian of the planetary problem can be obtained, up to a finite order of the planetary masses, by averaging over the planetary mean longitudes (construction of a resonant normal form), the symplectic transformation that maps the non-averaged Hamiltonian to the secular one being close to the identity. Precise estimates on the size of these transfor- mations are required in order to prove the existence of invariant tori using the KAM theory. These kinds of rigorous estimates have been established especially by Arnold (1963), F´ejoz (2004) and Chierchia and Pinzari (2011). When two planets are in co-orbital resonance, and more generally for two planets in mean-motion resonance, the transformation leading to the secular Hamiltonian is no more close to the identity, even at first order. Consequently, the secular motion does not provide a good representation of the real planetary motion. In these cases, it is still possible to use averaging methods, but for two planets the Hamiltonian is generally averaged over one fast angle, that is one of the planetary mean longitudes. Many authors work with the resonant averaged problem. Some of them use analytic approximations of the averaged Hamiltonian or averaged motion (e.g. ´Erdi, 1977; Namouni, 1999; Morais, 2001; Robutel and Pousse, 2013), while others prefer a numerical averaging (e.g. Nesvorn´y et al., 2002; Giup- pone et al., 2010). But in none of these works the size of the transformation and consequently the "distance" between the secular and the "complete" solution is rigorously estimated. In this paper, we estimate this distance and give an upper bound of the time for which this difference remains small enough. For this purpose, we derive, in section 2, a complex domain of holomorphy for the planetary Hamiltonian which allows to compute quantitatively the size of the trans- formations and of the perturbations involved in our construction. All the computations about perturbation methods are derived in section 3 and constitute the main novelty of this paper. The topology of the averaged Hamiltonian is studied in section 4 where we show the existence of an invariant manifold, associated to the quasi- circular motions, which carry an integrable dynamic. Section 5 is devoted to the construction of an integrable approximation of the averaged Hamiltonian and its degree of accuracy in the vicinity of the invariant manifold considered in section 4. We also give the general form of the solutions of this integrable system. Finally, in section 6, we can combine the quantitative estimates given in section 3 and the bounds on the remainders between the averaged Hamilto- 2 nian and its integrable approximation. This allows to prove rigorous theo- rems on the behaviour of co-orbital motions over large timescales. The last section concerns the proof of the technical propositions and lemma used in our reasonings. 2 Hamiltonian setting of the problem 2.1 Canonical heliocentric coordinates We consider two planets of respective masses (cid:101)m1 and (cid:101)m2 orbiting a central body (Sun, or star) of mass m0 dominant with respect to the planetary masses. As only co-orbital planets are considered, no planet is permanently farther from the central body than the other, so the heliocentric coordinate system seems to be the most adapted to this situation. Following Laskar and Robutel (1995), the Hamiltonian of the three-body problem reads (cid:101)H((cid:101)Rj, rj) = (cid:101)HK((cid:101)Rj, rj) + (cid:101)HP ((cid:101)Rj, rj) with (cid:33) − (cid:101)µj(cid:107)rj(cid:107) (cid:101)HK((cid:101)Rj, rj) = − G (cid:101)m1(cid:101)m2 (cid:88) (cid:101)HP ((cid:101)Rj, rj) = (cid:32)(cid:101)R2 2(cid:101)βj (cid:101)R1 · (cid:101)R2 j∈ {1,2} and j (1) (cid:107)r1 − r2(cid:107) , m0 where rj is the heliocentric position of the planet j, (cid:101)βj = m0(cid:101)mj(m0 + (cid:101)mj)−1 and(cid:101)µj = G(m0 +(cid:101)mj), G being the gravitational constant. The conju- gated variable of rj, denoted by (cid:101)Rj, is the barycentric linear momentum of the body of index j. In this expression, (cid:101)HK corresponds to the unperturbed (cid:101)βj around a fixe center of mass m0 +(cid:101)mj, while (cid:101)HP models the gravitational The Hamiltonian (cid:101)H is analytical on the whole phase space (R8 since we Keplerian motion of the two planets, more precisely the motion of a mass perturbations. consider the planar problem) except on the manifold which corresponds to collisions between two planets: (cid:110) ((cid:101)R1, r1,(cid:101)R2, r2) ∈ R8 such that r1 (cid:54)= r2 (cid:101)D = (cid:111) . If we introduce the small parameter ε given by ε = Max , (2) (cid:19) (cid:18)(cid:101)m1 m0 , (cid:101)m2 m0 3 one can verify that the Keplerian term of the planetary Hamiltonian is of order ε and that the other one is of order ε2 as long as the mutual distance between the planets remain large enough. This feature justifies a perturbative approach. Remark 1 In the sequel, we will not give explicit estimates of the constants independent of the small parameters involved in the problem in order to avoid cumbersome and meaningless expressions. We will just denote uniformly by M a positive constant chosen sufficiently large for our purpose but indepen- dent of the relevant quantities, we will try to specify this at each occurrence of such constant but sometimes it will be omitted. 2.2 The rescaled heliocentric coordinates According to (1), the momenta (cid:101)Rj and the Keplerian part of the Hamil- tonian are of order ε while the perturbation is quadratic in ε. In order to get more homogeneous quantities, it is convenient to rescale the planetary masses, the Hamiltonian and the canonical variables. the relations: First, we introduce new planetary masses and new reduced masses by , m2 = (cid:101)m2 m1 = (cid:101)m1 Moreover, for j ∈ {1, 2}, we rescale the impulsions by (cid:101)Rj = εrj while the and µj =G(m0+εmj) for j∈{1, 2}. m0 + εmj ; βj = (cid:101)βj ε m0mj = ε ε positions rj remain unchanged. Hence, we have made a conformal symplectic transformation T which changes the symplectic form with (3) 4 and the Hamiltonian linked to the considered system becomes This leads to the expression: (cid:88) j j drj ∧ drj = ε−1(cid:88) d(cid:101)Rj ∧ drj H(rj, rj) = ε−1(cid:101)H ◦ T (rj, rj). (cid:33) (cid:32) r2 H(rj, rj) = HK(rj, rj) + εHP (rj, rj) with HK(rj, rj) = (cid:88) and j∈ {1,2} HP (rj, rj) = j 2βj r1 · r2 m0 − µjβj (cid:107)rj(cid:107) − G m1m2 (cid:107)r1 − r2(cid:107) , the canonical variables rj and rj and the Keplerian part HK being now of order one. The three body Hamiltonian H is defined on D =(cid:8)(r1, r1, r2, r2) ∈ R8 such that r1 (cid:54)= r2 (cid:9) since the positions remain unchanged. 2.3 The Poincar´e variables. In order to define a canonical coordinate system related to the elliptic el- ements (aj, ej, λj, j) (respectively the semi-major axis, the eccentricity, the mean longitude and the longitude of the pericenter of the planet j), we use complex Poincar´e's rectangular variables (λj, Λj, xj,−ixj)j∈{1,2} ∈ (T × R × C × C)2: √ and xj =(cid:112)Λj (cid:114) 1 −(cid:113) 1 − e2 (4) Λj = βj µjaj j exp(ij), this coordinate system has the advantage to be regular when the eccentric- ities and the inclinations tend to zero. Consequently, we have the product of the analytic symplectic transfor- (cid:40) mations around the circular orbits (i.e. for a given constant c0 > 0): Φj : Ej (λj, Λj, xj,−ixj) −→ C4 (cid:55)−→ (rj, rj) (j = 1, 2) (5) (cid:26) with Ej = (λj, Λj, xj,−ixj) where (cid:12)(cid:12)(cid:12)(cid:12) λj ∈ T Λj ∈ R∗+ and xj ∈ C with xj ≤ c0 (cid:27) (cid:112)Λj which yields the new planetary Hamiltonian: (cid:101)H(λj, Λj, xj,−ixj) = (cid:101)HK(Λ1, Λ2) + (cid:101)HP (λj, Λj, xj,−ixj). (cid:101)H is analytic on the domain A ⊂ (T × R × C × C)2 defined by: (cid:18) Φ1(λ1, Λ1, x1,−ix1) (cid:19) (λj, Λj, xj,−ixj) ∈ Ej for j ∈ {1, 2} / (cid:26) Φ2(λ2, Λ2, x2,−ix2) (cid:27) ∈ D 5 2.4 The 1:1 Resonance The Keplerian part of the Hamiltonian expressed in terms of Poincar´e's variables reads: HK(rj, rj) = (cid:101)HK(Λ1, Λ2) = − (cid:88) 1 2 µ2 j β3 j Λ2 j (6) j∈ {1,2} Consequently, the 1:1 mean motion resonance, also called co-orbital reso- nance, is reached when (Λ1, Λ2) = (Λ0 2) such that the two mean motions are the same (let us denoted by ω this frequency), that is: 1, Λ0 ∂Λ1 or: (7) (8) (9) ∂(cid:101)HK ∂(cid:101)HK ∂Λ2 (Λ0 1, Λ0 2) = (Λ0 1, Λ0 2) := ω > 0 , 1β3 µ2 1 1)3 = (Λ0 2β3 µ2 2 2)3 := ω > 0 . (Λ0 In order to work in a neighborhood of this resonance, we introduce a new coordinate system with an affine unimodular transformation completed in a symplectic way Ψ(ζ1, ζ2, Z1, Z2, x1, x2) = (λ1, λ2, Λ1, Λ2, x1, x2) with: (cid:18)ζ1 (cid:19) ζ2 (cid:18)1 −1 (cid:19)(cid:18)λ1 (cid:19) 0 1 λ2 , = (cid:18)Z1 (cid:19) Z2 (cid:18)1 1 = (cid:19) (cid:19)(cid:18)Λ1 − Λ0 1 Λ2 − Λ0 2 0 1 which yields slow-fast angles. More precisely, in a neighborhood of the co-orbital resonance, the angular variables evolve at different rates: ζ2 is a "fast" angle with a frequency of order 1, ζ1 undergoes "semi-fast" variations at a frequency of order ε (see Section 5.2), while the variables xj related to the eccentricities are associated to the slow degrees of freedom evolving on a time scale of order ε (secular variations). With this set of variables, the planetary Hamiltonian becomes: √ H(ζj, Zj, xj,−ixj) =(cid:101)H ◦ Ψ(ζj, Zj, xj,−ixj) = HK(Z1, Z2) + HP (ζj, Zj, xj,−ixj). (10) For an arbitrary ∆ > 0, we consider the following domain centred at the 1:1 keplerian resonance defined in the (ζ, Z, x,−ix) variables: ∆ =(cid:8)((ζ1, 0, 0, 0), (ζ2, 0, 0, 0))∈ (T × R × C × C)2 such that ζ1 > ∆(cid:9) K(0) 6 where . denotes the usual distance over the quotient space T = R/2πZ. Hence, the angular separation ∆ over K(0) ∆ yields a minimal distance between the planets: (cid:18) ∆ (cid:19) 2 (cid:18)(Λ0 1)2 β2 1µ1 , (Λ0 2)2 β2 2µ2 (cid:19) δ := 2a sin where a := min (cid:27) (cid:111) is the lowest radius of the orbits. The condition ζ1 > ∆ for ζ1 ∈ T can also be considered with the real variable ζ1 ∈]∆, 2π − ∆[ since there exists an unique real representative in this interval for an angle with a modulus lowered by ∆. Hence K(0) ∆ has the structure of a cylinder in (]∆, 2π − ∆[×R × C × C) × (T × R × C × C). For ρ > 0 and σ > 0 small enough (this will be specified in the se- ∆ can be extended in a complex neighbourhood quel), our initial domain K(0) K(C) ∆,ρ,σ ⊂ C8 of the following type: (ζj, Zj, ξj, ηj)j∈{1,2}∈C8 / and Zj ≤ ρ, ξj ≤ √ (cid:19) (cid:18) (Re(ζ1), 0, 0, 0) ρσ, ηj ≤ √ (Re(ζ2), 0, 0, 0) ρσ, Im(ζj) ≤ σ ∈K(0) ∆ (cid:26) K(C) ∆,ρ,σ = actually K(0) ∆ ⊂ K(C) ∆,ρ,σ ∩ A = (cid:110) (ζj, Zj, ξj, ηj)j∈{1,2}∈ K(C) ∆,ρ,σ with ηj = iξj = xj . In this setting, we can define a complex domain of holomorphy for the planetary Hamiltonian where it will be possible to estimate the size of the transformations and the functions involved in our construction of a 1:1 res- onant normal form. More specifically, for ∆ > 0, ρ > 0, σ > 0 and for p > 0 we will consider the compact: Kp := K(C) ∆,pρ,pσ and the supremum norm .∞ on the space of holomorphic functions over the compact Kp which will be denoted .p. With HK of class C(3) on the compact {(z1, z2) ∈ C2 / max(z1,z2) ≤ ρ}, there exists a constant (cid:101)M > 0 which satisfies: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)∂Zp1 1 ,Zp2 2 HK (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ρ < (cid:101)M (11) ∀ (p1, p2) ∈ N2 such that p1 +p2 ≤ 3, one has: 7 where we denote .ρ the supremum norm .∞ on the space of holomorphic functions over the compact {(z1, z2) ∈ C2 / max(z1,z2) ≤ ρ}. By the real-analyticity of the transformation in Poincar´e resonant action- angle variables Υ = (Φ1 ◦ Ψ, Φ2 ◦ Ψ), there exist ρ0 > 0 and σ0 > 0 small enough such that Υ can be extended into a holomorphic function over the compact K(C) with: (cid:40) ∆,ρ0,σ0 Φj ◦ Ψ : K(C) ∆,ρ0,σ0 (ςj, zj, ξj, ηj)j∈{1,2} −→ C4 (cid:55)−→ (rj, rj) for j = 1, 2 and the differential of Ψ is bounded by a constant C > 0 with respect to the supremum norm .∆,ρ0,σ0 on the space of holomorphic functions over the compact K(C) Theorem 1 There exist constants ρ0 > 0, σ0 > 0 and M ≥ (cid:101)M independent Without loss of generality, we can assume that C > 1. ∆,ρ0,σ0 . of ε and ∆ (or equivalently of δ) such that if we assume: ρ ≤ ρ0, σ ≤ σ0, 0 < ρ < σ < 1 and σ ≤ δ 16C = a 8C sin then the following bounds are valid: (cid:18) ∆ (cid:19) 2 < δ (12) ε M < InfK1 (HP (ςj, zj, ξj, ηj)) and HP1 < M ε δ where we denote .1 the supremum norm .∞ on the space of holomorphic functions over the compact K1 = K(C) ∆,ρ,σ. Remark: As it was specified, from now on we will denote uniformly by M a positive constant independent of ε, ρ, σ and ∆ (or equivalently of δ) high enough for our purpose. We will try to specify this at each occurrence of such constant but sometimes it will be omitted 3 Hamiltonian perturbation theory In this paper, we only consider the averaged Hamiltonian at first order in the planetary masses. More precisely, we prove quantitatively that there 8 exists a canonical transformation C which maps the original Hamiltonian H in , Z. j, x. j,−ix. j) = HK(Z. 1, Z. 2)+H P (ζ. 1 H(cid:48)(ζ. j In this expression, the Keplerian part HK reads , Z. j, x. j,−ix. j)+H∗(ζ. j , Z. j, x. j,−ix. j). HK(Z. 1, Z. 2) = − 1µ2 β3 1 + Z. 1)2 − 1 2(Λ0 β3 2µ2 2 2(Λ0 2 − Z. 1 + Z. 2)2 , (13) while the averaged perturbation with respect to the second angle (which corresponds to a time averaging along the periodic orbits on the torus at the origin for the Kepler problem in our resonant action-angle variables) is given by: (cid:90) 2π (cid:90) 2π 0 ω 0 H P (ζ. 1 , Z. j, x. j,−ix. j) = ω 2π = 1 2π HP (ζ. 1 , ζ. 2 HP (ζ. 1 , ζ. 2 + ωt, Z. j, x. j,−ix. j)dt , Z. j, x. j,−ix. j)dζ. 2 (14) The remainder H∗(ζ. j whose size will be estimated in the next section. , Z. j, x. j,−ix. j) is a much smaller general perturbation 3.1 Hamiltonian perturbation theory Now, we specify our construction of the averaging transformation which will be the time-one map C = Φχ 1 of the Hamiltonian flow generated by some auxiliary function χ. Using the Poisson bracket: Lχ(f ) = {χ, f} = ∂1χ.∂2f + ∂3χ.∂4f − ∂2χ.∂1f − ∂4χ.∂3f, we have C = exp(Lχ) and H = exp(Lχ)(H). Hamiltonian can be written: In the new variables, the H(cid:48) = HK + HP + {χ, HK} + {χ, HP} + H(cid:48) − H − {χ, H}. With a generating function χ comparable with HP , the terms of order 1 in this expansion (with respect to HP ) are : [HP + {χ, HK}](ζ. j , Z. j, x. j,−ix. j) = HP (ζ. j , Z. j, x. j,−ix. j) + ∇HK(Z. 1, Z. 2). , Z. j, x. j,−ix. j) ∂χ ∂ζ. (ζ. j 9 We will choose χ as a solution, over K1 = K(C) ∆,ρ,, of the equation: actually, this equation is satisfied by : ω. ∂χ ∂ζ. 2 (ζ. j , Z. j, x. j,−ix. j) = H P (ζ. 1 (cid:90) 2π ω ω , Z. j, x. j,−ix. j) = t. 2π 0 , Z. j, x. j,−ix. j) , Z. j, x. j,−ix. j) − HP (ζ. j (cid:104) H P (ζ. 1 −HP (ζ. 1 , Z. j, x. j,−ix. j) , ζ. 2 (cid:105) + ωt, Z. j, x. j,−ix. j) χ(ζ. j and dt (cid:19) (cid:18) 0 ω , Z. j, x. j,−ix. j) = H(cid:48)(ζ. j HK(Z. 1, Z. 2) +H P (ζ. 1 with H∗ = (∇HK−−→ω ) . ∂χ ∂ζ. , Z. j, x. j,−ix. j) , Z. j, x. j,−ix. j) + H∗(ζ. j + {χ, HP} + H(cid:48) − H − {χ, H} for −→ω := 3.2 Quantitative Hamiltonian perturbation theory Here, we make the construction described in the previous section with ac- curate estimates on the size of the Hamiltonian and on the size of the trans- formations which are involved. As in the section 2.4, for ∆ > 0, ρ > 0 and σ > 0, we consider the compact K2 and the supremum norm .2 on the space of holomorphic functions over K2. With the expression of χ and the fact that H P2 ≤ HP2, we obtain : χ2 < HP2 2π ω (15) which allows to prove the following: Theorem 2 Let ∆ > 0, ρ > 0 and σ > 0 such that (∆, 2ρ, 2σ) satisfy the condition (12), hence: (cid:19) (cid:18)σ0 , δ 32C 2 ρ < ρ0 2 ; ρ < σ < min (cid:18)(Λ0 (cid:18)∆ (cid:19) 2 min 1)2 β2 1µ1 , 2)2 (Λ0 β2 2µ2 . (16) (cid:19) for δ := sin 10 In order to get a small enough transformation (or equivalently a small enough χ), we assume moreover that ε δρσ < ω 32πM . Then there exists a canonical transformation C : K3 ⊆ C(K3 C is one−to−one and K5 → K2 such that ) ⊆ K7 2 4 2 4 and, still using the notations (ζj, Zj, xj,−ixj)j∈{1,2} = C(cid:16) (cid:17) , Z. j, x. j,−ix. j ζ. j j∈{1,2} , (17) (18) the transformed Hamiltonian H(cid:48) = H ◦ C can be written as H(cid:48)(ζ. j with H P2 ≤ 2M , Z. j, x. j,−ix. j) = HK(Z. 1, Z. 2)+H P (ζ. 1 and : ε δ , Z. j, x. j,−ix. j)+H∗(ζ. j (cid:19) (cid:18) ρ M π ε + ω σ ρσδ , Z. j, x. j,−ix. j) (19) H∗3/2 ≤ ηM ε δ with η = 40 (cid:114) ε δ √ √ ε δ σ . M π ω The proof of this theorem is given in section 7.1. The decreasing factor η in the averaged perturbation can be minimized for a fixed lower bound on the mutual distance δ > 0 and a mass ratio ε. We first relate the analyticity width ρ to δ and ε by choosing ρ = such that the two terms in the factor η are of the same order, then η = 80 For δ ≤ 16Cσ0, the maximal analyticity width σ = factor: 32C δ gives the minimal min(η) = 2560 CM π ω δ3 for ρ = and σ = δ 32C (cid:114) ε (cid:114) ε δ which imposes a lower bound δ ≥ O(ε1/3) in order to get a decreased pertur- bation in the averaged system. Hence we recover the size of the Hill region inside which the averaged heliocentric Hamiltonian is not close to the initial one. 11 4 The averaged Hamiltonian and its topology 4.1 Lagrange and Euler configurations in the averaged sys- tem Euler and Lagrange configurations are central configurations of the three body problem1. The two Lagrange configurations correspond to the situ- ation where the three bodies occupy the vertices of an equilateral triangle (the equilibrium points L4 and L5 in the circular RTBP), while the three Euler's ones are the aligned configurations (L1, L2 and L3 in the circular RTBP). In terms of heliocentric elliptic elements, these are represented by two homothetical ellipses. The motion of the planets on these fixed ellipses is Keplerian and the difference of their mean longitude ζ1 = λ1 − λ2 is constant. More precisely, in the Lagrange's case, we have: ζ1 = λ1 − λ2 = 1 − 2 = ±π/3, e1 = e2, a1 = a2, (20) (21) (22) while Euler configurations lead to the relations ζ1 = λ1 − λ2 = 1 − 2 = 0, e1 = e2, a1 = a2 + O(ε1/3), if the two planets are on the same side of the Sun (L1, L2), and to ζ1 = λ1 − λ2 = 1 − 2 = π, e1 = e2, a1 = a2 + O(ε), if the Sun is between the planets (L3). In the both cases, these trajectories, in fixed reference frame, are periodic orbits whose period is the common mean motion of the planets. As a result, these central configurations corre- spond to fixed points of the averaged problem. More precisely, each type of configuration (L1 to L5) defines, in the averaged problem, a one-parameter family of equilibrium points. This implies that a given fixed point of one of these families possesses an eigenvector (tangent to the family) associated to a zero eigenvalue. This is the source of the degeneracy that will be discussed in section 5.3. 4.2 Some properties of the averaged Hamiltonian In this section we will study the main properties of the averaged Hamiltonian of order 1: H(ζ. 1 , Z. j, x. j,−ix. j) = H(cid:48)(ζ. j , Z. j, x. j,−ix. j) − H∗(ζ. j , Z. j, x. j,−ix. j) = HK(Z. 1, Z. 2)+H P (ζ. 1 , Z. j, x. j,−ix. j) (23) 1See the webpages by A. Chenciner (2012) and by R. Moeckel (2014) at www.scholarpedia.org/article/Three body problem, and at www.scholarpedia.org/article/Central configurations#Euler. 12 and of its associated dynamics. The Hamiltonian H P being an analytic function on the domain K3 , it can be expanded in Taylor series in a neighborhood of (x. 1, x. 2) = (0, 0) as: 2 (cid:88) (p,¯p)∈N4 H P (ζ. 1 , Z. j, x. j,−ix. j) = Cp,q(ζ. 1 , Z. 1, Z. 2)x. p1 1 x. p2 2 x. ¯p1 1 x. ¯p2 2 . (24) In the previous summation, only the coefficients Cp,q satisfying the relation p1 + p2 = ¯p1 + ¯p2 (25) are different from zero. This propriety, known as D'Alembert rule, is equiv- alent to the fact that the quantity x. 12 + x. 22 (26) is an integral of the averaged motion. The existence of this constant of motion makes possible the reduction of the averaged problem, leading to a Hamiltonian system which depends on two angles: the difference of the mean longitudes and the difference of the longitudes of the perihelion (see Giuppone et al., 2010). This reduction, which decreases the number of degrees of freedom of the averaged problem from 3 to 2, introduces some technical issues (addition of a parameter, singularity when the eccentricities tend to zero). For this reason, we prefer not to reduce the problem. Besides the reduction, the relation (25) has many implications on the dynamics of the averaged system. One of these properties lies in the fact that the manifold CC 0 = {(ζ. j , Z. j, x. 1, x. 2, x. 1, x. 2) ∈ K3 2 , such that x. j = x. j = 0} (27) is an invariant manifold of the averaged Hamiltonian (23). Indeed, the relation (25) implies that the Taylor series (24) starts at degree two. As a consequence ∂x. jH P (ζ. 1 , Z. j, 0, 0) = ∂x. jH P (ζ. 1 which proves the invariance of the manifold CC Actually, with the previous definition, CC , Z. j, 0, 0) = 0, (28) 0 by the flow of (23). 0 is a complex manifold and we will consider the real manifold C0 = CC 0 ∩ (]∆, 2π − ∆[×R × C × C) × (T × R × C × C) (29) which is also invariant by the flow of (23) since it is a real Hamiltonian. 13 4.3 The invariant manifold C0 4.3.1 Hamiltonian dynamics on C0 The dynamics on C0 is given by the restriction of the averaged Hamiltonian (23) to this manifold, that is: (cid:48) 0(ζ. 1 H , Z. 1, Z. 2) = HK(Z. 1, Z. 2) +H P (ζ. 1 , Z. 1, Z. 2, 0, 0, 0, 0) (30) with H 0 = − β1µ1 2a1 − β2µ2 2a2 +εGm1m2 and ∗ 0 = ε H Gm1m2 √ a1a2 the expression: a1 = µ−1 1 β−2 1 (cid:0)Λ0 (cid:1)2 (31)  (32) 1 a2 1 + a2 2 − 2a1a2 cos ζ. 1 (cid:48) 0 = H 0 +H ∗ 0, H  cos ζ. 1√ (cid:18) β1β2 a1a2 m1m2 (cid:113) − √ µ1µ2 Gm0 − 1 (cid:19) (cid:0)Λ0 cos ζ. 1 . (33) (cid:1)2 In the expression (32), aj can be easily expressed in term of Z. j using 1 + Z. 1 a2 = µ−1 2 β−2 2 , 2 + Z. 2 − Z. 1 (34) deduced from (9). With the expressions βj = mj(1 + O(ε)) (resp. µj = Gm0 + O(ε)) and x.y) over R2 the analyticity of the functions f (x, y) = x.y (resp. g(x, y) = (resp. R∗ × R∗), we can write √ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3 2 ∗ 0 14 where M > 0 is independent of the small parameters in the problem. Moreover, we have a uniform upper bound on (cid:12)(cid:12)(cid:12) cos ζ. 1√ a1a2 (cid:12)(cid:12)(cid:12) (or equivalently (cid:12)(cid:12)(cid:12)(cid:12) < M ε √ µ1µ2 Gm0 − 1 m1m2 (cid:12)(cid:12)(cid:12)(cid:12) β1β2 (cid:12)(cid:12)(cid:12)) over the compact K3 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H 2 (cid:12)(cid:12)(cid:12) cos ζ. 1 Λ1Λ2 on and gathering these estimates yield: < M ε2. (35) 4.3.2 Phase portrait on C0 The phase portrait of the Hamiltonian H 0 is represented on Fig. 1 in the plan (ζ. 1 , u) where u is a dimensionless quantity related to the action Z. 1 and to the semi-major axes by −2/3 0 (36) with µ0 = Gm0. This figure has been obtained for particular values of the masses and of the parameters given in the caption of the figure 1, but the qualitative structure of the phase space does not depend on these values. 2ω1/3µ m1 + m2 µ0 + O(ε) a1 − m2 Z. 1 = u = √ √ a2 √ m1 2 ) (cid:55)−→ (Z. 1, 2π−ζ. 1 The phase portrait ofH 0 is invariant by the symmetry with respect to the Z. 1axis (Z. 1, ζ. 1 ). When the two planetary masses are equal, the Hamiltonian H 0, and consequently its phase portrait, is also invariant by the symmetry (Z. 1, ζ. 1 ) (cid:55)−→ (−Z. 1, ζ. 1 The shaded areas indicate the outside of the co-orbital resonance. In the upper grey region where Z. 1 > 0, the angle ζ. 1 circulates clockwise while its circulation is anti-clockwise in the lower grey region (Z. 1 < 0). The others domains correspond to the different kind of resonant motions. ). The two elliptic fixed points in the middle of two green areas, located at ∈ {π/3, 5π/3} 0 ω−1/3 + O(ε2) , µ2/3 and Z. 1 = m1m2 (37) m1 − m2 m1 + m2 ε 6 m0 ζ. 1 are associated to the Lagrange equilateral configurations (see section 4.1). These points correspond to the relative equilibria where the three bodies occupy the vertices of an equilateral triangle rotation at the constant an- gular velocity ω. Each of these points, named L4 and L5 in the RTBP, is surrounded by tadpole orbits (green regions) corresponding to periodic de- formations of the equilateral triangle. These regions, whose maximal width in the Z. 1-direction is of order ε1/2 (see Robutel and Pousse, 2013, for more details) are bounded by the separatrix S3 that originates from the hyperbolic fixed point L3 at ζ. 1 = π, Z. 1 = ε 2 m1 − m2 m1 + m2 m1m2 m0 0 ω−1/3 + O(ε2) , µ2/3 (38) for which the three bodies are aligned and the Sun is between the two planets and its separatrix. Outside this curve, in the blue domain, one find the horseshoe orbits that surround the three fixed points mentioned above. Along these orbits, the angle ζ. 1 undergoes large variation such that ζ. 1 √ ∈ [ζm, 2π − ζm] with 0 < O(ε1/3) < ζm < 2 arcsin(( 2 − 1)/2) + O(ε). 15 The horseshoe region (in blue), whose vertical extend is of order ε1/3, is enclosed by the separatrices associated to the two last Euler configura- = 0, Z. 1 = ±O(ε1/3). These points, corresponding to tions located at ζ. 1 the equilibria point L1 en L2 in the RTBP, are associated with the Euler configurations for which the two planets are on the same side of the Sun. The last domain (in red), centred at the singularity of H 0, that is the collision between the two planets, is surrounded by the separatrix connecting the fixed point located at ζ. 1 = 0 and Z. 1 > 0 to itself. Inside this small region, the two planets seem to be subjected to a prograde satellite-like motion, the one revolving the other one clockwise. This last region (in red) ζ. 1 Figure 1: Phase portrait of the Hamiltonian H 0 in the coordinates (ζ. 1 , u). The units and the parameter are chosen such that Z. 2 = 0, G = 1, ω = 2π, εm1 = 10−3, εm2 = 3 × 10−4. See the text for more details. and its neighbourhood that includes L1 and L2 are located at a distance to the collision of order ε1/3 and consequently, is outside the validity domain of the resonant normal form (see Theorem 2 in the section 3.2) . Indeed, in this region the remainder H∗ is at least of the same order that the perturbation H P . Finally, let us remark that this figure is similar to the well known Hill's diagram (or zero-velocity curves) of the non averaged planar circular RTBP (see Szebehely, 1967) although the zero-velocity curves are not orbits of the system. It is also topologically equivalent to the phase space of the averaged planar circular RTBP when the eccentricity of the massless body is equal to zero (Nesvorn´y et al., 2002; Morbidelli, 2002). 16 5 An integrable approximation of the averaged Hamil- tonian 5.1 Expansion around the resonance In order to get a more tractable expression of the averaged Hamiltonian H 0 in the domaine K3 , it will be expanded in the neighborhood of Z. j = 0 so that the remainder is always smaller than HP . We will prove that the last condition is fulfilled when the expansion of the Keplerian part is truncated at the second order and H at zero order. 2 Let us start with the Keplerian part of the Hamiltonian. If its constant part 2−1(cid:16) β1µ2/3 1 + β2µ2/3 2 5.1.1 Approximation of the Keplerian part HK(Z. 1, Z. 2) = ωZ. 2 + Q(Z. 1, Z. 2) + R1(Z. 1, Z. 2) + R2(Z. 1, Z. 2) ω2/3 is omitted, HK can be written as: (cid:17)  R1(Z. 1, Z. 2) = (cid:101)Q(Z. 1, Z. 2) − Q(Z. 1, Z. 2) R2(Z. 1, Z. 2) = HK(Z. 1, Z. 2) − ωZ. 2 − (cid:101)Q(Z. 1, Z. 2) (39) (40) with where the quadratic form (cid:101)Q reads: (cid:101)Q(Z. 1, Z. 2) = − 3 ω4/3(β−1 1 µ +3ω4/3β−1 2 µ −2/3 1 −2/3 2 2 + β−1 2 µ Z. 1Z. 2 − 3 2 2 1 −2/3 )Z. 2 ω4/3β−1 2 µ −2/3 2 Z. 2 2, (41) and its approximation Q(Z. 1, Z. 2): −2/3 1 + m−1 (m−1 2 )Z. 0 −2/3 0 m−1 2 Z. 1Z. 2 − 3 2 Q(Z. 1, Z. 2) = − 3 2 +3ω4/3µ ω4/3µ 2 1 ω4/3µ −2/3 0 m−1 2 Z. 2 2, (42) We first look at the terms R1, as: βj = mj + O(ε) and µj = Gm0 + O(ε) = µ0 + O(ε), the quantity R1 = (cid:101)Q − Q satisfies the relation: R1 = O(ε(Z. 1, Z. 2)2) (43) 17 and more specifically, there exists a large enough constant M > 0 indepen- dent of ε, ρ, σ and δ such that R13 2 ≤ M ερ2. As regards the estimate of R2, the application of the Taylor formula on the function g(t) = HK(tZ. 1, tZ. 2) for (Z. 1, Z. 2) ∈ K3 R2(Z. 1, Z. 2) = HK(Z. 1, Z. 2) − ωZ. 2 − (cid:101)Q(Z. 1, Z. 2) = Using the inequality (11), we have for all t ∈ [0, 1]: g(3)(t) ≤ M(Z. 1, Z. 2)3, which finally leads to: (1 − t)2 (cid:90) 1 g(3)(t)dt leads to: 2 0 2 R2(Z. 1, Z. 2) = O3(Z. 1, Z. 2), (44) and more specifically to: R23 2 ≤ 9M 16 ρ3. 5.1.2 Approximation of the perturbation We consider the function G(ζ. 1 to the modified function H P (ζ. 1 equal to µ2/3 notation: j βjω−1/3 are replaced by µ2/3 , x. j, x. j) on the compact K3 which is equal , 0, 0, x. j, x. j) where the resonant actions Λ0 j 0 βjω−1/3. Equivalently, using the (cid:115) 2 ∆Z0 j = 6 µ2 j ω βj − 3 µ2 j ω βj = O(ε) for j ∈ {1, 2} (cid:115) 3 (cid:114) µ0 µj and using the transformation (9) we obtain: G(ζ. 1 , x. j, x. j) = H P (ζ. 1 , ∆Z0 1 , ∆Z0 1 + ∆Z0 2 , x. j, x. j). (45) Then the averaged perturbationH P can be split in the sum of three term as follows: H P (ζ. 1 , Z. 1, Z. 2, x. j, x. j) = G(ζ. 1 , x. j, x. j) + R3(ζ. 1 , x. j, x. j) + R4(ζ. 1 , Z. j, x. j, x. j) (46) with R3(ζ. 1 R4(ζ. 1 , x. j, x. j) = H P (ζ. 1 , Z. j, x. j, x. j) = H P (ζ. 1 , 0, 0, x. j, x. j) − G(ζ. 1 , Z. 1, Z. 2, x. j, x. j) −H P (ζ. 1 , x. j, x. j) (47) , 0, 0, x. j, x. j) 18 In order to estimate theses remainders, we need to consider the smaller where the neglected terms are smaller than the averaged ⊂ K3 compact K(κ) perturbation H P(κ) 3 2 2 ≤ HP3 2 . 3 2 With similar reasonings as in the previous section, we use the mean value theorem to evaluate the remainder in the truncation at order 0 of H P we consider K(κ) p ⊂ Kp which is defined for p > 0 and κ > 0 by: (cid:110) (cid:111) (ζj, Zj, ξj, ηj)j∈{1,2}∈ Kp such that max(Z1,Z2) ≤ κρ K(κ) p :== . The supremum norm .∞ on the space of holomorphic functions over the compact K(κ) p . For p = 3/2, we obtain on this smaller compact K(κ) p will be denoted by .(κ) for a small enough κ > 0: ⊂ K3 3 2 2 R3(κ) 3 2 = H P (ζ. 1 , 0, 0, x. j, x. j) −H P (ζ. 1 , ∆Z0 1 , ∆Z0 1 + ∆Z0 2 , x. j, x. j)(κ) 3 2 which implies that ≤ ∂ZH P3 R3(κ) 3 2 (∆Z0 2 1 , ∆Z0 1 + ∆Z0 We have an upper bound (∆Z0 1 , ∆Z0 C > 0 and our estimate on HP yields: 2 ) 2 ) ≤ 2 H P2(∆Z0 ρ 2 ) ≤ Cε for some constant 1 + ∆Z0 1 + ∆Z0 1 , ∆Z0 H P2 ≤ 2M ε δ =⇒ R3(κ) 3 2 ≤ M ε2 ρδ (48) for a large enough constant M > 0. In the same way, the mean value theorem yields: R4(κ) 3 2 ≤ ∂ZH P3 2 (Z. 1, Z. 2) ≤ 2 ρ H P2(Z. 1, Z. 2) =⇒ R4(ζ. 1 , Z. 1, Z. 2, x. j, x. j)(κ) 3 2 ≤ 4M ε ρδ κρ = 4M ε δ κ. (49) 19 5.1.3 The final approximation of H the constant term 2−1(cid:16) Gathering the approximation given in sections 5.1.1 and 5.1.2, and omitting ω2/3, the averaged Hamiltonian H β1µ2/3 1 + β2µ2/3 2 takes the following form: H(ζ. j Z. j, x. j, x. j) = ωZ. 2 + Q(Z. 1, Z. 2) + G(ζ. 1 , x. j, x. j) + R(ζ. j Z. j, x. j, x. j), (50) (cid:17) where the quadratic part Q is given by the expression (42), the perturbation G is defined in the top of Section 5.1.2, and the remainder R is defined by the sum R = R1 + R2 + R3 + R4 where the Rj are given in (40) and (47). If, as in the end of the section 3.2, we relate the analyticity width ρ to δ and ε by ρ2δ = ε, we get the following upper bound with M large enough: (cid:19) ≤ M ρ(cid:0)3ρ2 + 4κρ(cid:1) ρ2 + ε + 4κρ (51) (cid:18) 3 2 ≤ M ρ R(κ) 9 16 since ρ < 1 and δ < 1 =⇒ ε < ρ2. Especially, we have R(κ) ερ + H P(κ) 3 2 < 3M ρ3 + 4M κρ2 < mρ2 < H P3 ≤ 2 3 2 for κ small enough and ε small enough (since ρ < ε). 5.2 The dynamics on C0 and its implication for the initial problem Using the approximation (50) of the averaged Hamiltonian where the re- mainder R(ζ. j Z. j, x. j, x. j) is neglected, its restriction to the invariant manifold C0 reads: (cid:101)H0 = ωZ. 2 + Q(Z. 1, Z. 2) + εµ2/3 F (ζ. 1 0 ω2/3 m1m2 m0 1(cid:113) 2 − 2 cos ζ. 1 − ) (52) (53) with F (ζ. 1 ) = G(ζ. 1 , 0, 0) = cos ζ. 1 In order to uncouple the fast and semi-fast degrees of freedom, we define , Z. 1, Z. 2) = L(ϕ1, ϕ2, I1, I2) on the cylinder the symplectic linear map (ζ. 1 (cid:19) ]∆, 2π − ∆[×T × R × R by: (cid:19) (cid:18) (cid:19) (cid:19) , ζ. 2 m1 (cid:19)(cid:18)ϕ1 ϕ2 0 1 (cid:18)Z. 1 Z. 2 , (cid:18)1 0 = m1+m2 1 (54) (cid:19)(cid:18)I1 I2 = 1 − m1 m1+m2 (cid:18)ζ. 1 ζ. 2 20 and completed by the identity in the x. j and x. j variables. As a consequence, we have: H0(ϕj, Ij) = (cid:101)H0 ◦ L(ϕj, Ij) = H(1) (cid:18) 1 = − 3 2 ω4/3µ −2/3 0 (cid:19) 0 (ϕ1, I1) + H(2) 1 + εµ2/3 I 2 + 0 (ϕ2, I2) 0 ω2/3 m1m2 m0 F (ϕ1) (55) 1 m2 I 2 2 m1 + m2 m1 −2/3 0 + ωI2 − 3 2 ω4/3µ 0 0 0 3 2 The dynamics of the fast variables (ϕ2, I2) is now governed by H(2) the dynamics of the semi-fast variables (ϕ1, I1) is given by H(1) these two Hamiltonians are integrable. But the dynamics of H(2) while that of H(1) is less. 0 while 0 . Of course, is trivial m1+m2 (cid:113) 27 In the domain K(κ) , the phase portrait of the two approximations H(1) (already explored by several authors, i.e. Morais, 2001, and references therein) andH 0 of the averaged Hamiltonian are topologically equivalent. They both have two elliptic fixed points corresponding to L4 and L5 and an unstable equilibrium associated to the Euler configuration L3. The stable equilib- ria are located at (ϕ1, I1) = (±π/3, 0), and their eigenvalues, which are the √ same for both points, are equal to ±i ω+O(ε3/2). The unstable point is located at (ϕ1, I1) = (π, 0), its eigenvalues are ±√ ω + O(ε3/2). As we have seen in the section 4.2, in the domain enclosed by the separatrices emanating from this fixed point, one find tadpole orbits surrounding L4 or L5. Outside these invariant manifolds are the horseshoe orbits that encompass the three fixed points mentioned above. At this points, we know at least qualitatively what are the orbits on the invariant manifold C0. To go further, we would like to have the tem- poral parametrization of the corresponding trajectories. However, even if the Hamiltonian H(1) is integrable, its trajectories cannot be given explic- itly. Consequently, in the sequel, we will assume that these trajectories that satisfy the canonical differential equations: (cid:113) 21 m1+m2 m0 4 m0 ε ε 8 0 (cid:16)  I1 = εµ2/3 ϕ1 = −3 0 ω2/3 m1m2 m0 m1 + m2 ω4/3µ 1 − (2 − 2 cos ϕ1) −2/3 0 I1 m1m2 −3/2(cid:17) sin ϕ1 (56) are perfectly known, once given its initial conditions (ϕ1(0), I1(0)). Ac- 0 are closed tually, these solutions are all periodic since the level curves of H(1) and without singularities. 21 5.3 The normal stability of the manifold C0 Now, we study the linearized dynamic around the invariant manifold C0. Consider an arbitrary trajectory on C0. It is shown in Robutel and Pousse (2013) that the variational equation in the (x. j, x. j) direction around this solution reads: X = M (t)X where X = (cid:18)x1 (cid:19) x2 and M (t) = iεω m1m2 m0 (cid:16) with A(ζ. 1 ) = 5 cos 2ζ. 1 1 )5 4D(ζ. 1 −2iζ. 1 − (cid:113) 1 8D(ζ. 1 2 − 2 cos ζ. 1 , )5 B(ζ. 1 ) = e D(ζ. 1 ) = (57) (58)  (t)) m1m2 (t)) B(ζ. 1 √ A(ζ. 1 m2 (t)) m1 B(ζ. 1 √  A(ζ. 1 (cid:17) − cos ζ. 1 (t)) m1m2 , (cid:16) − 13 + 8 cos ζ. 1 −3iζ. 1 + 16e e −2iζ. 1 − 26e −iζ. 1 + 9eiζ. 1 (cid:17) , (59) (t) being a solution of the canonical equation associated to the Hamilto- ζ. 1 nian (52) According to the Floquet theorem (see Meyer and Hall, 1992), if the frequency of the considered periodic solution is ν, the solutions of the vari- ational equation take the form Y (t) = P (νt) exp(U t), (60) where U is a constant matrix and P (ψ) is a matrix whose coefficients are 2π-periodic functions of ψ. As, if Y is a fundamental matrix solution to the variational equation along a 2π/ν-periodic solution, one has the relation Y (t + 2πν−1) = Y (t) exp(cid:0)2πν−1U(cid:1) . (61) It turns out that the stability of the solutions of the variational equation (57) depends on the eigenvalues of the matrix U . As stated in section 4.2, the quantity x. 1x. 1 + x. 2x. 2 is an integral of the variational equation (57). This implies that the solutions of (57) are bounded, and as a consequence, U is 22 diagonalisable and the real parts of its eigenvalues are equal to zero. Al- though we cannot exclude that one of the eigenvalue vanishes, the invariant manifold C0 is normally stable. There exists, on C0, at least three trajectories for which one eigenvalue of the variational equation (57) vanishes. These are the stationary solutions of the differential system (56), that is the Euler configuration L3 and the Lagrange ones L4 and L5. For the collinear configuration associated to L3, which corresponds to (ϕ1, I1) = (π, 0), we have As a consequence, the equilibrium has two eigendirections collinear to A(π) = 7 8 , B(π) = 7 8 . (cid:19) (cid:18)√ m2√ m1 V 1 π = and V 2 π = (cid:19) (cid:18) √ m1−√ m2 (62) (63) (64) associated respectively to the eigenvalue v1 π = iεω 7 8 m1 + m2 m0 and v2 π = 0. The reason why one of the eigenvalues vanishes has been given in section 4.1. Indeed it is easy to verify that the eigenvector V 2 π corresponds to the elliptic Euler's configurations where the two planet are in the two sides of the Sun. The other eigendirection corresponds to a non trivial family of periodic orbits. According to the expressions (63), the configurations corresponding to V 1 π are two ellipses in conjonction (1 = 2) with equal semi-major axis and whose eccentricities satisfy the relation m1e1 = m2e2. Contrary to the previous Euler's configurations, the ellipses are not fixed, π = O(ε). This but precess at the same rate defined by the frequency −iv1 eigendirection gives rise to a one-parameter family unstable periodic orbits of the averaged problem (periodic in rotating frame in the non-averaged prob- lem) described by Hadjidemetriou et al. (2009) and related to the Poincar´e solutions of second sort (see Robutel and Pousse, 2013, for more details). Similar phenomena occur in the neighborhood of the two equilateral fixed points L4 and L5 at ϕ1 = ±π/3 and I1 = 0. Without entering into details (see Robutel and Pousse, 2013), let us just mention what we get for L4 (the results are similar for the other equilateral equilibrium). The coefficients of the matrix (58) satisfy √ (1 − i 3). (65) A( π 3 ) = − 27 8 , B( π 3 ) = 27 16 23 As a consequence, its eigenvectors are V 1 π/3 = and V 2 π/3 = (cid:19) (cid:18)√ m2eiπ/3 −√ m1 (cid:18)√ (cid:19) m1eiπ/3 √ m2 and the corresponding eigenvalues are π/3 = −iεω v1 27 8 m1 + m2 m0 and v2 π/3 = 0. (66) (67) It is easy to verify that the configurations associated to V 2 π/3 are the elliptic equilateral configurations that are fixed points of the averaged Hamiltonian. This is the reason why v2 π/3 = 0. As in the case of the Euler configuration, the eigenvector V 1 π/3 is tangent to a one-parameter family of periodic orbits, called anti-Lagrange by Giuppone et al. (2010). For small eccentricities, the elliptic elements of the corresponding orbits that precess simultaneously at the frequency iv1 λ1 − λ2 + π. π/3 satisfy the relation m1e1 = m2e2 and 1 − 2 = 6 Consequences for the co-orbital motion So far, the study of the manifold C0 and its dynamics was carried out in the context of the averaged problem. To end this section, we will study what happens to the dynamics of C0 when we go back to the initial variables. The following theorem gives a partial approximation, on a finite but large time, of the dynamics on the manifold C0 in the initial variables (λj, Λj, xj,−ixj). Actually, we explain at the end of this section the way to obtain a complete result of approximation with our methods. We consider a solution in the initial variables (λj(t), Λj(t), xj(t),−ixj(t)) starting with at t = 0 in the image of the manifold C0 by the transformation (λj, Λj, xj,−ixj) = Ψ ◦ C ◦ L(ϕj, Ij, x. j,−ix. j) considered in section 2, 3 and 5, hence: (λj(0), Λj(0), xj(0),−ixj(0)) = Ψ ◦ C ◦ L(ϕj(0), Ij(0), 0, 0). 24 Using these linear and averaging transformations, we can formally write: λ1(t) =(cid:101)ω(I2(0))t + λ2(t) =(cid:101)ω(I2(0))t − m1 m2 m1 + m2 Λ1(t) = Λ0 Λ2(t) = Λ0 1 + I1(t) + 2 − I1(t) + m1 + m2 m2 m1 + m2 m1 m1 + m2 ϕ1(t) + ϕ2(0) + ρ1(t) ϕ1(t) + ϕ2(0) + ρ2(t) I2(0) + τ1(t) I2(0) + τ2(t) (68) xj(t) = τj+2(t) where the function (ϕ1(t), I1(t)) in these expressions is the solution of the differential system (56) with the initial conditions given as follow: ϕ1(0) = λ1(0) − λ2(0) I1(0) = m2 m1 + m2 m1 ϕ2(0) = m1 + m2 I2(0) = Λ1(0) − Λ0 (cid:101)ω(I2) = ∂I2H(2) m1 + m2 (Λ1(0) − Λ0 1) − m1 m2 λ1(0) + m1 + m2 1 + Λ2(0) − Λ0 2 λ2(0) (Λ2(0) − Λ0 2) (69) 0 (ϕ2, I2) = ω − 3ω4/3µ −2/3 0 I2 m1 + m2 and the remainders ρ1(t), ρ2(t), τ1(t), τ2(t), τj+2(t) should be small over large times. We give a partial theorem in this direction. We first recall bounds on the remainders in the previous computations with the choices δρ2 = ε and 32Cσ = δ < 16Cσ0, there exists a large enough constant M independent of ε, ρ, σ and δ: (cid:114) and we denote (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H∗ +H ∗ 0 + R (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(κ) 3 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3 2 ∗ 0 H∗3/2 ≤ M = εO(ε) = O(ε2) ≤ M ε2 ε3 δ5 ≤ M ρ(cid:0)3ρ2 + 4κρ(cid:1) (cid:34)(cid:114) R(κ) 3 2 δ5 + ε2 + ρ(cid:0)3ρ2 + 4κρ(cid:1)(cid:35) ε3 (70) (71) (72) . (73) < L := M 25 Theorem 3 We denote: p = m1 + m2 2m1 + m2 for the solution ΦH(cid:48)◦L eraged Hamiltonian H(cid:48). t and Tp the first time of escape out of K(κ) p (ϕj(0), Ij(0), 0, 0) along the flow governed by the av- With the previous assumptions and notations, we have the following bounds on the remainders: τj+2(t) ≤ ρ for t ≤ min (cid:18) κρ2 4L (cid:19) ,Tp and Λ1(t) + Λ2(t) = Λ1(0) + Λ2(0) + τ (t) with τ (t) ≤ ρ for t ≤ min (cid:18) κρ2 4L (cid:19) . ,Tp (74) (75) The proof of this theorem is given in the section 7.3 and exactly the same reasonings would give a complete approximation of the solutions on C0 except that we need moreover the action-angles variables linked to the integrable Hamiltonian H(1) 0 which can be built by classical techniques (cf Arnold) The first step is to bound the difference between the flows linked to two nearby Hamiltonian in the normalized variables (ϕj, Ij, x. j,−ix. j) ∈ K(κ) p and then gives a sharp timescale such that these two flows remain close in the initial variables (λj, Λj, xj,−ixj). 7 Proof of the theorems 7.1 Proof of theorem 1 Using the notations of section 2.4 and the mean value theorem, we obtain a lower bound for the minimal distance (with C2 equipped with the Hermitian norm) between two orbits in the complex domain K(C) . ∆,ρ0,σ0 Actually, denoting , r(C) (r(C) , r(C) , r(C) 1 1 2 2 ) = Υ (ςj, zj, ξj, ηj)j∈{1,2} with Υ = (Φ1 ◦ Ψ, Φ2 ◦ Ψ), 26 then the triangular inequality yields: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) 1 − r(C) 2 ρσ + σ) (ςj, zj, ξj, ηj) − (Re(ςj), 0, 0, 0) ≤ ρ + 2 j − r(R) j =⇒(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = ≥ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(R) ≥ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(R) √ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C(ρ + 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) − 2C(ρ + 2 1 − r(R) 1 + r(R) 2 1 − r(R) 1 − r(R) 1 − r(R) 2 1 − r(R) √ 1 2 + r(R) If we assume that ρ < σ and σ ≤ δ 16C = a 8C sin √ ρσ + σ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 − r(R) 2 2 ρσ + σ) 2 − r(C) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) (cid:19) (cid:18) ∆ (cid:18) ∆ (cid:19) 2 , (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (76) (77) we have: (r(R) 1 , r(R) 1 , r(R) 2 , r(R) 2 ) = Ψ ((Re(ς1), 0, 0, 0); (Re(ς2), 0, 0, 0)) , and the mean value theorem together with our bound C on the differential of Υ allows to write for j ∈ {1, 2}: we obtain with the lower bound on the mutual distance in the real domain: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) 1 − r(C) 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ − 8Cσ ≥ δ 2 = a sin 2 and the expression of the planetary Hamiltonian gives the upper bounds on the size of HP in the complex domain. In the same way, under the assumption ρ < σ < 1, we obtain with the upper bound on the mutual distance in the real domain: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)r(C) 1 − r(C) 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ M + 8Cσ < M + 8C (78) and the expression of the planetary Hamiltonian gives the lower bounds on the size of HP in the complex domain provided that M is large enough to satisfy < M + 8C. ε M 27 7.2 Proof of theorem 2 Let ∆ > 0, ρ > 0 and σ > 0 such that (∆, 2ρ, 2σ) satisfy (12), hence: ρ < σ ≤ δ 32C = a 16C sin with δ := 2a sin (79) (cid:18) ∆ (cid:19) 2 (cid:18) ∆ (cid:19) 2 ε δ then there exists a constant M > 0 independant of ε, ∆, ρ and σ such that HK2ρ < M and HP2 = HP∆,2ρ,2σ < M (80) .2 = .∆,2ρ,2σ) on the space of for the supremum norm .2ρ (resp. holomorphic functions over the compact {(Z1, Z2) ∈ C2 such that max(Z1,Z2) ≤ 2ρ} (resp. over the compact K(C) ∆,2ρ,2σ = K2). In view of our bounds (15) and (80), we obtain: χ2 < 2π ω (HP2) < M 2πε ωδ . (81) Moreover, with our rescaling of the masses, the derivatives of the Ke- plerian part HK remains of order one and we can also assume that the constant M > 0 independant of ε, ∆, ρ and σ gives also an upper bound on the Hessian of HK : ∀(Z1, Z2) ∈ C2 with max(Z1,Z2) ≤ 2ρ : (Z1, Z2) (cid:12)(cid:12)(cid:12)(cid:12) ∂2HK ∂Zi∂Zj (cid:12)(cid:12)(cid:12)(cid:12) < M. (82) We must first estimate the size of the partial derivatives and the Poisson bracket of analytical functions on K2 by classical applications of Cauchy inequalities which can be found in Poschel (1993) and Giorgilli (2003). Let f be analytical on K2 (continuous on the boundary), we can write : (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂f (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = Sup(e1,e2)=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) d (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (ζ, Z, x,−ix) f (ζ + te, Z, x,−ix) ∂ζ dtt=0 for (ζ, Z, x,−ix) = (ζj, Zj, xj,−ixj)j∈{1,2} ∈ K2. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 28 One then applies Cauchy formula to the function t (cid:55)→ f (ζ + te, Z, x,−ix) of the complex variable t, holomorphic and continuous on the boundary for t ≤ σ/4 when (ζ, Z, x,−ix) ∈ K(C) 7/4, and obtains : The same reasoning yields the equivalent inequalities for the other partial derivatives : (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂f ∂Z (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 f2 ; ≤ 4 ρ f2 . ≤ 4 σ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂f ∂x (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 and (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂ζ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂f (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂f (cid:18) (cid:20) ∂x ≤ 4√ ρσ f2 . (cid:19)(cid:21) To estimate the size of the Poisson brackets for two analytical functions on K2, we write in a similar way that ζ−t {f, g}(ζ, Z, x,−ix) = g d dtt=0 ∂f ∂Z , Z +t ∂f ∂ζ , x−it ∂f ∂x ,−ix+t ∂f ∂x and the function of the complex variable t in the right hand side is defined for t ≤ ρσ 16f2 (we apply the Cauchy formula to f ), hence we can write: {f, g}7/4 ≤ 16 {f, g}3/2 ≤ 4 f2g2 ρσ f2g2 . ρσ , in the same way we obtain Since the averaging transformation ϕ is the time-one map Φχ 1 of the Hamiltonian flow generated by some auxiliary hamiltonian χ, for any func- tion K defined on the phase space: (K ◦ Φχ t ) = {K, χ} ◦ Φχ d dt (cid:90) 1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3/2 0 t =⇒ ζ. j =⇒(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ζ. j − ζj = − ζj ◦ Φχ ∂χ ∂Zj ≤ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ ∂Zj t dt (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 for j ∈ {1, 2}. Actually, we prove that starting inside K3/2 along the flow 1 yields a time of escape out of K7/4 which is bigger than 1 and we can Φχ use our estimates on χ. 29 with our threshold (17). − ζj In the same way, we obtain: Hence , we have:(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ζ. j (cid:12)(cid:12)(cid:12)(cid:12)Z. j − Zj (cid:12)(cid:12)(cid:12)(cid:12)3/2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ (cid:12)(cid:12)(cid:12)(cid:12)3/2 (cid:12)(cid:12)(cid:12)(cid:12)x. j − xj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ (cid:12)(cid:12)(cid:12)(cid:12)x. j − xj (cid:12)(cid:12)(cid:12)(cid:12)3/2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 ∂xj ≤ ≤ ∂ζj ≤ ⊆ C(K3 2 ∂xj ) ⊆ K7 . 4 which yields K5 4 ≤ 4 ρ χ2 < 8M π ε ω ρδ < σ 4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 χ2 < ≤ 4 σ 8M π ε ω σδ < ρ 4 ≤ 4√ ρσ χ2 < 8M π ω ≤ 4√ ρσ χ2 < 8M π ω ε√ ρσδ ε√ ρσδ < < √ ρσ 4 √ ρσ 4 As H(cid:48)(ζ. j , Z. j, x. j,−ix. j) = HK(Z. 1, Z. 2) +H P (ζ. 1 , Z. j, x. j,−ix. j) + H∗(ζ. j , Z. j, x. j,−ix. j), where H∗ = (∇HK−−→ω ) . ∂χ ∂ζ. + {χ, HP} + H(cid:48) − H − {χ, H} for −→ω := (cid:18) 0 ω (cid:19) , (cid:90) 1 we may use Taylor's formula at order two to write: H(cid:48)−H−{χ, H} = H◦Φχ 1−H◦Φχ 0− d dt (H◦Φχ t dt. 1 (K3/2) ⊂ K7/4, using the upper bound M on the Hessian of HK 0 ) = 0 (1−t){χ,{χ, H}}◦Φχ Since Φχ and Cauchy inequalities, one finds the following estimate : ≤ 7πM 2 ω χ2 σ/4 ≤ M ρ 7 4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(∇HK−−→ω ) . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ ∂ζ. ρε σδ moreover : {χ, HP}7/4 < 16 ≤ 16πM 2 ω ε2 ρσδ2 . χ2HP2 ρσ 30 (83) (84) To estimate the third term in H∗, we insert again the definition of χ given in section (3.1) into the Poisson bracket to get : {χ, H} = H P − HP + (∇HK − −→ω ) . + {χ, HP} ∂χ ∂ζ. hence (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:40) {χ,{χ, H}}3/2 ≤ {χ, HP −H P}}3/2 χ, (∇HK − −→ω ) . ∂χ ∂ζ. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + (cid:41)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3/2 + {χ,{χ, HP}}3/2 Using again Cauchy inequalities, we can estimate the previous terms and the sum of these inequalities yields the given value of η, indeed: ε2 ρσδ2 . χ2HP −H P2 ≤ 16πM 2 ρσ ω {χ, HP −H P}}3/2 < 4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:40) χ, (∇HK − −→ω ) . (cid:41)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ ∂ζ. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 (cid:33)2 ∂χ ∂ζ. ω ρε σδ ρσ < 8 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 4πM (cid:32) χ2 ≤ 7M (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3/2 χ2{χ, HP}3/4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(∇HK − −→ω ) . (cid:33)2 (cid:32) (cid:41)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3/2 =⇒ {χ,{χ, HP}}3/2 < (cid:19) ∂χ ∂ζ. ≤ M 8πM 2 (cid:18) 21 gives: < ω 7 2 ω ε ω ε ρ σ 2 + 40 ρσδ M ε δ πM 2 ρσδ 16πM ε ρσδ ε δ ρ ε σ δ ε δ (cid:18) ρ σ (85) (86) (cid:19) . + ε ρσδ moreover the threshold ρσ {χ,{χ, HP}}3/2 < 8 (cid:40) χ, (∇HK − −→ω ) . ρ =⇒ 32πM δρσ < < ω ω ε ε δσ 32πM (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Finally, the sum of all the previous estimates yields: ε δρσ < ω 32πM H∗3/2 ≤ M π ω and ensures the claimed value for the upper bound η = 40 M π ω 31 7.3 Proof of the theorem 3 The first part of the proof of theorem 3 comes from the use of Cauchy inequalities to bound the size of the Hamiltonian vector field linked to (H∗ + ∗ 0 + R) ◦ L over the compact K(κ) H p . The second part comes from the size of Ψ◦C◦L composed of the averaging transformation C and two linear transformations Ψ and L and the choice of a time T which give terms of the same order in the upper bound on the error in the approximate solution. formation L admits the upper bound p−1 = 2m1+m2 More specifically, by straightforward computations the norm of the trans- hence L(cid:16)K(κ) (cid:17) ⊂ K(κ) p 1 . m1+m2 With our notations, the function (ϕ1(t), ϕ2(t), I1(t), I2(t), 0, 0) = (ϕ1(t), ϕ2(0)+∂I2H(2) 0 (I2(0))t, I1(t), I2(0), 0, 0) (cid:19) (cid:18) d(cid:101)I2 dt (ϕ1(t), ϕ2(0) + ω − 3ω4/3µ −2/3 0 I2(0) m1 + m2 t, I1(t), I2(0), 0, 0) (87) is a solution of the Hamiltonian system linked to H(2) solution of the averaged system governed by H. on C0, hence it is a Then, the complete averaged Hamiltonian H(cid:48) satisfies H(cid:48) −H = H∗ + ∗ H the system linked to H(cid:48) starting at (ϕ1(0), ϕ2(0), I1(0), I2(0), 0, 0), we can write 0 + R and if we denote ((cid:101)ϕ1(t),(cid:101)ϕ2(t),(cid:101)I1(t),(cid:101)I2(t), x. (t), x. (t)) the solution of 0 = −∂ϕ2(H∗ +H ∗ 0 + R) ◦ L. and Cauchy inequality yields: Then, with our assumption on the time of escape Tp, we have: for t < Tp 1 L((cid:101)ϕ1(t),(cid:101)ϕ2(t),(cid:101)I1(t),(cid:101)I2(t), x. (t), x. (t)) ∈ K(κ) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(κ) =⇒(cid:101)I2(t) =(cid:101)I2(0) + τ2(t) with τ2(t) ≤ 2L (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)H∗ +H (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) d(cid:101)I2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∗ 0 + R 2 κρ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt < 3 2 t for t ≤ Tp. κρ 2L κρ . (88) We have the same bound in the averaged variable Z. 2 = I2. 32 Finally, in the initial variables Z2, we have: (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∂χ ∂ζ2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)7/4 Z. 2 − Z23/2 ≤ < 8M π ε ω σδ < ρ 4 et Z2(t) − Z2(0) ≤ Z2(t) − Z. 2(t) +Z. 2(t) − Z. 2(0) +Z. 2(0) − Z2(0) =⇒ Z2(t) − Z2(0) ≤ 2Z. 2 − Z23/2 + (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:101)I2(t) −(cid:101)I2(0) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19) (cid:18) κρ2 ,Tp 4L =⇒ Z2(t) − Z2(0) ≤ ρ 2 + 2L κρ t < ρ for t ≤ min (89) which gives the formula for Λ in the theorem. The formula for (xj, xj) is proved in the same way. References Arnold, V. I. (1963). Small denominators and problems of stability of motion in classical and celestial mechanics. Russ. Math. Survey, 18(6):85 -- 192. Chierchia, L. and Pinzari, G. (2011). The planetary n-body problem: sym- plectic foliation, reductions and invariant tori. Invent math, 186(1-77). Delaunay (1860). Th´eorie du mouvement de la Lune. M´emoires de l'acad´emie des sciences de l'institut imp´erial de France, XXVIII. ´Erdi, B. (1977). An asymptotic solution for the trojan case of the plane elliptic restricted problem of three bodies. Celestial Mechanics, 15:367 -- 383. F´ejoz, J. (2004). D´emonstration du th´eor`eme d'Arnold sur la stabilit´e du syst`eme plan´etaire (d'apr`es Michael Herman). Ergod. Th. Dyn. Sys., 24:1 -- 62. Giorgilli, A. (2003). Notes on exponential stability of hamiltonian systems. In Dynamical Systems. Part I. Hamiltonian Systems and Celestial Me- chanics, pages 87 -- 198. Pubblicazioni del Centro di Ricerca Matematica Ennio de Giorgi, Proceedings. Scuola Normale Superiore, Pisa. Giuppone, C. A., Beaug´e, C., Michtchenko, T. A., and Ferraz-Mello, S. (2010). Dynamics of two planets in co-orbital motion. MNRAS, 407:390 -- 398. 33 Hadjidemetriou, J. D., Psychoyos, D., and Voyatzis, G. (2009). The 1/1 resonance in extrasolar planetary systems. Celest. Mech. Dyn. Astron., 104:23 -- 38. Haghighipour, N. (2002). Resonance dynamics and partial averaging in a re- stricted three-body system. Journal of Mathematical Physics, 43(7):3678 -- 3694. Lagrange (1778). Recherches sur les ´equations s´eculaires des mouvements des noeuds et des inclinaisons des plan`etes. M´emoires de l'Acad´emie des Sciences de Paris, ann´ee 1774. Laskar, J. and Robutel, P. (1995). Stability of the planetary three-body problem I: Expansion of the planetary hamiltonian. Celestial Mechanics and Dynamical Astronomy, 62:193 -- 217. Meyer, K. R. and Hall, G. R. (1992). Introduction to Hamiltonian dynamical systems and the n-body problem. Springer-Verlag. Morais, M. H. M. (2001). Hamiltonian formulation of the secular theory for Trojan-type motion. Astron. Astrophys., 369:677 -- 689. Morbidelli, A. (2002). Modern celestial mechanics : aspects of solar system dynamics. Taylor & Francis, London, 2002, ISBN 0415279399. Namouni, F. (1999). Secular Interactions of Coorbiting Objects. Icarus, 137:293 -- 314. Nesvorn´y, D., Thomas, F., Ferraz-Mello, S., and Morbidelli, A. (2002). A perturbative treatment of the co-orbital motion. Celest. Mech. Dyn. As- tron., 82(4):323 -- 361. Poincar´e, H. (1892). M´ethodes nouvelles de la M´ecanique C´eleste, volume I. Gauthier Villars Paris, reprinted by Blanchard, 1987. Poschel, J. (1993). Nekhoroshev estimates for quasi-convex hamiltonian systems. Math. Z., 213:187 -- 216. Robutel, P. and Pousse, A. (2013). On the co-orbital motion of two planets in quasi-circular orbits. Celest. Mech. Dyn. Astron., 117:17 -- 40. Szebehely, V. (1967). Theory of orbits: the restricted problem of three bodies. Academic Press, New-York. 34
1209.0608
1
1209
2012-09-04T11:16:12
A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System
[ "astro-ph.EP", "astro-ph.SR" ]
In 2009, the discovery of two planets orbiting the evolved binary star system HW Virginis was announced, based on systematic variations in the timing of eclipses between the two stars. The planets invoked in that work were significantly more massive than Jupiter, and moved on orbits that were mutually crossing - an architecture which suggests that mutual encounters and strong gravitational interactions are almost guaranteed. In this work, we perform a highly detailed analysis of the proposed HW Vir planetary system. First, we consider the dynamical stability of the system as proposed in the discovery work. Through a mapping process involving 91,125 individual simulations, we find that the system is so unstable that the planets proposed simply cannot exist, due to mean lifetimes of less than a thousand years across the whole parameter space. We then present a detailed re-analysis of the observational data on HW Vir, deriving a new orbital solution that provides a very good fit to the observational data. Our new analysis yields a system with planets more widely spaced, and of lower mass, than that proposed in the discovery work, and yields a significantly greater (and more realistic) estimate of the uncertainty in the orbit of the outermost body. Despite this, a detailed dynamical analysis of this new solution similarly reveals that it also requires the planets to move on orbits that are simply not dynamically feasible. Our results imply that some mechanism other than the influence of planetary companions must be the principal cause of the observed eclipse timing variations for HW Vir. If the sys- tem does host exoplanets, they must move on orbits differing greatly from those previously proposed. Our results illustrate the critical importance of performing dynamical analyses as a part of the discovery process for multiple-planet exoplanetary systems.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1–?? (2012) Printed 2 May 2014 (MN LATEX style file v2.2) A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System J. Horner1(cid:63), T. C. Hinse2,3, R. A. Wittenmyer1, J. P. Marshall4 & C. G. Tinney1 1Department of Astrophysics and Optics, School of Physics, University of New South Wales, Sydney 2052, Australia 2Korea Astronomy and Space Science Institute, 776 Daedeokdae-ro, Yuseong-gu, 305-348, Daejeon, Republic of Korea (South) 3Armagh Observatory, College Hill, BT61 9DG, NI, UK 4Departmento F´ısica Te´orica, Facultad de Ciencias, Universidad Aut´onoma de Madrid, Cantoblanco, 28049, Madrid, Espana Accepted for publication in Monthly Notices of the Royal Astronomical Society, 3rd September 2012 ABSTRACT In 2009, the discovery of two planets orbiting the evolved binary star system HW Virginis was announced, based on systematic variations in the timing of eclipses between the two stars. The planets invoked in that work were significantly more massive than Jupiter, and moved on orbits that were mutually crossing - an architecture which suggests that mutual encounters and strong gravitational interactions are almost guaranteed. In this work, we perform a highly detailed analysis of the proposed HW Vir planetary system. First, we consider the dynamical stability of the system as proposed in the discovery work. Through a mapping process involving 91,125 individual simulations, we find that the system is so unstable that the planets proposed simply cannot exist, due to mean lifetimes of less than a thousand years across the whole parameter space. We then present a detailed re-analysis of the observational data on HW Vir, deriving a new orbital solution that provides a very good fit to the observational data. Our new analysis yields a system with planets more widely spaced, and of lower mass, than that proposed in the discovery work, and yields a significantly greater (and more realistic) estimate of the uncertainty in the orbit of the outermost body. Despite this, a detailed dynamical analysis of this new solution similarly reveals that it also requires the planets to move on orbits that are simply not dynamically feasible. Our results imply that some mechanism other than the influence of planetary companions must be the principal cause of the observed eclipse timing variations for HW Vir. If the sys- tem does host exoplanets, they must move on orbits differing greatly from those previously proposed. Our results illustrate the critical importance of performing dynamical analyses as a part of the discovery process for multiple-planet exoplanetary systems. Key words: binaries: close, binaries: eclipsing, stars: individual: HW Vir, planetary systems, methods: N-body simulations 2 1 0 2 p e S 4 . ] P E h p - o r t s a [ 1 v 8 0 6 0 . 9 0 2 1 : v i X r a 1 INTRODUCTION Since the discovery of the first planets around other stars (Wol- szczan & Frail 1992; Mayor & Queloz 1995), the search for exo- planets has blossomed to become one of the most exciting fields of modern astronomical research. The great majority of the hundreds of exoplanets that have been discovered over the past two decades have been found orbiting Sun-like stars by dedicated international radial velocity programs. Among these programmes are HARPS (the High Accuracy Radial velocity Planetary Search project, e.g. Pepe et al. 2004; Udry et al. 2007; Mayor et al. 2009), AAPS (the (cid:63) E-mail: [email protected] (JH) c(cid:13) 2012 RAS Anglo-Australian Planet Search, e.g. Tinney et al. 2001, 2011; Wit- tenmyer et al. 2012b), California (e.g. Howard et al. 2010; Wright et al. 2011), Lick-Carnegie (e.g. Rivera et al. 2010; Anglada- Escud´e et al. 2012), and Texas (e.g. Endl et al. 2006; Robertson et al. 2012a). The other main method for exoplanet detection is the transit technique, which searches for the small dips in the bright- ness of stars that result from the transit of planets across them. Ground-based surveys such as WASP (Hellier et al. 2011; Smith et al. 2012) and HAT (Bakos et al. 2007; Howard et al. 2012) have pi- oneered such observations, and resulted in the discovery of a num- ber of interesting planetary systems. In the coming years, such sur- veys using space-based observatories will revolutionise the search for exoplanets. Indeed, a rapidly growing contribution to the cata- 2 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney logue of known exoplanets comes from the Kepler spacecraft (e.g. Borucki et al. 2011; Welsh et al. 2012; Doyle et al. 2011), which will likely result in the number of known exoplanets growing by an order of magnitude in the coming years. In recent years, a number of new exoplanet discoveries have been announced featuring host stars that differ greatly from the Sun-like archetype that make up the bulk of detections. The most striking of these are the circumbinary planets, detected around eclipsing binary stars via the periodic variations in the timing of observed stellar eclipses. A number of these unusual systems fea- ture cataclysmic variable stars, interacting binary stars composed of a white dwarf primary and a Roche lobe filling M star secondary. Sharply-defined eclipses of the bright accretion spot, with periods of hours, can be timed with a precision of a few seconds. In these systems, the eclipse timings are fitted with a linear ephemeris, and the residuals (O−C) are found to display further, higher-order vari- ations. These variations can be attributed to the gravitational effects of distant orbiting bodies which tug on the eclipsing binary stars, causing the eclipses to appear slightly early or late. This light-travel time (LTT) effect can then be measured and used to infer the pres- ence of planetary-mass companions around these highly unusual stars. Some examples of circumbinary companions discovered in this manner include UZ For (Potter et al. 2011), NN Ser (Beuer- mann et al. 2010), DP Leo (Qian et al. 2010), HU Aqr (Schwarz et al. 2009; Qian et al. 2011) and SZ Her (Hinse et al. 2012b; Lee et al. 2012) The first circumbinary planets to be detected around hosts other than pulsars were those in the HW Vir system (Lee et al. 2009), which features a subdwarf primary, of spectral class B, and a red dwarf companion which display mutual eclipses with a pe- riod of around 2.8 hours (Menzies & Marang 1986). The detection of planets in this system was based on the timing of mutual eclipses between the central stars varying in a fashion that was best fit by including two sinusoidal timing variations. The first, attributed to a companion of mass M sin i=19.2 MJup, had a period of 15.8 years, and a semi-amplitude of 77 s, while the second, attributed to a com- panion of mass M sin i = 8.5 MJup, had a period of 9.1 years and semi-amplitude of 23 s. Whilst these semi-amplitudes might ap- pear relatively small, the precision with which the timing of mutual eclipses between the components of the HW Vir binary can be mea- sured means that such variations are relatively easy to detect. Over the last decade, a number of studies have shown that, for systems that are found to contain more than one planetary body, a detailed dynamical study is an important component of the planet discovery process that should not be overlooked (e.g. Stepinski et al. 2000; Go´zdziewski & Maciejewski 2001; Ferraz-Mello et al. 2005; Laskar & Correia 2009). However, despite these pioneering works, the great majority of exoplanet discovery papers still fail to take account of the dynamical behaviour of the proposed systems. Fortunately, this situation is slowly changing, and recent discovery papers such as Robertson et al. (2012a), Wittenmyer et al. (2012b) and Robertson et al. (2012b) have shown how studying the dynam- ical interaction between the proposed planets can provide signifi- cant additional constraints on the plausible orbits allowed for those planets. Such studies can even reveal systems in which the observed signal cannot be explained by the presence of planetary compan- ions. The planets proposed to orbit HU Aqr are one such case, with a number of studies (e.g. Horner et al. 2011; Funk et al. 2011; Hinse et al. 2012a; Wittenmyer et al. 2012a; Go´zdziewski et al. 2012) showing that the orbital architectures allowed by the observa- tions are dynamically unstable on astronomically short timescales. In other words, whilst it is clear that the observed signal is truly Parameter HW Vir b M sin i a sin i e ω T P 0.00809 ± 0.00040 3.62 ± 0.52 0.31 ± 0.15 60.6 ± 7.1 2,449,840 ± 63 3316 ± 80 HW Vir c Unit 0.01836 ± 0.000031 M(cid:12) AU 5.30 ± 0.23 0.46 ± 0.05 90.8 ± 2.8 2,454,500 ± 39 5786 ± 51 deg HJD days Table 1. Parameters for the two planetary bodies proposed in the HW Vir system, taken from Lee et al. (2009) (their Table 7). there, it seems highly unlikely that it is solely the result of orbiting planets. Given these recent studies, it is clearly important to consider whether known multiple-planet exoplanetary systems are truly what they seem to be. In this work, we present a re-analysis of the 2-planet system proposed around the eclipsing binary HW Vir. In section 2, we briefly review the HW Vir planetary system as proposed by Lee et al. (2009). In section 3, a detailed dynamical analysis of the planets proposed in that work is performed. In sec- tion 4, we present a re-analysis of the observations of HW Vir that led to the announcements of the exoplanets, obtaining a new orbital solution for those planets which is dynamically tested in section 5. Finally, in section 6, we present a discussion of our work, and draw conclusions based on the results herein. 2 THE HW VIR PLANETARY SYSTEM The HW Vir system consists of a subdwarf B primary and an M6- 7 main sequence secondary. The system eclipses with a period of 2.8 hours (Menzies & Marang 1986), and the stars have masses of M1 = 0.48 M(cid:12) and M2 = 0.14 M(cid:12) (Wood & Saffer 1999). Changes in the orbital period of the eclipsing binary were first noted by Kilkenny et al. (1994); further observations led other authors to suggest that the period changes were due to LTT effects aris- ing from an orbiting substellar companion (Kilkenny et al. 2003; Ibanoglu et al. 2004). Lee et al. (2009) obtained a further 8 years of photometric observations of HW Vir. Those data, in combination with the previously published eclipse timings spanning 24 years, indicated that the period changes consisted of a quadratic trend plus two sinusoidal variations with periods of 15.8 and 9.1 years. Lee et al. (2009) examined alternative explanations for the cyclical changes, ruling out apsidal motion and magnetic period modula- tion via the Applegate mechanism (Applegate 1992). They con- cluded that the most plausible cause of the observed cyclic pe- riod changes is the light-travel time effect induced by two com- panions with masses 19.2 and 8.5 MJup. The parameters of their fit can be found in Table 1. Formally, the planets are referred to as HW Vir (AB)b and HW Vir (AB)c, but for clarity, we refer to the planets as HW Vir b and HW Vir c. A first look at the fitted parameters for the two proposed planets reveals an alarming result: the planets are both massive, in the regime that borders gas giants and brown dwarfs, and oc- cupy highly eccentric, mutually crossing orbits with separations that guarantee close encounters - generally a surefire recipe for dynamical instability (as seen for the proposed planetary system around HU Aqr (e.g. Horner et al. 2011; Hinse et al. 2012a; Wit- tenmyer et al. 2012a; Go´zdziewski et al. 2012). 3 A DYNAMICAL SEARCH FOR STABLE ORBITS The work by Lee et al. (2009) derived relatively high masses (8.5 and 19.2 MJup) and orbital eccentricities (0.31 and 0.46), and so c(cid:13) 2012 RAS, MNRAS 000, 1–?? A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 3 significant mutual gravitational interactions are expected. To assess the dynamical stability of the proposed planets in the HW Vir sys- tem, we performed a large number of simulations of the planetary system, following a successful strategy used on a number of pre- vious studies (e.g. Marshall et al. 2010; Horner et al. 2011; Wit- tenmyer et al. 2012a; Wittenmyer et al. 2012b; Robertson et al. 2012a,b; Horner et al. 2012). We used the Hybrid integrator within the n-body dynamics package MERCURY (Chambers & Miglior- ini 1997; Chambers 1999), and followed the evolution of the two giant planets proposed by Lee et al. (2009) for a period of 100 mil- lion years. In order to examine the full range of allowed orbital solutions, we composed a grid of plausible architectures for the HW Vir planetary system, each of which tested a unique combi- nation of the system’s orbital elements, spanning the ± 3-σ range in the observed orbital parameters. Following our earlier work, the initial orbit of HW Vir c (the planet with the best constrained or- bit in Lee et al. (2009) was held fixed at its nominal best-fit values (i.e. a = 5.3 AU, e = 0.46 etc.). The initial orbit of HW Vir b was varied systematically such that the full 3-σ error ellipse in semi- major axis, eccentricity, longitude of periastron and mean anomaly were sampled. In our earlier work, we have found that the two main drivers of stability or instability were the orbital semi-major axis and eccentricity (e.g. Horner et al. 2011), and so we sampled the 3-σ region of these parameters in the most detail. In total, 45 distinct values of initial semi-major axis were tested for the orbit of HW Vir b, equally distributed across the full ± 3-σ range of allowed values. For each of these unique semi- major axes, 45 distinct eccentricities were tested, evenly distributed across the possible range of allowed values (i.e. between eccentrici- ties of 0.00 and 0.76). For each of the 2025 a−e pairs tested in this way, fifteen unique values of ω were tested (again evenly spread across the ± 3-σ range, whilst for each of the a−e−ω values, three unique values of mean anomaly were considered. In total, therefore, we considered 91,125 unique orbital configurations for HW Vir b, spread in a 45× 45× 15× 3 grid in a− e− ω− M space. In each of our simulations, the masses of the planets were set to their min- imal Msini values, in order to maximise the potential stability of their orbits. The orbital evolution of the planets was followed for a period of 100 million years, or until one of the planets was ei- ther ejected (defined by that planet reaching a barycentric distance of 20 AU), a collision between the planets occurred, or one of the planets collided with the central stars. If such a collision/ejection event occurred, the time at which it happened was recorded. In this way, the lifetime of each of the unique systems was determined. This, in turn, allowed us to construct a map of the dy- namical stability of the system, which can be seen in Fig 6. As can be seen in that figure, none of the orbital solutions tested were dy- namically stable, with few a-e locations displaying mean lifetimes longer than 1,000 years. 1 Remarkably, we find that the proposed orbits for the HW Vir planetary system are even less dynamically stable than those pro- posed for the now discredited planetary system around HU Aqr (Horner et al. 2011; Wittenmyer et al. 2012a; Hinse et al. 2012a; Horner et al. 2012; Go´zdziewski et al. 2012). Simply put, our result 1 The four yellow/orange/red hotspots in that plot between a ∼ 3.3 and a ∼ 4.2 AU are the result of four unusually stable runs, with lifetimes of 120 kyr (a = 3.31 AU,e = 0.12), 250 kyr (a = 3.74 AU,e = 0.22), 56 kyr (a = 4.02 AU,e = 0.19) and 49 kyr (a = 4.24 AU,e = 0.05). Such “long- live” outliers are not unexpected, given the chaotic nature of dynamical in- teractions, but given the typically very short lifetimes observed can signifi- cantly alter the mean lifetime in a given bin. c(cid:13) 2012 RAS, MNRAS 000, 1–?? proves conclusively that, if there are planets in the HW Vir system, they must move on orbits dramatically different to those proposed by Lee et al. (2009). This instability is not particularly surprising, given the high orbital eccentricity of planet c, which essentially en- sures that the two planets are on orbits that intersect one another, irrespective of the initial orbit of planet b. Given that the two plan- ets are not trapped within mutual mean-motion resonance (MMR), such an orbital architecture essentially guarantees that they will ex- perience strong close encounters within a very short period of time, ensuring the system’s instability. 4 ECLIPSE TIMING DATA ANALYSIS AND LTT MODEL Given the extreme instability exhibited by the planets proposed by Lee et al. (2009), it seems reasonable to ask whether a re-analysis of the observational data will yield significantly different (and more reasonable) orbits for the planets in question. We therefore chose to re-analyse the observational data, following a similar methodology as applied in an earlier study of HU Aqr (Hinse et al. 2012a). At the basis of our analysis we use the combined mid-eclipse timing data set compiled by Lee et al. (2009), including the times of secondary eclipses. The timing data used in Lee et al. (2009) were recorded in the UTC (Coordinated Universal Time) time stan- dard, which is known to be non-uniform (Bastian 2000; Guinan & Ribas 2001). To eliminate timing variations introduced by acceler- ated motion within the Solar System, we therefore transformed2 the HJD (Heliocentric Julian Date) timing records in UTC time stan- dard into Barycentric Julian Dates (BJD) within the Barycentric Dynamical Time (TDB) standard (Eastman et al. 2010). A total of 258 timing measurements were used spanning 24 years from Jan- uary 1984 (HJD 2 445 730.6) to May 2008 (HJD 2 454 607.1). We assigned 1-σ timing uncertainties to each data point by following the same approach as outlined in Lee et al. (2009). For an idealised, unperturbed and isolated binary system, the linear ephemeris of future/past mid-eclipse (usually primary) events can be computed from TC(E) = T0 + P0E, (1) where E denotes the (independent) ephemeris cycle number, T0 is the reference epoch, and P0 measures the eclipsing period ((cid:39) 2.8 hrs) of HW Vir. A linear regression performed on the 258 recorded light-curves allows P0 to be determined with high precision. In this work, we chose to place the reference epoch close to the middle of the observing baseline to avoid parameter correlation between T0 and P0 during the fitting process. In the following we briefly outline the LTT model as used in this work. 4.1 Analytic LTT model The model adopted in this work is similar to that described in Hinse et al. (2012a), and is based on the original formulation of a sin- gle light-travel time orbit introduced by Irwin et al. (1952). In this model the two components of the binary system are assumed to represent one single object with a total mass equal to the sum of the masses of the two stars. This point mass is then placed at the original binary barycentre. If a circumbinary companion exist, then the combined binary mass follows an orbit around the total system barycentre. The eclipses are then given by Eq. 1. This defines the LTT orbit of the binary. The underlying reference system has its origin at the total centre of mass. 2 http://astroutils.astronomy.ohio-state.edu/time 4 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney Following Irwin et al. (1952), if the observed mid-eclipse times exhibit a sinusoidal-like variation (due to one or more un- seen companion(s)), then the quantity O−C defines the light-travel time effect and is given by (O−C)(E) = TO(E)− TC(E) = 2 ∑ i=1 τi, (2) where TO denotes the measured time of an observed mid-eclipse, and TC is the computed time of that mid-eclipse based on a linear ephemeris. We note that τ1 + τ2 is the combined LTT effect from two separate two-body LTT orbits. The quantity τi is given by the following expression for each companion (Irwin et al. 1952): τi = Kb,i 1− e2 b,i 1 + eb,i cos fb,i sin( fb,i + ωb,i) + eb,i sinωb,i (3) (cid:105) , (cid:104) where Kb,i = ab,i sinIb,i/c is the semi-amplitude of the light-time effect (in the O−C diagram) with c measuring the speed of light and Ib,i is the line-of-sight inclination of the LTT orbit relative to the sky plane, eb,i the orbital eccentricity, fb,i the true longi- tude and ωb,i the argument of pericenter of the LTT orbit. The 5 model parameters for a single LTT orbit are given by the set (ab,i sinIb,i,eb,i,ωb,i,Tb,i,Pb,i). The time of pericentre passage Tb,i and orbital period Pb,i are introduced through the expression of the true longitude as a time-like variable via the mean anomaly M = nb,i(TO − Tb,i), with nb,i = 2π/Pb,i denoting the mean mo- tion of the combined binary in its LTT orbit. Computing the true anomaly as a function of time (or cycle number) requires the solu- tion of Kepler’s equation. We direct the interested reader to Hinse et al. (2012a) for further details. In Eq. 3, the origin of the coordinate system is placed at the centre of the LTT orbit (see e.g. Irwin et al. 1952). A more natural choice (from a dynamical point of view) would be to use the system centre of mass as the origin of the coordinate system. However, the derived Keplerian elements are identical in the two coordinate systems (e.g. Hinse et al. 2012b). Finally, we note that our model does not include mutual gravitational interactions. We also only consider the combination of two LTT orbits from two circumbinary companions. From first principles, some similarities exist between the LTT orbit and the orbit of the circumbinary companion. First, the ec- centricities (eb,i = ei) and orbital periods (Pb,i = Pi) are the same. Second, the arguments of pericenter are 180◦ apart from one an- other (ωi = 180◦ − ωb,i). Third, the times of pericenter passage are also identical (Tb,i = Ti). Information of the mass of the unseen companion can be ob- tained from the mass function given by f (mi) = 4π2(ab,i sinIb,i)3 4π2(Kb,ic)3 GP2 b,i = GP2 b,i = (mi sinIi)3 (mb + mi) , i = 1,2 (4) The least-squares fitting process provides a measure for Kb,i and Pb,i, and hence the minimum mass of the companion can be found from numerical iteration. In the non-inertial astrocentric reference frame, with the combined binary mass at rest, the companion’s semi-major axis relative to the binary is then calculated using Ke- pler’s third law. In Lee et al. (2009) the authors also accounted for additional period variations due to mass transfer and/or magnetic interactions between the two binary components. These variations usually occur on longer time scales compared to orbital period variations due to unseen companions. Following Hilditch (2001), the corresponding ephemeris of calculated times of mid-eclipses then takes the form i = 1,2 TC = T0 + P0E + βE2 + τi, (5) where β is an additional free model parameter and accounts for a secular modulation of the mid-eclipse times resulting from interac- tions between the binary components. Assuming the timing data of HW Vir are best described by a two-companion system, and to be consistent with Lee et al. (2009), we have used Eq. 5 as our model which consists of 13 parameters. 5 METHODOLOGY AND RESULTS FROM χ2-PARAMETER SEARCH To find a stable orbital configuration of the two proposed circumbi- nary companions, we carried out an extensive search for a best-fit in χ2 parameter space. The analysis, methodology and technique follow the same approach as outlined in Hinse et al. (2012a). Here we briefly repeat the most important elements in our analysis. We used the Levenberg-Marquardt least-square minimisation algorithm as implemented in the IDL3-based software package MPFIT4 (Markwardt 2009). The goodness-of-fit statistic χ2 of each fit was evaluated from the weighted sum of squared errors. In this work, we use the reduced chi-square statistic χ2 r which takes into account the number of data points and the number of freely varying model parameters. We seeded 28,201 initial guesses within a Monte Carlo experi- ment. Each guess was allowed a maximum of 500 iterations before termination, with all 13 model parameters (including the secular r (cid:54) 10 were term) kept freely varying. Converged solutions with χ2 accepted with the initial guess and final fitting parameters recorded to a file. After each converged iteration we also solved the mass function (Eq. 4) for the companion’s minimum mass and calculated the semi-major axis relative to the system barycentre from Kepler’s third law. Initial guesses of the model parameters were chosen at random following either a uniform or normal distribution. For example, ini- tial orbital eccentricities were drawn from a uniform distribution within the interval e ∈ [0.0,0.8]. Our initial guesses for the orbital periods were guided by a Lomb-Scargle (LS) discrete Fourier trans- formation analysis on the complete timing data set. For the LS anal- ysis we used the PERIOD04 software package (Lenz & Breger 2005) capable of analysing unevenly sampled data sets. Fig. 1 shows the normalised LS power-spectrum. The LS algorithm found two significant periods with frequencies f1 (cid:39) 1.4×10−4 cycles/day and f2 (cid:39) 2.1× 10−4 cycles/day. These frequencies correspond to periods of 7397 and 4672 days, respectively. Hence the short-period variation is covered more than twice during the observing interval. Due to a lower amplitude, it contains less power within the data set. Our random initial guesses for the companion periods were then drawn from a Gaussian distribution centred at these periods with standard deviation of ± 5 years. We call this approach a “quasi- global” search of the underlying χ2-parameter space. 5.1 Results - finding best fit and confidence levels Our best fit model resulted in a χ2 r = 0.943 and is shown in Fig. 2 along with the LTT signal due to the inner and outer companion and the secular term. The corresponding root-mean-square (RMS) 3 The acronym IDL stands for Interactive Data Language and is a trademark of ITT Visual Information Solutions. For further details see http://www.ittvis.com/ProductServices/IDL.aspx. 4 http://purl.com/net/mpfit c(cid:13) 2012 RAS, MNRAS 000, 1–?? A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 5 scatter of data around the best fit is 8.7 seconds, which is close to the RMS scatter reported in Lee et al. (2009). The fitted model elements and derived quantities of our best fit are shown in Ta- ble 2. Compared to the system in Lee et al. (2009) we note that we now obtain a lower eccentricity for the inner companion and a larger eccentricity for the outer companion. Furthermore, our two- companion system has also slightly expanded, with larger semi- major axes (and therefore longer orbital periods) for both compan- ions compared to Lee et al. (2009). The next question to ask is how reliable or significant our best-fit solution is in a statistical sense. Assuming that the errors are normally distributed one can establish confidence levels for a multi-parameter fit (Bevington & Robinson 1992). We therefore carried out detailed two-dimensional parameter scans covering a large range around the best-fit value in order to study the χ2 r -space topology in more detail. In particular, we explored relevant model parameter combinations including T0,P0 and β. In all our experiments we allowed the remaining model param- eters to vary freely while fixing the two parameters of interest in the considered parameter range (Bevington & Robinson 1992; Press et al. 1992). Assuming parameter errors are normally distributed, our 1-, 2- and 3-σ level curves provide the 68.3, 95.4 and 99.7% confidence levels relative to our best-fit, respectively. In Fig. 3 we show a selection of our two-dimensional parameter scan consider- ing various model parameters. Ideally, one would aim to work with parameters with little correlation between the two parameters. The lower-right panel in Fig. 3 shows the relationship between T0 and P0. The near circular shape of the level curves reveals that little correlation between the two parameters exists. This is most likely explained by our choice to locate the reference epoch in the middle of the dataset. The re- maining panels in Fig. 3 show some correlations between the pa- rameters. However, we have some indication of an unconstrained outer orbital period in the lower-left panel of Fig. 3. We show the location of orbital mean motion resonances in the (P1,P2) plane. Our best fit is located close to the 2:1 mean motion resonance. To demonstrate that the outer period is unconstrained we have generated χ2 r -parameter scans in the (P2,e2)-plane as shown in Fig. 4. Our best fit model is shown in the left-most panel of Fig. 4, along with the 1-,2- and 3-σ confidence levels. It is readily evident that the 1-σ confidence level does not simply surround our best-fit model in a confined or ellipsoidal manner. We rather observe that all three level curves are significantly stretched towards solutions featuring longer orbital periods for the outer companion. This is demonstrated in the middle and right panels of Fig. 4. Any longer orbital period for the outer companion therefore results in a χ2 r with similar statistical significance as our best-fit model. In addition, we studied the (asini,P)-parameter plane for the outer companion, as shown in the lower-middle panel of Fig. 3. Here, we also observe that asini is unconstrained. The uncon- strained nature of the outer companion’s asini and orbital period has a dramatic effect on the derived minimum mass and the corre- sponding error bounds. 5.2 Results - parameter errors When applying the LM algorithm formal parameter errors are ob- tained from the best-fit covariance matrix. However, in our study, we sometimes encounter situations where some of the matrix el- ements become zero, or have singular values. However, at others times, the error matrix is returned with non-zero elements. In those cases we have observed that the outer orbital period is often better c(cid:13) 2012 RAS, MNRAS 000, 1–?? determined than that of the inner companion. In the case of HW Vir, such solutions are clearly incongruous, given the relatively poor orbital characterisation of the outer body compared to that of the innermost. Being suspicious about the formal covariance errors, we have resorted to two other methods to determine parameter errors. First, we attempted to determine errors by the use of the bootstrap method (Press et al. 1992). However, we found that the resulting error ranges are comparable to the formal errors extracted from the best- fit covariance matrix. Furthermore, the bootstrap error ranges were clearly incompatible with the 1-σ error “ellipses” discussed above. For example, from the top-left panel in Fig. 3 we estimate the 1-σ error on the inner orbital period to be on the order of 50-100 days. In contrast, the errors for the inner orbital period obtained from our bootstrap method were of order just a few days. For this reason, we consider the bootstrap method to have failed, and it has therefore not been investigated further in this study. However, it is interest- ing to speculate on the possibility of dealing with a dataset which is characterised by “clumps of data”, as seen in Fig. 2. When gen- erating random (with replacement) bootstrap ensembles, there is the possibility that only a small variation is being introduced for each random draw, as a result of the clumpiness of the underlying dataset. That clumpiness might be mitigated for, and the bootstrap method rendered still viable for the establishment of reliable errors, by enlarging the number of bootstrap ensembles to compensate for the lack of variation within single bootstrap data sets. One other possibility would be to replace clumps of data by a single data point reflecting the average of the clump. However, we have instead in- voked a different approach. Having located a best-fit minimum, we again seeded a large number of initial guesses around the best-fit parameters. This time, we considered only a relatively narrow range around the best-fit values (e.g. as shown in Fig. 3 and Fig. 4. This ensured that the LM algorithm would iterate towards our best-fit model depend- ing on the underlying χ2 r topology and inter-parameter correlations. The initial parameter guesses were randomly drawn from a uniform distribution within a given parameter interval. We then iterated to- wards a best-fit value using LM, and recorded the best-fit parame- ters along with the corresponding χ2 r . Generating a large number of guesses enabled us to establish statistics on the final best-fit parameters with χ2 r within 1-, 2-, and 3-σ confidence levels. We therefore performed a Monte Carlo ex- periment that considered several tens of thousands of guesses. To establish 1-σ error bounds (assuming a normal distribution for each parameter), we then considered only those models that yielded χ2 r within the 1-σ confidence limits (inner level curves), as shown in Fig. 3 and Fig. 4. The error for a given parameter is then obtained from the mean and standard deviation, and listed in Table 2. In or- der to test our assumption of normally distributed errors, we plotted histograms for the various model parameters in Fig. 5. For each his- togram distribution, we fitted a Gaussian and established the corre- sponding mean and standard deviation. While some parameters fol- low a Gaussian distribution (for example the outer companion’s ec- centricity), other parameters show no clear sign of “Gaussian tails”. This is especially true for the orbital period of the outer compan- ion. Following two independent paths of analysis, we have demon- strated that the outer period is unconstrained, based on the present dataset, and should be regarded with some caution. However, we also point out a short-coming of our method of determining random parameter errors. The parameter estimates depend on the proximity of the starting parameters to the best-fit parameter. In principle, our method of error determination assumes that the best-fit model pa- 6 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney rameters are well-determined in terms of well-established closed- loop confidence levels around the best-fit parameters. The true ran- dom parameter error distribution for the two ill-constrained param- eters might turn out differently. To put stronger constraints on the model parameters is clearly only possible by augmenting the ex- isting timing data through a program of continuous monitoring of HW Vir over the coming years (Pribulla et al. 2012; Konacki et al. 2012). As more data is gathered, the confidence levels in Fig. 4 will eventually narrow down. 6 THE β COEFFICIENT AND ANGULAR MOMENTUM LOSS One of the features of our best-fit orbital solution is that it results in a relatively large β coefficient, which can be related to a change in the period of the binary resulting from additional, non-planetary ef- fects. Potential causes of such a period change include mass trans- fer, loss of angular momentum, magnetic interactions between the two binary components and/or perturbations from a third body on a distant and unconstrained orbit. In this study, the β factor repre- sents a constant binary period change (see Hilditch 2001, p. 171) with a linear rate of dP/dt = −9.57×10−9 days/yr, which is about 15 percent larger than the value reported in Lee et al. (2009). We retained the β coefficient in our model in order that our treatment be consistent with that detailed in Lee et al. (2009), such that our results might be directly compared to their work. Lee et al. (2009) examined a number of combinations of mod- els that incorporated a variety of potential causes for the observed period modulation. They found that the timing data is best de- scribed by two LTT and a quadratic term in the linear ephemeris model. In their work, Lee et al. (2009) carefully examined the contribution of period modulation by various astrophysical effects. They were able to provide arguments that rule out the operation of the Applegate mechanism, due to the lack of small-scale variations in the observed luminosities that would have an influence on the J2 oblateness coefficient of the magnetically active component. A change in J2 would, in turn, affect the binary period. Furthermore, Lee et al. (2009) reject the idea that the ob- served O−C variation could be the result of apsidal motion, based on the circular orbit of the HW Vir binary system. In addition, they estimated the secular period change of the HW Vir binary or- bit due to angular momentum loss through gravitational radiation and magnetic breaking. They found that the most likely explana- tion for the observed linear decrease in the binary period is that it is the result of angular momentum loss by magnetic stellar wind breaking in the secondary M-type component. From first princi- ples (e.g. Brinkworth et al. 2006), the period change observed in this work corresponds to a angular momentum change of order dJ/dt = −2.65× 1036 erg. This is approximately 15 percent larger than the value reported by Lee et al. (2009), but still well within the range where magnetic breaking is a reasonable astrophysical cause for period modulation. Finally, we note that, whilst this work was under review, Beuermann et al. (2012) published a similar study based on new timing data, in which they also considered the influence of period changes due to additional companions. In their work, they obtained a markedly different orbital solution than those discussed in this work, one which they found to be dynamically stable on relatively long timescales. In light of their findings, it is interesting to note that they do not include a quadratic term in their linear ephemeris model. This could point at the possibility that the inclusion of a quadratic term is somehow linked to the instability of the best-fit system found in this work. Beuermann et al. (2012) found stable orbits for a solution in- volving two circumbinary companions. However, despite this, we note that the two models share some qualitative characteristics. A careful examination of Fig. 2 in Beuermann et al. (2012) reveals that the orbital period of the outer companion is unconstrained from a period analysis since the χ2-contour curves are open towards longer outer orbital periods - a result mirrored in our current work. As we noted earlier, we were unable to place strict confidence lev- els on the best-fit outer companion’s orbital period and semi-major axis. Although Beuermann et al. (2012) do find a range of stable scenarios featuring their outer companion, we note that they fixed the eccentricity of that companion’s orbit to be near-circular, with period of 55 years. Such an assumption (i.e. fixing some orbital pa- rameters) is somewhat dangerous, since it can lead to the produc- tion of dynamically stable solutions that are not necessarily sup- ported by the observational data (e.g. Horner et al. 2012). A more rigorous strategy would be to generate an ensemble of models with each model (all parameters freely varying) tested for orbital stabil- ity (using some criterion like non-crossing orbits or non-overlap of mean-motion resonances, etc.) resulting in a distribution of stable best-fit models. Using the new data set, a study of the distribution of the best-fit outer planet’s eccentricity would be interesting. It is certainly possible that the new data set constrains this parameter sufficiently in order to validate their assumptions. Although the re- sults presented in Beuermann et al. (2012) are clearly promising, it is definitely the case that more observations are needed before the true origin of the observed variation for HW Vir is established beyond doubt. The question of whether a period damping factor is truly nec- essary for the HW Vir system would require a statistically self- consistent re-examination of the complete data set taking account of a range of model scenarios. We refer the interested reader to Go´zdziewski et al. (2012), who recently carried out a detailed in- vestigation of the influence of the quadratic term for various sce- narios in their attempt to explain the timing data of the HU Aqr system. 7 DYNAMICAL ANALYSIS OF THE BEST-FIT LTT MODEL Since a detailed re-analysis of the observational data on the HW Vir system yields a new orbital solution for the system, it is interesting to consider whether that new solution offers better prospects for dy- namical stability than that proposed in Lee et al. (2009). We there- fore repeated our earlier dynamical analysis using the new orbital solution. Once again, we held the initial orbit of planet HW Vir c fixed at the nominal best-fit solution, and ran an equivalent grid of unique dynamical simulations of the planetary system, varying the initial orbit of HW Vir b such that a total of 45 distinct values of a and e, 15 distinct values of ω and 3 values of M were tested, each distributed evenly as before across the ± 3-σ range of allowed val- ues. As before, the two simulated planets were assigned the nomi- nal M sin i masses obtained from the orbital model. The results of our simulations can be seen in Fig. 7. As was the case for the original orbital solution proposed in Lee et al. (2009), and despite the significantly reduced uncertain- ties in the orbital elements for the resulting planets, very few of the tested planetary systems survived for more than 1,000 years (with just twenty six systems, 0.029% of the sample, surviving for more than 3,000 years, and just three systems surviving for more than c(cid:13) 2012 RAS, MNRAS 000, 1–?? A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 7 10,000 years). As was the case for the planetary system proposed to orbit the cataclysmic variable system HU Aqr (e.g. Horner et al. 2011; Wittenmyer et al. 2012a), it seems almost certain that the pro- posed planets in the HW Vir system simply do not exist - at least on orbits resembling those that can be derived from the observational data. 8 CONCLUSION AND DISCUSSION The presence of two planets orbiting the evolved binary star sys- tem HW Vir was proposed by Lee et al. (2009), on the basis of periodic variations in the timing of eclipses between the two stars. The planets proposed in that work were required to move on rela- tively eccentric orbits in order to explain the observed eclipse tim- ing variations, to such a degree that the orbit of the outer planet must cross that of the innermost. It is obvious that, when one ob- ject moves on an orbit that crosses that of another, the two will eventually encounter one another, unless they are protected from such close encounters by the influence of a mutual mean-motion resonance (e.g. Horner et al. 2004a,b). Even objects protected by the influence of such resonances can be dynamically unstable, al- beit on typically longer timescales (e.g. Horner & Lykawka 2010; Horner, Muller & Lykawka 2012; Horner et al. 2012). Since the two planets proposed by Lee et al. (2009) move on calculated or- bits that allow the them to experience close encounters and yet are definitely not protected from such encounters by the influence of mutual mean-motion resonance, it is clear that they are likely to be highly dynamically unstable. To test this hypothesis, we performed a suite of highly detailed dynamical simulations of the proposed planetary system to examine its dynamical stability as a function of the orbits of the proposed planets. We found the proposed system to be dynamically unstable on extremely short timescales, as was expected based on the proposed architecture for the system. Following our earlier work (Wittenmyer et al. 2012a; Hinse et al. 2012a), we performed a highly detailed re-analysis of the observed data, in order to check whether such improved analy- sis would yield better constrained orbits that might offer better prospects for dynamical stability. Our analysis resulted in calcu- lated orbits for the candidate planets in the HW Vir system that have relatively small uncertainties. Once these orbits had been obtained, we performed a second suite of detailed dynamical simulations to ascertain the dynamical stability of the newly determined orbits. Following the same procedure as for the original orbits, we con- sidered the stability of all plausible architectures for the HW Vir system. Despite the increased precision of the newly determined orbits, we find that the planetary system proposed is dynamically unstable on timescales as short as a human lifetime. For that rea- son, we must conclude that the eclipse-timing variations observed in the HW Vir system are not solely down to the gravitational in- fluence of perturbing planets. Furthermore, if any planets do exist in that system, they must move on orbits dramatically different to those considered in this work. Our results highlight the importance of performing comple- mentary dynamical studies of any suspected multiple-exoplanet system – particularly in those cases where the derived planetary or- bits approach one another closely, are mutually crossing and/or de- rived companion masses are large. Following a similar strategy as applied to the proposed planetary system orbiting HU Aqr (Horner et al. 2011; Wittenmyer et al. 2012a; Go´zdziewski et al. 2012), we have found that the proposed 2-planet system around HW Vir does not stand up to a detailed dynamical scrutiny. In this work we have shown that the outer companion’s period (among other parameters) c(cid:13) 2012 RAS, MNRAS 000, 1–?? is heavily unconstrained by establishing confidence limits around our best-fit model. However, we also point out the fact that the two circumbinary companions have brown-dwarf masses. Hence, a more detailed n-body LTT model which takes account of mutual gravitational interactions might provide a better description of the problem. To further characterise the HW Vir system and constrain or- bital parameters we recommend further observations within a mon- itoring program as described in Pribulla et al. (2012). In a recent work on HU Aqr, Go´zdziewski et al. (2012) pointed out the pos- sibility that different data sets obtained from different telescopes could introduce systematic errors resulting in a false-positive de- tection of a two-planet circumbinary system. Finally, we note that, whilst this paper was under referee, Beuermann et al. (2012) independently published their own new study of the HW Vir system. Based on new observational timing data, those authors determined a new LTT model that appears to place the 2-planet system around HW Vir on orbits that display long-term dynamical stability. Based on the results presented in this work, we somewhat doubt their findings of a stable 2-planet system and question whether such a system is really supported by the new data set given that no strict confidence levels were found for the best-fit outer period. Since performing a full re-analysis of their newly compiled data, including dynamical mapping of their new architecture, would be a particularly time intensive process, we have chosen to postpone this task for a future study. ACKNOWLEDGMENTS The authors wish to thank an anonymous referee, whose extensive comments on our work led to significant changes that greatly im- proved the depth of this work. We also thank Dr. Lee Jae Woo for useful discussions and suggestions that resulted in the cre- ation of Fig. 1. JH gratefully acknowledges the financial support of the Australian government through ARC Grant DP0774000. RW is supported by a UNSW Vice-Chancellor’s Fellowship. JPM is partly supported by Spanish grants AYA 2008/01727 and AYA 2011/02622.. TCH gratefully acknowledges financial support from the Korea Research Council for Fundamental Science and Tech- nology (KRCF) through the Young Research Scientist Fellowship Program, and also the support of the KASI (Korea Astronomy and Space Science Institute) grant 2012-1-410-02. The dynamical sim- ulations performed in this work were performed on the EPIC super- computer, supported by iVEC, located at the Murdoch University, in Western Australia. The Monte Carlo/fitting simulations were car- ried out on the “Beehive” computing cluster at Armagh Observa- tory (UK) and the “Pluto” high performance computing cluster at the Korea Astronomy and Space Science Institute. Astronomical research at Armagh Observatory (UK) is funded by the Department of Culture, Arts and Leisure. REFERENCES Anglada-Escud´e, G., Arriagada, P., Vogt, S. S., et al. 2012, arXiv:1202.0446 Applegate, J. H. 1992, ApJ, 385, 621 Bakos, G. ´A., Noyes, R. W., Kov´acs, G., et al. 2007, ApJ, 656, 552 Barlow, B. N., Dunlap, B. H., Clemens, J. C., 2011 ApJ, 737, 2 Bastian, U. 2000, IBVS, 4822 Beuermann K., et al. 2010, A&A, 521, L60 Beuermann K., et al. 2012, A&A, in press, 2012arXiv1206.3080B Bevington, P. R., & Robinson, K. D., McGraw-Hill, 2nd ed., 1992 8 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 728, 117 Brinkworth, C. S., Marsh, T. R., Dhillon, V. S., Knigge, C., 2006, MNRAS, 365, 287 Chambers, J. E., Migliorini, F. 1997, BAAS, 27-06-P Chambers, J. E. 1999, MNRAS, 304, 793 Doyle, L. R., Carter, J. A., Fabrycky, D. C., et al. 2011, Science, 333, 1602 Eastman, J., Siverd, R., Gaudi, B. S. 2010, PASP, 122, 935 Endl, M., Cochran, W. D., Wittenmyer, R. A., & Hatzes, A. P. 2006, AJ, 131, 3131 Ferraz-Mello, S., Michtchenko, T. A., & Beaug´e, C. 2005, ApJ, 621, 473 Funk, B., et al. 2011, (submitted to MNRAS) Funk, B., Eggl, S., Gyergyovits, M., Schwarz, R., & Pilat- Lohinger, E. 2011, EPSC-DPS Joint Meeting 2011, 1725 Go´zdziewski, K., & Maciejewski, A. J. 2001, ApJL, 563, L81 Go´zdziewski, K., Konacki, M., Maciejewski, A. J., 2003, ApJ, 594, 1019 Go´zdziewski, K., Nasiroglu, I., Slowikowska, A., et al. 2012, MN- RAS, in press, (eprint arXiv:1205.4164) Go´zdziewski, K., Konacki, M., Maciejewski, A. J., 2005, ApJ, 622, 1136 Guinan, E. F., Ribas, I., 2001, ApJL, 546, L43 Hellier, C., Anderson, D. R., Collier Cameron, A., et al. 2011, A&A, 535, L7 Hilditch, R. W., “An introduction to close binary stars”, 2001, 1st ed., Cambridge University Press Hinse, T. C., Lee, J. W., Go´zdziewski, K., Haghighipour, N., Lee, C.-U., Scullion, E. M., 2012, MNRAS, 420, 3609 Hinse, T. C., Go´zdziewski, K., Lee, J. W., Haghighipour, N., Lee, C.-U., 2012, AJ, 144, 34 Horner, J., Evans, N. W., & Bailey, M. E. 2004, MNRAS, 354, 798 Horner, J., Evans, N. W., & Bailey, M. E. 2004, MNRAS, 355, 321 Horner, J., & Lykawka, P. S. 2010, MNRAS, 405, 49 Horner, J., Muller, T. G. & Lykawka, P. S., 2012, MNRAS, 423, 2587 Horner, J., Lykawka, P. S., Bannister, M. T., & Francis, P., 2012, MNRAS, 422, 2145 Horner, J., Marshall, J., Wittenmyer, R., Tinney, Ch. 2011, MN- RAS, 416, L11 Horner, J., Wittenmyer, R. A., Hinse, T. C., & Tinney, C. G. 2012, MNRAS, in press Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, ApJ, 721, 1467 Howard, A. W., Bakos, G. ´A., Hartman, J., et al. 2012, ApJ, 749, 134 Ibanoglu, C., C¸ akırlı, O., Tas¸, G., & Evren, S. 2004, A&A, 414, 1043 Irwin, J. B., 1952, ApJ, 116, 211 Irwin, J. B., 1959, AJ, 64, 149 Kilkenny, D., Marang, F., & Menzies, J. W. 1994, MNRAS, 267, 535 Kilkenny, D., van Wyk, F., & Marang, F. 2003, The Observatory, 123, 31 Konacki, M., Sybilski, P., Kozłowski, S. K., Ratajczak, M., & Hełminiak, K. G. 2012, IAU Symposium, 282, 111 Laskar, J., & Correia, A. C. M. 2009, A&A, 496, L5 Lee J., et al., 2009, AJ, 137, 3181 Lee J., et al., 2012, AJ, 143, 34 Lenz P., Breger M. 2005, CoAst, 146, 53 Malmberg, D., Davies, M. B., 2009, MNRAS, 394, 26 Markwardt, C. B., 2009, ASPCS, “Non-linear Least-squares Fit- ting in IDL with MPFIT”, Astronomical Data Analysis Software and Systems XVIII, eds. Bohlender, D. A., Durand, D., Dowler, P. Marshall, J., Horner, J., & Carter, A. 2010, International Journal of Astrobiology, 9, 259 Mayor, M., Queloz, D., 1995, Nature, 378, 355 Mayor, M., Udry, S., Lovis, C., et al. 2009, A&A, 493, 639 Menzies, J. W., & Marang, F. 1986, Instrumentation and Research Programmes for Small Telescopes, 118, 305 Murray, C. D., Dermott, S. F., 2001, Solar System Dynamics, Cambridge University Press. Pepe, F., Mayor, M., Queloz, D., et al. 2004, A&A, 423, 385 Potter, S. B., Romero-Colmenero, E., Ramsay, G., Crawford, S., Gulbis, A., Barway, S., Zietsman, E., Kotze, M., Buckley, D. A. H., O’Donoghue, D., Siegmund, O. H. W., McPhate, J., Welsh, B. Y., Vallerga, J., MNRAS, 416, 2202 Press, W. H., Saul, A. T., Vetterling, W. T., Flannery, B. P., Cam- bridge University Press, 2nd edition, 1992. Pribulla, T., Vanko, M., Ammler- von Eiff, M., Andreev, M. et al. 2012, AN (submitted), eprint arXiv:1206.6709 Rivera, E. J., Laughlin, G., Butler, R. P., et al. 2010, ApJ, 719, 890 Robertson, P., Endl, M., Cochran, W. D., et al. 2012a, ApJ, 749, 39 Robertson, P., Horner, J., Wittenmyer, R. A. et al. 2012b, ApJ, 754, 50. Schwarz, R., Schwope, A. D., Vogel, J., Dhillon, V. S., Marsh, T. R., Copperwheat, C., Littlefair, S. P., Kanbach, G., 2009, A&A, 496, 833 Schwope, A. D., Schwarz, R., Sirk, M., Howell, S. B., 2001, A&A, 375, 419 Schwope, A. D.,Horne, K., Steeghs, D., Still, M., A&A, 531, 34 Smith, A. M. S., Anderson, D. R., Collier Cameron, A., et al. 2012, A. J., 143, 81 Stepinski, T. F., Malhotra, R., & Black, D. C. 2000, ApJ, 545, 1044 Tinney, C. G., Butler, R. P., Marcy, G. W., et al. 2001, ApJ, 551, 507 Tinney, C. G., Wittenmyer, R. A., Butler, R. P., et al. 2011, ApJ, 732, 31 Qian, S.-B., et al., 2011, MNRAS, 414, L16 Qian, S.-B., Liao, W.-P., Zhu, L.-Y., & Dai, Z.-B. 2010, ApJL, 708, L66 Udry, S., Bonfils, X., Delfosse, X., et al. 2007, A&A, 469, L43 Welsh, W. F., Orosz, J. A., Carter, J. A., et al. 2012, Nature, 481, 475 Wittenmyer, R. A., Horner, J., Marshall, J. P., Butters, O. W., & Tinney, C. G. 2012a, MNRAS, 419, 3258 Wittenmyer, R. A., et al., 2012b, ApJ, in press Wolszczan, A., Frail, D. A., 1992, Nature, 355, 145 Wood, J. H., & Saffer, R. 1999, MNRAS, 305, 820 Wright, J. T., Veras, D., Ford, E. B., et al. 2011, ApJ, 730, 93 c(cid:13) 2012 RAS, MNRAS 000, 1–?? A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 9 Figure 1. Power spectrum of HW Vir timing data set using BJD(TDB) times with the O−C residuals measured in seconds. Two periods were found with 1/ f1 = 7397 and 1/ f2 = 4672 days, corresponding to 20.3 and 12.8 years, respectively. S/N denotes the frequency signal-to-noise ratio, which are significantly larger than the spectrum’s noise level. Normalisation was done by division of the maximum amplitude in each spectrum. The f2-frequency was determined from the residuals after subtracting the period 1/ f1 from the original timing data set. Additional peaks in both panels represent 1-year alias frequencies due to the repeating annual observing cycle of HW Vir. c(cid:13) 2012 RAS, MNRAS 000, 1–?? 10 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney Figure 2. Best-fit two-Kepler LTT model with χ2 root-mean-square scatter around the best fit. The lower part of the figure shows the residuals between the best fit model and the observed timing data set. r = 0.943. The best-fit model parameters (including reference epoch) are shown in Table 2.RMS denotes the Figure 3. Colour-coded χ2 symbol. Contour curves show the 1,2,3σ confidence level curves around our best-fit model. See electronic version for colour figures. r parameter scans of orbital parameters with remaining parameters to vary freely. The best-fit parameter is indicated by a star-like c(cid:13) 2012 RAS, MNRAS 000, 1–?? 3600 3700 3800 3900 4000 4100 4200 4300initial inner period, Pb,1 (days) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8initial inner eccentricity, eb,1 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂ 0.05 0.06 0.07 0.08 0.09 0.1 0.11initial ab,1sin Ib,1 (AU)-1.7-1.65-1.6-1.55-1.5-1.45-1.4initial β (10-12 days/cycle2) 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂-400-200 0 200 400initial Tb,1 (BJD(TT)-2448880.0)-0.6-0.4-0.2 0 0.2 0.4 0.6initial ωb,1 (radians) 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂ 5e-05 0.0001 0.00015 0.0002 0.00025initial reference epoch, T0 (BJD + 2,450,280.2858) 2e-09 4e-09 6e-09 8e-09 1e-08 1.2e-08initial binary period, P0 (days + 0.116719519) 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂ 0.15 0.2 0.25 0.3 0.35 0.4initial outer semi-major axis, ab,2sin Ib,2 (AU) 8000 10000 12000 14000 16000 18000initial outer period, Pb,2 (days) 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂ 2000 4000 6000 8000 10000 12000 14000 16000 18000initial outer period, Pb,2 (days) 3000 3500 4000 4500 5000initial inner period, Pb,1 (days) 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂1:12:15:23:14:1 A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 11 Figure 4. Same as Fig. 3 but considering the (P2,e2)-plane. The middle and right panel shows the χ2 Based on the used data set, no firm confidence levels can be established around out best-fit value. r topology for longer orbital periods of the outer companion. r < 1− σ where considered to Figure 5. Histogram distribution of six model parameters as obtained from a Monte Carlo experiment. Only models with χ2 assess the 68% confidence levels for each parameter. Solid curves show fitted normal distribution with mean and standard deviation indicated in each panel. However, we used the mean and standard deviation derived from the underlying dataset to derive our 1-σ errors as quoted in Table 2. c(cid:13) 2012 RAS, MNRAS 000, 1–?? 6000 6500 7000 7500 8000 8500 9000 9500 10000initial outer period, Pb,2 (days) 0.3 0.4 0.5 0.6 0.7 0.8 0.9initial outer eccentricity, eb,2 1 1.2 1.4 1.6 1.8 2 2.2 2.4❂ 10000 10500 11000 11500 12000 12500 13000 13500 14000initial outer period, Pb,2 (days) 0.3 0.4 0.5 0.6 0.7 0.8 0.9initial outer eccentricity, eb,2 1 1.2 1.4 1.6 1.8 2 2.2 2.4 14000 14500 15000 15500 16000 16500 17000 17500 18000initial outer period, Pb,2 (days) 0.3 0.4 0.5 0.6 0.7 0.8 0.9initial outer eccentricity, eb,2 1 1.2 1.4 1.6 1.8 2 2.2 2.4 12 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney Figure 6. The stability of the HW Vir planetary system as proposed by Lee et al. (2009), as a function of the semi-major axis, a, and eccentricity, e, of planet HW Vir b. The initial orbit of HW Vir c was the same in each integration, set to the nominal best fit orbit from that work. The mean lifetime of the planetary system (in log10(li f etime/yr)) at a given a− e co-ordinate is denoted by the colour of the plot. The lifetime at each a− e location is the mean value of 45 separate integrations carried of orbits at that a− e position (testing a combination of 15 unique ω values, and 3 unique M values). The nominal best-fit orbit for HW Vir b is located within the open square, from which lines radiate showing the extend of the ± 1-σ errors on a and e. As can be seen, the orbits of the system are incredibly unstable, no matter what initial orbit is considered for HW Vir b. c(cid:13) 2012 RAS, MNRAS 000, 1–?? A Dynamical Analysis of the Proposed Circumbinary HW Virginis Planetary System 13 Figure 7. The stability of the HW Vir planetary system, given the orbital solution derived in this work, as a function of the semi-major axis, a, and eccentricity, e, of planet HW Vir b. The initial orbit of HW Vir c was the same in each integration, set to the nominal best fit orbit as detailed in Table 2. The mean lifetime of the planetary system (shown as log10(li f etime/yr) at a given a− e co-ordinate is denoted by the colour of the plot. The lifetime at each a− e location is the mean value of 45 separate integrations carried of orbits at that a− e position (testing a combination of 15 unique ω values, and 3 unique M values). The nominal best-fit orbit for HW Vir b is located within the open square, from which lines radiate showing the extend of the ± 1-σ errors on a and e. Once again, the orbits of the system are found to be incredibly unstable, no matter what initial orbit is considered for HW Vir b. The two red hotspots in that plot are the result of two unusually stable runs, with lifetimes of 33 kyr (a = 4.185 AU, e = 0.137) and 38 kyr (a = 4.15 AU, e = 0.11). Even these most extreme outliers are dynamically unstable on astronomically short timescales. c(cid:13) 2012 RAS, MNRAS 000, 1–?? 14 J. Horner, T. C. Hinse, R.A. Wittenmyer, J. P. Marshall & C.G. Tinney Parameter two-LTT τ1 (i = 1) τ2 (i = 2) χ2 r RMS β T0 P0 ab,i sinIb,i eb,i (or e1,2) ωb,i Tb,i Pb,i (or P1,2) Kb,i mi sinIi ai sinIi ei ωi Ti Pi 0.943 8.665 −1.529· 10−12 ± 1.25· 10−13 2,450,280.28596± 2.3· 10−5 0.116719519± 4.6· 10−9 0.081± 0.002 0.17± 0.02 0.05± 0.01 2,448,880± 57 4021± 64 0.196± 0.012 0.61± 0.02 2.09± 0.08 2,448,629± 42 7992± 551 (!) 4.6· 10−4 ± 1.3· 10−5 1.13· 10−3 ± 7.04· 10−5 12± 3 4.26± 0.05 0.17± 0.02 (π− 0.05)± 0.01 2,448,880± 57 4021± 64 11± 8 6.8± 0.3 0.61± 0.02 (π− 2.09)± 0.08 2,448,629± 42 7992± 551 (!) Unit - seconds day/cycle2 BJD(TDB) days AU - rad. BJD(TT) days days MJup AU - rad. BJD(TT) days Table 2. Best-fit parameters for the LTT orbits of HW Vir corresponding to Fig. 2. Subscripts 1,2 refer to the circumbinary companions with i = 1, the inner, and i = 2, the outer, companions. RMS measures the root-mean-square scatter of the data around the best fit. 1-σ uncertainties have been obtained as described in the text. The last five entries are quantities of the two companions in the astrocentric coordinate system. Note that our values for a and P are somewhat larger than those of Lee et al. (2009): ab = 3.62 and ac = 5.30 AU, Pb = 3316 and Pc = 5786 days. c(cid:13) 2012 RAS, MNRAS 000, 1–??
1804.06595
1
1804
2018-04-18T08:22:29
Equation of state and optical properties of shock-compressed C:H:N:O molecular mixtures
[ "astro-ph.EP" ]
Water, ethanol, and ammonia are the key components of the mantles of Uranus and Neptune. To improve structure and evolution models and give an explanation of the magnetic fields and luminosities of the icy giants, those components need to be characterised at planetary conditions (some Mbar and a few $10^3$ K). Those conditions are typical of the Warm Dense Matter regime, which exhibits a rich phase diagram, with the coexistence of many states of matter and a large variety of chemical processes. H$_2$O, C:H:O, and C:H:N:O mixtures have been compressed up to 2.8 Mbar along the principal Hugoniot using laser-driven decaying shocks. The experiments were performed at the GEKKO XII and LULI 2000 laser facilities using standard optical diagnostics (Doppler velocimetry and pyrometry) to characterise equation of state and optical reflectivity of the shocked states. The results show that H$_2$O and the C:H:N:O mixture share the same equation of state with a density scaling, while the reflectivity behaves differently by what concerns both the onset pressures and the saturation values. The reflectivity measurement at two frequencies allows to estimate the conductivity and the complex refractive index using a Drude model.
astro-ph.EP
astro-ph
EQUATION OF STATE AND OPTICAL PROPERTIES OF SHOCK-COMPRESSED C:H:N:O MOLECULAR MIXTURES M. GUARGUAGLINI1,2, J.-A. HERNANDEZ1,2, T. OKUCHI3, P. BARROSO4, A. BENUZZI-MOUNAIX1,2, R. BOLIS1,2, E. BRAMBRINK1,2, Y. FUJIMOTO5, R. KODAMA5,6,7, M. KOENIG1,2,6, F. LEFEVRE1, K. MIYANISHI7, N. OZAKI5,7, T.SANO7, Y. UMEDA5, T. VINCI1,2, AND A. RAVASIO1,2 Abstract. Water, ethanol, and ammonia are the key components of the mantles of Uranus and Neptune. To improve structure and evolution models and give an explanation of the magnetic fields and luminosities of the icy giants, those components need to be characterised at planetary conditions (some Mbar and a few 103 K). Those conditions are typical of the Warm Dense Matter regime, which exhibits a rich phase diagram, with the coexistence of many states of matter and a large variety of chemical processes. H2O, C:H:O, and C:H:N:O mixtures have been compressed up to 2.8 Mbar along the principal Hugoniot using laser-driven decaying shocks. The experiments were performed at the GEKKO XII and LULI 2000 laser facilities using standard optical diagnostics (Doppler velocimetry and pyrometry) to characterise equation of state and optical reflectivity of the shocked states. The results show that H2O and the C:H:N:O mixture share the same equation of state with a density scaling, while the reflectivity behaves differently by what concerns both the onset pressures and the saturation values. The reflectivity measurement at two frequencies allows to estimate the conductivity and the complex refractive index using a Drude model. Introduction Composite mixtures behaviour at extreme pressures and temperatures shows intriguing chemical and physical processes, involving complex bounding scenarios. At different temperatures and densities, a variety of states ex- ist. These include combinations of many chemical species in distinct states of ions, atoms, molecules, clusters and lattices, depending on the specific conditions. Of particular interest are C:H:N:O mixtures (also called planetary ices), as they comprise the major component of the interiors of our icy giant planets Uranus and Neptune. Their structure are indeed composed by an outer layer of hydrogen and helium, an "icy" mantle made of a water (H2O) - methane (CH4) - ammonia (NH3) mixture, and possibly a rocky core [1]. As pressure and temperature increase from the outer layers towards the core, their interiors are expected to exhibit a wide range of different states embracing atomic and molecular fluids, dissociated plasmas, and superionic lattices. The complexity in describing the behaviour of these mixtures at planetary conditions (few Mbar, few 1000 K) is at the basis of the numerous lacunae in our understanding of Uranus and Neptune. Their internal structures are inferred from the observed gravitational fields, masses, internal rotation and radii. However mass distribution remains ambiguous [2]. Accurate analysis of Voyager 2 data [3, 4, 5, 6] even open the possibility for a dichotomy in their structures, indicating that the two planets could have very different interiors, despite being similar in mass and radius. Lack of precise information on transport properties of the C:H:N:O mixture is also casting serious issues in explaining Uranus and Neptune's magnetic fields [7]. Similarly, the simplified approach adopted in the ice characterisation fails in describing Uranus' low luminosity [8, 9]. Resolving this situation is even more urgent today as the discovery of exoplanets is incredibly active. Since solar planets are used as prototypes for extrasolar planets [10], the loose description of planetary ice and the resulting approximate portrait of Uranus and Neptune not only prevent the understanding of extrasolar giant planets such as GJ 436b or HAT-P-11b but also affect our capability 1LULI - CNRS, ´Ecole Polytechnique, CEA, Universit´e Paris-Saclay, route de Saclay, 91128 Palaiseau cedex, France 2Sorbonne Universit´e, UPMC Univ. Paris 06, CNRS, Laboratoire d'Utilisation des Lasers Intenses (LULI), place Jussieu, 75252 Paris cedex 05, France 3Institute for Planetary Materials, Okayama University, Misasa, Tottori 682-0193, Japan 4GEPI, Observatoire de Paris, PSL Universit´e, CNRS, 77 avenue Denfert Rochereau, 75014 Paris, France 5Graduate School of Engineering, Osaka University, Suita, Osaka 565-0871, Japan 6Open and Transdisciplinary Research Initiatives, Osaka University, Suita, Osaka 565-0871, Japan 7Institute of Laser Engineering, Osaka University, Suita, Osaka 565-0871, Japan E-mail addresses: [email protected]. to distinguish Earth-like planets candidates. As a result there is actually a great need to establish benchmarking values for the equations of states, phase diagrams, and transport properties of H2O-NH3-CH4 mixtures at Mbar pressures and temperatures of some 103 K. So far, our knowledge of water/methane/ammonia mixtures mainly re- lies on ab initio calculations [11, 12, 13, 14] since experimental data at planetary conditions [15, 16, 11] are limited, while water has been experimentally probed up to high pressures [17, 18, 19, 20, 21]. In the present work, we have compressed water and two C:H:(N):O mixtures relevant for ice giant interiors up to 2.8 Mbar through laser-driven shocks. We have measured the equation of state of the shocked sample and the optical reflectivity of the shock front using optical diagnostics (VISARs and SOP). An estimation of the electronic contribution to conductivity is given using a Drude model. Methods and experimental setup Mixtures. Liquid water/ethanol (C:H:O) and water/ethanol/ammonia (C:H:N:O) mixtures have been prepared by adding up different amounts (see the Supporting Information) of pure water, pure ethanol, and a liquid wa- ter/ammonia (28% wt.) mixture to obtain the following atomic ratios: C:H:O = 4:22:7; C:H:N:O = 4:25:1:7. The latter - synthetic Uranus [16] - reproduces the chemical composition of Uranus and Neptune's mantles, with the C:N:O abundance ratios comparable to those of the Solar System [22]. The density of the mixtures at ambient conditions was ρmix 0 = 0.885 g/cm3. Laser facilities. Experiments were performed at the GEKKO XII laser facility of the Institute of Laser Engineer- ing, Osaka University (Japan) [23] and at the LULI 2000 laser facility of the Laboratoire d'Utilisation des Lasers Intenses, ´Ecole Polytechnique (France). At GEKKO XII, we used 3 up to 9 beams (corresponding to energies on target from 120 - 440 J) at 351 nm, with a 600 µm focal spot diameter. At LULI 2000 we used 1 or 2 beams (energies on target from 200 - 500 J) at 527 nm, with a 500 µm focal spot diameter. In both cases, the laser pulse duration was 2.5 ns and phase plates were used to obtain a uniform irradiation spot. Targets. To optimise the target design and ensure there were no shock reverberations in the sample, we simulated laser-target interaction and the shock loading into the cell with the Lagrangian 1-D hydrodynamic code MULTI [24]. The equation of state of the target components were extracted from the SESAME tables [25, 26]. The table 7154 for water has been used for the mixtures. The multi-layered target cells were composed by a 10-15 µm thick CH ablator, a 40 µm thick Al shield, a 50 µm thick α−SiO2 standard, the sample (4 mm thick), and a rear α−SiO2 window (200 µm thick). We also performed some high-intensity shots at GEKKO XII with CH 50 µm / Au 3 µm / Al 5 µm / α−SiO2 20 µm / mixture 4 mm / α−SiO2 200 µm targets, the gold layer serving as X-ray shield to prevent any pre-heating of the sample. As probe lasers, we used a YAG at 532 nm at GEKKO XII and at 532 and 1064 nm at LULI 2000, with a full-width half-maximum pulse duration of ∼ 10 ns. Data analysis. Time-resolved shock velocity Us(t) has been extracted using the Neutrino software [27] from the output of two VISARs [28] (Doppler velocity interferometers, see the Supporting Information), both working at 532 nm (GEKKO XII) or one at 532 and one at 1064 nm (LULI 2000). The thermodynamic conditions (the mass density ρ, the pressure p, and the internal energy density E) reached in the mixture have been obtained from the Rankine-Hugoniot relations [29, 30] through impedance mismatching [31], using quartz as in situ standard. To span a range of thermodynamical conditions with a single shot, we employed a decaying shock technique. We determined the shock velocity at the exit from quartz and at the entrance in the mixture (U Qz , respectively) with a linear fit on Us(t) on a time window of some ns before and after the crossing of the quartz/mixture interface. Shock velocity versus fluid downstream velocity (Us-Up) Sandia Z-pinch data [32] have been used as reference for quartz. The adiabatic release of quartz on the lower-impedance mixture at the shock crossing of the interface has been modeled using a mirror Hugoniot approximation. This method agrees with the use of a quartz release model [33] within the error bars in the region where the latter can be applied. Time-resolved self emission has been measured through a streaked optical pyrometer (SOP) working at λSOP = 455 and U mix s s 2 Figure 1. Top. VISAR and SOP raw output for a GEKKO shot on C:H:N:O mixture. The three time periods indicate when the probe laser is reflected by aluminum (Al), when a reflecting shock front is propagating through the quartz layer (Qz) and the mixture. The transverse target dimension is ∼ 180 µm. Bottom. Shock velocity temporal profile from the VISAR raw output. nm. Temperature has been obtained from Planck's law T = T0/ ln(1 + AλSOP /Nc), where T0 = hc/kBλSOP , A is a calibration factor, λSOP is the emissivity of the shock front at the working wavelength and Nc is the number of counts on the SOP. To get the emissivity at 455 nm we used the reflectivity measured at 532 nm under a grey-body hypothesis: λSOP = 1 − R(532 nm). SOP calibration has been made in situ, by determining the A factor using quartz as standard (GEKKO XII) or using a calibration lamp with known emission temperature (LULI 2000, see the Supporting Information). A typical VISAR and SOP output is shown in Figure 1, together with the extracted shock velocity temporal profile. Since the shocked sample has different thermodynamical conditions with respect to the un-shocked one, its refrac- tive index changes. According to the Fresnel equations, the shock front reflects a fraction of the incident light from the probe laser. The reflectivity of the shock front has been measured with the VISARs as the ratio between the shot signal and a reference signal reflected on the aluminum/quartz interface. A VISAR-independent reflectivity measurement at 532 nm was included in the setup for some shots at LULI 2000. Since VISARs and reflectometer can measure only a relative value, a quartz reflectivity fit [34] based on previous measurements [35], has been used to calibrate the measure. 3 Results and discussion Equation of state. Figure 2 (left) shows Us-Up velocity data for pure water and the C:H:O and C:H:N:O mixtures, together with the best linear fit on previous water high pressure shock data [19]: Us = 1.35Up + 2.16 km/s. We extracted ρ-p-E thermodynamic conditions from the Us-Up relation via the Rankine-Hugoniot equations: (1a) (1b) (1c) Us ρ = ρ0 Us − Up p = p0 + ρ0UsUp E = E0 + 1 2 (p + p0) (cid:18) 1 ρ0 (cid:19) . − 1 ρ ρ-P -E results are shown in Table 1. We observed that the Us-Up relation does not significantly change between water and C:H:(N):O mixture. Therefore, the only discrepancy in their p-ρ relation along the principal Hugoniot is due to the different initial density (1.00 vs 0.88 g/cm3). The mixtures p-ρ relation is shown in Figure 2 (right) and compared with a fit on previous water data [19]. This fit has been rescaled to take into account the different initial density of water and mixtures in order to be immediately compared with mixture data. The common Us-Up relation between water and mixtures indicates that they share a similar structural behaviour. This confirms previous first-principles molecular dynamics (FPMD) simulations [11] which identify the regime we explored as an electronic conducting phase. At these temperatures, carbon-carbon and carbon-nitrogen bond life- times are predicted to be very short by first-principles calculations [13]. This prevents polymerisation and clustering, and causes the existence of an atomic fluid above 5000− 6000 K. Therefore, no structural effect of carbon and nitro- gen atoms on the Us-Up relation is expected at those conditions. These results confirm recent FPMD calculations [14] which validate the use of the linear mixing approximation when dealing with C:H:N:O mixtures at planetary conditions. Figure 2. Left. Water, C:H:O, and C:H:N:O mixture Us − Up relation along the principal Hugo- niot with a fit on previous results [19]. Right. C:H:O and C:H:N:O mixture p − ρ relation along the principal Hugoniot. The blue line is the p− ρ transposition of the Us − Up linear fit on previous water data [19]. The orange line is the same fit rescaled to take into account the initial density dif- ference between water and mixtures. The magenta lines are Uranus profiles according to water-only model [36], a model with a thermal boundary layer [37], and an icy model [14]. The temperature-pressure (T -p) relations of water and C:H:N:O mixture are shown in Figure 3 and 4, respec- tively. Figures 3 and 4 also show the predicted planetary isentropes of Uranus [36, 37, 14] and Neptune [36]. Our 4 6 8 10 12 14 16 18 20 22 24 4 6 8 10 12 14 16Us (km/s)Up (km/s)waterCHO mixtureCHNO mixtureNellis, 1997 (CHNO)fit on Knudson, 2012 0 0.5 1 1.5 2 2.5 3 3.5 4 1.8 2 2.2 2.4 2.6 2.8 3 3.2Pressure (Mbar)Density (g/cc)CHO mixtureCHNO mixturefit on Knudson, 2012rescaled fitUranus profile (Redmer, 2011)Uranus profile (Nettelmann, 2016)Uranus profile (Bethkenhagen, 2017) Figure 3. Water temperature vs pressure along the principal Hugoniot. Each color dot is a decaying shock measurement. The green-shaded area corresponds to the fit within the errors. Uranus and Neptune isentropes are from a water-only planetary model [36]. Figure 4. C:H:N:O mixture temperature vs pressure along the principal Hugoniot. Color dots are decaying shock measurements, each color corresponds to a different shot. The blue-shaded area corresponds to the fit within the errors. Violet points with error bars are previous data along the principal Hugoniot [15]. A water-only planetary model [36] is shown for Uranus and Neptune. A thermal boundary layer [37] and an icy [14] model are shown for Uranus. data have been fitted with the function T (Us) = θ0 +γU δ s , rescaled to be pressure-dependent using the p(Us) relation given by the equation of state of water. An extrapolation of our fit to lower pressures is compatible with previous gas-gun data [18]. While our data agree within the errors with recent laser shock results [20], our temperatures are higher than those given by FPMD simulations [38], although it is worth noticing that when quantum corrections 5 5000 10000 20000 50000 0.5 1 1.5 2 2.5 3Temperature (K)’Pressure (Mbar)French, 2009Kimura, 2015French, 2009 (b)Lyzenga, 1982SESAME table 7150SESAME table 7154fit on our dataUranus profile (Redmer, 2011)Neptune profile (Redmer, 2011) 5000 10000 20000 50000 0.5 1 1.5 2 2.5 3Temperature (K)’Pressure (Mbar)GEKKO shotLULI shot 1000 2000 5000 10000 20000 0 0.5 1 1.5 2 2.5 3Temperature (K)Pressure (Mbar)Radousky, 1990Uranus profile (Redmer, 2011)Neptune profile (Redmer, 2011)Uranus profile (Nettelmann, 2016)Uranus profile (Bethkenhagen, 2017) from molecular vibrations are taken into account [39] the predicted temperatures increase of (cid:39) 700 K and become more similar to our data. Our T -p results for C:H:N:O agree with a previous low-pressure experimental study of the same mixture [15]. Water and C:H:N:O temperatures on the principal Hugoniot are comparable, although the relatively high error bars on temperature make difficult to point out possible discrepancies. Hereafter we compare our T -p data with the available models of planetary interiors profiles. Most of them [36, 14] predict an adiabatic profile inside the icy giants, implying that temperatures stay relatively low (3 - 4 ·103 K) even at the highest pressures we explored (about 3 Mbar). Indeed, we compressed the sample through single shock loading, which is a process causing high entropy increase, reaching higher temperatures than those of isentropic models. Nevertheless, recent interior profile models include a thermal boundary layer [37], predicting 2-3 times higher temperatures, consistent with our data up to 1.5 Mbar. Moreover, the recent discovery of a large amount of exoplanets exhibits a wide range of structures including hot Neptunes, whose interior profiles can match the thermodynamical conditions we explored. Finally, our data are useful for the validation of FPMD simulations at extreme conditions. Figure 5. Left. Water reflectivity at 532 and 1064 nm vs pressure along the principal Hugoniot. Color dots are decaying shock measurements, each color corresponds to a different shot. Green and gold-shaded areas correspond to the fit on our data at 532 and 1064 nm within the error bars, respectively. Dashed and dotted lines correspond to DFT / Kubo-Greenwood calculations of the reflectivity [40] at 1064 or 532 nm, respectively, using the HSE (red) or PBE (blue) exchange- correlation functionals. Right. C:H:N:O mixture reflectivity at 532 and 1064 nm vs pressure along the principal Hugoniot. The blue and pink-shaded area correspond to the fit on our data at 532 and 1064 nm within the error bars, respectively. Some typical error bars for reflectivity measurements at 532 and 1064 nm have been shown. Error bars at 1064 nm are larger because of the limited number of available shots. Optical reflectivity. The optical reflectivity R of the water shock front at 532 and 1064 nm as a function of pressure is shown in Figure 5 (left). The gradual increase of reflectivity up to a saturation value along the principal Hugoniot indicates a smooth transition from an insulating to an electronically conducting ("metallic") state with the increase of density, pressure, and temperature. We performed a best fit on each R-Us relation using a Hill function R(Us) = R0 + (Rsat − R0)· U k 0 ), which is suitable to model this gradual transition. The function has been then rescaled to be pressure-dependent using an experimental water equation of state [19] to link pressure s + U k s /(U k 6 and shock velocity. The error bars of the fit are discussed in the Supporting Information and mainly depend on the calibration. According to the Hill fit on 532 nm data, the onset of reflectivity in water occurs at 1.1 − 1.2 Mbar. At these pressures the reflectivity of the shock front reaches the 10 − 20% of the saturation value, respectively. We found a saturation value of reflectivity of 24% at 532 nm and 34% at 1064 nm. Our 532 nm reflectivity data are in quantitative agreement with previous experiments [19, 20]. When compared with existing calculations [40], at 532 nm, in the low-pressure regime (P < 1.5 Mbar), our measured reflectivity is lower than the results using two different exchange-correlation functionals (HSE and PBE). At higher pressures our results are in qualitative agreement with the calculated reflectivity using the PBE functional, while the HSE one fails in providing the correct pressure-dependence, as observed by [19]. At 1064 nm we always obtained data in qualitative agreement with the calculations. The reflectivity of the C:H:N:O mixture shock front at 532 nm and 1064 nm is shown in Figure 5 (right) against the shocked sample pressure, with Hill fits on the R(Us) relations rescaled to be pressure-dependent. These are the first reflectivity data of the C:H:N:O mixture along the principal Hugoniot. There are no calculations of shock- compressed C:H:N:O reflectivity in the literature. The onset of reflectivity in C:H:N:O occurs at lower pressures Indeed, reflectivity at 532 nm reaches the 10 − 20% of the saturation value at 0.8 − 0.9 Mbar, than in water. respectively. While water reflectivity at 1064 nm is always greater than at 532 nm, this is not true for C:H:N:O at low pressures: the onset value for the 1064 nm reflectivity is about 1.2 Mbar. Starting from 1.5 Mbar, the 1064 nm reflectivity becomes greater and saturates at 41%. The crossing between the two reflectivity values is an indication of the metallisation of the sample via a gap-closure mechanism. Indeed, frequency-dependent conductivity of a semiconducting state has a maximum at a non-zero frequency. As the gap progressively closes with the increase of density and temperature, conductivity (thus reflectivity) becomes to decrease monotonically with frequency as in a free electron gas [41, 42, 43]. The fact that the onset pressure of the C:H:N:O reflectivity at 532 nm occurs at pressures around 26% lower than in water can not be fully explained by the 12% difference between their initial densities. Different dissociation mechanisms that occur in pure water and in carbon-rich mixtures could be at the origin of the mixture higher reflectivity due to the higher free electron density associated to the breaking of carbon-carbon bonds. For similar reasons, the reflectivity saturation value at 532 nm of C:H:N:O is 29%, higher than that of water (24%). Conductivity. Electrical conductivity is one of the most important parameters to understand the planetary mag- netic field generation and structure. Indeed, a dynamo effect can be sustained if magnetic induction dominates over diffusion. This is usually expressed by the requirement that the magnetic Reynolds number Rm = µ0σuL (cid:38) 100 (where σ is the electrical conductivity of the active planetary layer component and u and L are the velocity and length scale of the fluid motion inside the layer, respectively). In gas-gun experiments, the DC electrical conductivity can be directly measured using electrodes. This approach can not be applied to laser shock experiments. Instead, they would need a measurement of the complex refractive index of the shocked sample n = n+ik since, from the wave solution of the Maxwell equations, σ(ω) = 20n(ω)k(ω). In a restricted range of pressure and temperature the absorption coefficient α(ω) = 2ωk(ω)/c and the reflectivity (2) R(ω) = [n(ω) − n0(ω)]2 + k2(ω) [n(ω) + n0(ω)]2 + k2(ω) can be simultaneously measured [21]. In this case, the evaluation of the conductivity is straightforward. Never- theless, this approach is very delicate and remains restricted to few experiments and conditions. In laser shock experiments only reflectivity is usually measured. In this case, a model has to be considered in order to infer the complex refractive index. A common approach employs a Drude modelisation of optical properties, modified to account for both free and bound carriers. Even if this model is too simplistic for a well-established conductivity estimation, no first-principles calculations on these mixtures are led to date. This approach can therefore be fol- lowed to compare mixture conductivity with water data found in literature and obtained with the same method. , where nb is the contribution of the bound electrons to the refractive index. σ(ω) = σdc/(1 − iωτ ), where σdc is the direct current conductivity and τ the electron-ion scattering time. The free parameters in this model are σdc, nb, and τ . Reflectivity is generally measured at one probe laser frequency In the context of the Drude model, n =(cid:0)n2 b + iσ(ω)/0ω(cid:1)1/2 7 (usually in the green: 532 nm), requiring two of the three parameters to be fixed in an arbitrary way. As σdc is the physical quantity of interest, both τ and nb must be estimated. A reasonable choice for τ is the Ioffe-Regel limit τIR = ls/vth, which depends on the scattering length ls = 2 (3M V /4πNANF )1/3 and on the electron thermal veloc- ity vth = (kBT /m(cid:63))1/2. M is the molar mass, V the molar volume, NA the Avogadro number and NF the number of atoms in the chemical formula of the mixture; m(cid:63) = me/2 is the reduced mass in a semiconductor formalism. nb is much more delicate to be estimated over a wide range of pressures and temperatures. It is usually considered either as a constant or linearly dependent on density (Gladstone-Dale model), which is a simplistic assumption for conducting states. The simultaneous measurement of reflectivity at two frequencies in our experiments removes this difficulty, reducing the number of parameters to fix to one (τ = τIR). Following this approach, for each Hugoniot state we find the best couple (nb, σdc) which matches the two reflectivity measurements. Figure 6. Conductivity vs temperature for water (blue) and C:H:N:O mixture (red). We show our results by applying the Drude model to direct reflectivity measurements (inverted triangles, with error bars) and to the Hill fit on our reflectivity datasets (continuous lines). Temperatures from Chau [11] have been corrected transposing the correction made by Millot [21] on another dataset of the same author [44]. In Figure 6 we show the temperature-dependent DC conductivity of water and C:H:N:O mixture. After a rapid arise from 4000 to 10000 K, conductivity quasi-saturates, following the reflectivity behaviour. In this region, at 15000 K ((cid:39) 2 Mbar), conductivity values are ∼ 2.1 and ∼ 4.1 · 102 S/cm for water and C:H:N:O, respectively. We notice that the C:H:N:O mixture conductivity is higher than water. A different behaviour was observed in multiple shock experiments. As already underlined, in this conditions the electronic contribution dominates over the ionic one. This situation is different from previous experiments [16, 11] where the main contribution was ionic. These data highlight that the use of water as a representative of planetary ices can be a too simplified picture for dynamo modelisation. 8 0.001 0.01 0.1 1 10 100 1000 10000 0.5 1 2 5 10 20Conductivity (S/cm)Temperature (1000 K)Waterour dataour data (Hill fit)total - Yuknavech, 1964total - Hamann, 1966total - Mitchell, 1982electronic - Celliers, 2004total - French, 2010electronic - French, 2010electronic - Millot, 2018 0.001 0.01 0.1 1 10 100 1000 10000 0.5 1 2 5 10 20Conductivity (S/cm)Temperature (1000 K)CHNO mixtureour dataour data (Hill fit)total - Nellis, 1997total - Chau, 2011 Refractive index. The combined use of the reflectivity at two different frequencies also allows us to infere a measurement of the complex refractive index of the shocked sample. From the couple (nb, σdc), extracted as explained in Subsection , we can compute the complex refractive index along the Hugoniot: (cid:18) (cid:18) n2 b + i σdc 0ω(1 − iωτ ) (cid:19)1/2 (cid:19)1/2 . (3a) (3b) n(ω) = n(2ω) = σdc n2 b + i 20ω(1 − 2iωτ ) The real and imaginary part of the refractive indices of water and C:H:N:O mixture at both laser frequencies are shown in Figure 7. Low-density real refractive index values are comparable to the results in the literature [45, 46, 47, 48]. At densities around 2.6 g/cm3, the water real refractive index starts to increase from values com- parable to those given by the Gladstone-Dale model [49] to a saturation value of around 3.5 − 4. This value is very similar to previous results [49, 46], although they found it at around 2.4 g/cm3, in an opaque regime where reflec- tivity could not be measured. It has been recently noticed [21] that this is in contrast with the Fresnel reflectivity estimation obtained with n = 3.5. Our data are not affected by this inconsistency, since we find n (cid:39) 3.5 at 2.8 g/cm3, where water reflectivity is ∼ 20% . The increase of the real and imaginary part of the refractive index are a marker of the transition to a metallic state. Our data show that this transition takes place between 2.5 and 2.8 g/cm3. For C:H:N:O, the sudden increase in both the real and imaginary refractive indices takes place around 2.5 g/cm3. Figure 7. Left. Real (n) and imaginary (k) part of the shock-compressed water refractive index vs density. The dashed line is a Gladstone-Dale model for n shown in literature [49]. Right. Real and imaginary part of the shock-compressed C:H:N:O mixture refractive index vs density. Conclusions We studied the behaviour of shock-compressed water and C:H:N:O mixtures at extreme conditions in the Warm Dense Matter regime, reaching pressures up to 2.8 Mbar and temperatures of 24000 K. We obtained ρ-p-E equation of state, temperature, optical reflectivity, and electronic contribution to the electrical conductivity of pure H2O and C:H:N:O mixtures along their principal Hugoniot. We found that the only difference in the ρ-p relations of water and C:H:N:O can be completely explained by the difference in the initial densities. Their T -p relations are comparable, although possible small discrepancies could not be distinguished. 9 0 1 2 3 4 5 1.6 1.8 2 2.2 2.4 2.6 2.8 3 3.2waterRefractive indexDensity (g/cc)n at 1064 nmk at 1064 nmn at 532 nmk at 532 nmGladstone-Dale modelZeldovich, 1961Henry, 2003Dewaele, 2003Zha, 2007 0 1 2 3 4 5 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8CHNODensity (g/cc)n at 1064 nmk at 1064 nmn at 532 nmk at 532 nm The similarity between the equations of state of water and C:H:N:O confirms the validity of the Linear Mixing Ap- proximation at planetary conditions [14]. Moreover, the studied Hugoniot states are consistent with the existence of an atomic fluid above 5000 − 6000 K as recently expected by first-principles calculations [13]. The reflectivity behaviour of water and C:H:N:O mixture are different. The reflectivity onset for C:H:N:O is at 0.8 − 0.9 Mbar (R532 nm = 2.9 − 5.8%), slightly lower than water which is found at 1.1 − 1.2 Mbar (R532 nm = 2.4 − 4.8%). At 1.5 Mbar, C:H:N:O reflectivity at 1064 nm gets higher than at 532 nm showing a strong metallic behaviour. The reflectivity saturation values are higher for C:H:N:O than for water (29% against 24% at 532 nm, 41% against 34% at 1064 nm). Using the dual reflectivity measurement, conductivity and complex refractive index of shocked water and C:H:N:O mixture are obtained through a modified Drude model. Saturation values for conductivities are ∼ 2600 and ∼ 4000 S/cm for water and C:H:N:O, respectively. Our results suggest that, in a mantle composed by C:H:N:O mixtures, planetary dynamo could be sustained differently than expected if water is assumed as unique component. Future experimental work should consider high pressure off-Hugoniot states to enlarge the studied scenario and explore thermodynamic conditions more relevant to planetary interiors. shot # ρ (g/cm3) p (Mbar) E − E0 (kJ/g) GK-680 L1-19 GK-687 GK-706 GK-725 GK-744 GK-753 L1-29 L2-82 L2-86 GK-694 GK-696 GK-712 GK-715 GK-745 L2-102 L2-130 2.74 ± 0.18 3.00 ± 0.15 Pure water 1.23 ± 0.05 2.73 ± 0.06 C:H:O mixture C:H:N:O mixture 2.48 ± 0.15 2.24 ± 0.20 2.65 ± 0.13 2.32 ± 0.14 2.38 ± 0.22 2.86 ± 0.19 2.37 ± 0.16 2.31 ± 0.19 2.56 ± 0.19 2.78 ± 0.19 2.79 ± 0.23 2.14 ± 0.19 2.18 ± 0.19 2.65 ± 0.26 2.44 ± 0.30 1.42 ± 0.04 0.74 ± 0.03 2.55 ± 0.06 1.16 ± 0.04 0.94 ± 0.04 2.67 ± 0.07 2.79 ± 0.11 1.84 ± 0.08 1.73 ± 0.06 1.58 ± 0.05 2.60 ± 0.08 0.60 ± 0.03 0.67 ± 0.03 1.57 ± 0.07 2.08 ± 0.10 39.2 ± 2.7 91.3 ± 3.7 51.9 ± 2.7 25.5 ± 2.0 96.8 ± 3.8 41.1 ± 2.6 33.5 ± 2.7 104.8 ± 5.1 99.6 ± 6.8 64.8 ± 5.3 63.6 ± 4.0 60.4 ± 3.4 99.4 ± 5.9 19.5 ± 2.1 22.3 ± 2.1 56.7 ± 4.8 74.4 ± 7.8 Table 1. Experimental data on pure water, C:H:O mixture, and C:H:N:O mixture. The shot number prefix GK, L1, and L2 identify the campaigns at GEKKO XII in January 2016, at LULI 2000 in February 2017, and at LULI 2000 in September 2017, respectively. References [1] Tristan Guillot. Interiors of giant planets inside and outside the solar system. Science, 286(5437):72–77, 1999. [2] M. Podolak and R. Helled. What do we really know about Uranus and Neptune? Astrophysical Journal Letters, 759:L32(2), 2012. [3] G.F. Lindal, J.R. Lyons, D.N. Sweetnam, V.R. Eshleman, D.P. Hinson, and G.L. Tyler. The atmosphere of Uranus: results of radio occultation measurements with Voyager 2. Journal of Physical Research, 92(A13), 1987. 10 [4] J. D. Anderson, J. K. Campbell, R. A. Jacobson, D. N. Sweetnam, and A. H. Taylor. Radio science with Voyager 2 at Uranus - Results on masses and densities of the planet and five principal satellites. Journal of Geophysical Research, 92(A13):14877–14883, 1987. [5] G. L. Tyler, D. N. Sweetnam, J. D. Anderson, S. E. Borutzki, J. K. Campbell, E. R. Kursinski, G. S. Levy, G. F. Lindal, J. R. Lyons, and G. E. Wood. Voyager radio science observations of Neptune and Triton. Science, 246:1466–1473, 1989. [6] G. F. Lindal. The atmosphere of Neptune - an analysis of radio occultation data acquired with Voyager 2. Astrophysical Journal, 103(3):967–982, 1992. [7] D.J. Stevenson. Planetary magnetic fields: Achievements and prospects. Space Science Reviews, 152:651–664, 2010. [8] J. C. Pearl, B. J. Conrath, R. A. Hanel, and J. A. Pirraglia. The albedo, effective temperature, and energy balance of Uranus, as determined from Voyager IRIS data. Icarus, 84:12–28, March 1990. [9] J. C. Pearl and B. J. Conrath. The albedo, effective temperature, and energy balance of Neptune, as determined from Voyager data. Journal of Geophysical Research Supplement, 96:18, October 1991. [10] D. C. Swift, J. H. Eggert, D. G. Hicks, S. Hamel, K. Caspersen, E. Schwegler, G. W. Collins, N. Nettelmann, and G. J. Ackland. Mass-radius relationships for exoplanets. The Astrophysical Journal, 744(1):59, 2012. [11] R. Chau, S. Hamel, and W.J. Nellis. Chemical processes in the deep interior of Uranus. Nature Communications, 2(203), 2011. [12] M. Lee and S. Scandolo. Mixtures of planetary ices at extreme conditions. Nature communications, 2:185, 2011. [13] Edmund R. Meyer, Christopher Ticknor, Mandy Bethkenhagen, Sebastien Hamel, Ronald Redmer, Joel D. Kress, and Lee A. Collins. Bonding and structure in dense multi-component molecular mixtures. The Journal of Chemical Physics, 143(16):164513, 2015. [14] M. Bethkenhagen, E. R. Meyer, S. Hamel, N. Nettelmann, M. French, L. Scheibe, C. Ticknor, L. A. Collins, J. D. Kress, J. J. Fortney, and R. Redmer. Planetary ices and the linear mixing approximation. The Astrophysical Journal, 848(1):67, 2017. [15] H.B. Radousky, A.C. Mitchell, and W.J. Nellis. Shock temperature measurements of planetary ices: NH3, CH4, and synthetic Uranus. Journal of Chemical Physics, 93(11):8235–8239, 1990. [16] W.J. Nellis, N.C. Holmes, A.C. Mitchell, D.C. Hamilton, and M. Nicol. Equation of state and electrical conductivity of synthetic Uranus, a mixture of water, ammonia, and isopropanol, at shock pressures up to 200 GPa. J. Chem. Phys., 107(21):9096–9100, 1997. [17] A.C. Mitchell and W.J. Nellis. Equation of state and electrical conductivity of water and ammonia shocked to the 100 GPa (1 Mbar) pressure range. Journal of Chemical Physics, 76(12):6273–6281, 1982. [18] G. A. Lyzenga, T. J. Ahrens, W. J. Nellis, and A. C. Mitchell. The temperature of shock-compressed water. Journal of Chemical Physics, 76:6282–6286, June 1982. [19] M.D. Knudson, M. P. Desjarlais, R.W. Lemke, T. R. Mattsson, M. French, N. Nettelmann, and R. Redmer. Probing the interiors of the ice giants: Shock compression of water to 700 GPa and 3.8 g/cm3. Physical Review Letters, 108:091102, Feb 2012. [20] T. Kimura, N. Ozaki, T. Sano, T. Okuchi, T. Sano, K. Shimizu, K. Miyanishi, T. Terai, T. Kakeshita, Y. Sakawa, and R. Kodama. P-ρ-T measurements of H2O up to 260 GPa under laser-driven shock loading. Journal of Chemical Physics, 142(16):164504, 2015. [21] Marius Millot, Sebastien Hamel, J. Ryan Rygg, Peter M. Celliers, Gilbert W. Collins, Federica Coppari, Dayne E. Fratanduono, Raymond Jeanloz, Damian C. Swift, and Jon H. Eggert. Experimental evidence for superionic water ice using shock compression. Nature Physics, 14:297–302, 2018. [22] A. G. W. Cameron. Abundances of the Elements in the Solar System. Space Science Reviews, 15:121–146, September 1973. [23] N. Ozaki, K. A. Tanaka, T. Ono, K. Shigemori, M. Nakai, H. Azechi, T. Yamanaka, K. Wakabayashi, M. Yoshida, H. Nagao, and K. Kondo. GEKKO/HIPER-driven shock waves and equation-of-state measurements at ultrahigh pressures. Physics of Plasmas, 11:1600–1608, April 2004. [24] R. Ramis, R. Schmalz, and J. Meyer-Ter-Vehn. MULTI: A computer code for one-dimensional multigroup radiation hydrodynamics. Computer Physics Communications, (3):475 – 505, 1988. [25] S.P. Lyon and J.D. Johnson. Los Alamos technical report, la-ur-92-3407 edition, 1992. [26] J.D. Johnson. National Technical Information Service Document, 1994. [27] A. Flacco and T. Vinci. Neutrino: expandable and full A light, featured image analysis tool for research. https://github.com/NeutrinoToolkit/Neutrino, 2011. [28] D.H. Dolan. Foundations of VISAR analysis, 2006. [29] W.J. Macquorn Rankine. On the thermodynamic theory of waves of finite longitudinal disturbance. Philosophical Transactions of the Royal Society of London, 160:277–288, 1870. [30] H. Hugoniot. M´emoire sur la propagation des mouvements dans les corps et sp´ecialement dans les gaz parfaits. Journal de l'Ecole Polytechnique, 57:3–97. [31] J.W. Forbes. Shock Wave Compression of Condensed Matter. Springer, 2012. [32] M. D. Knudson and M. P. Desjarlais. Shock Compression of Quartz to 1.6 TPa: Redefining a Pressure Standard. Physical Review Letters, 103(22):225501, November 2009. [33] M. D. Knudson and M. P. Desjarlais. Adiabatic release measurements in α-quartz between 300 and 1200 GPa: Characterization of α-quartz as a shock standard in the multimegabar regime. Physical Review B, 88(18):184107, November 2013. [34] M. Millot, N. Dubrovinskaia, A. Cernok, S. Blaha, L. Dubrovinsky, D. G. Braun, P. M. Celliers, G. W. Collins, J. H. Eggert, and R. Jeanloz. Shock compression of stishovite and melting of silica at planetary interior conditions. Science, 347:418–420, 2015. [35] D. G. Hicks, T. R. Boehly, J. H. Eggert, J. E. Miller, P. M. Celliers, and G. W. Collins. Dissociation of Liquid Silica at High Pressures and Temperatures. Physical Review Letters, 97(2):025502, July 2006. 11 [36] R. Redmer, T.R. Mattsson, N. Nettelmann, and M. French. The phase diagram of water and the magnetic fields of Uranus and Neptune. Icarus, 211(1):798 – 803, 2011. [37] N. Nettelmann, K. Wang, J.J. Fortney, S. Hamel, S. Yellamilli, M. Bethkenhagen, and R. Redmer. Uranus evolution models with simple thermal boundary layers. Icarus, 275:107 – 116, 2016. [38] M. French, T.R. Mattsson, N. Nettelmann, and R. Redmer. Equation of state and phase diagram of water at ultrahigh pressures as in planetary interiors. Physical Review B, 79:054107, Feb 2009. [39] M. French and R. Redmer. Estimating the quantum effects from molecular vibrations of water under high pressures and tempera- tures. Journal of Physics Condensed Matter, 21:375101, September 2009. [40] M. French and R. Redmer. Optical properties of water at high temperature. Physics of Plasmas, 18(4):043301, 2011. [41] Yann Laudernet, Jean Cl´erouin, and St´ephane Mazevet. Ab initio simulations of the electrical and optical properties of shock- compressed SiO2. Phys. Rev. B, 70:165108, Oct 2004. [42] Jean Clerouin, Patrick Renaudin, Yann Laudernet, Pierre Noiret, and Michael P. Desjarlais. Electrical conductivity and equation- of-state study of warm dense copper: Measurements and quantum molecular dynamics calculations. Phys. Rev. B, 71:064203, Feb 2005. [43] T. Qi, M. Millot, R. G. Kraus, S. Root, and S. Hamel. Optical and transport properties of dense liquid silica. Physics of Plasmas, 22(6):062706, June 2015. [44] R. Chau, A. C. Mitchell, R. W. Minich, and W. J. Nellis. Electrical conductivity of water compressed dynamically to pressures of 70 - 180 GPa (0.7 - 1.8 Mbar). The Journal of Chemical Physics, 114(3):1361–1365, 2001. [45] Y. B. Zel'Dovich, S. B. Kormer, M. V. Sinitsyn, and K. B. Yushko. A Study of the Optical Properties of Transparent Materials under High Pressure. Soviet Physics Doklady, 6:494, December 1961. [46] E. Henry. ´Equation d'´etat et m´etallisation de l'eau comprim´ee par choc laser. PhD thesis, ´Ecole Polytechnique, 2003. [47] A. Dewaele, J. H. Eggert, P. Loubeyre, and R. Le Toullec. Measurement of refractive index and equation of state in dense He, H2, H2O, and Ne under high pressure in a diamond anvil cell. Phys. Rev. B, 67:094112, Mar 2003. [48] Chang-Sheng Zha, Russell J. Hemley, Stephen A. Gramsch, Ho-Kwang Mao, and William A. Bassett. Optical study of H2O ice to 120 GPa: Dielectric function, molecular polarizability, and equation of state. The Journal of Chemical Physics, 126(7):074506, 2007. [49] D. Batani, K. Jakubowska, A. Benuzzi-Mounaix, C. Cavazzoni, C. Danson, T. Hall, M. Kimpel, D. Neely, J. Pasley, M. Rabec Le Gloahec, and B. Telaro. Refraction index of shock compressed water in the megabar pressure range. Europhysics Letters, 112(3):36001, 2015. Acknowledgements We want to thank the GEKKO XII and LULI 2000 laser and support teams. We are grateful to Ronald Redmer, Martin French, and Mandy Bethkenhagen for the useful discussions. This research was supported by a French ANR grant to the POMPEI project (ANR-16-CE31-0008), the JSPS core-to-core program on International Alliance for Material Science in Extreme States with High Power Laser and XFEL, and the International Joint Research Promotion Program at the Osaka University. This work has taken advantage of the MECMATPLA international collaboration. Author contributions statement T.O., A.B.-M., R.K., N.O., and A.R.. conceived the project. M.G., J.-A.H., T.O., P.B., A.B.-M., R.B., E.B., Y.F., F.L., K.M., N.O., Y.U., T.V., and A.R. conducted the experiments. M.G. analysed the results. M.G., J.-A.H, and A.R. wrote the paper. All authors reviewed the manuscript. 12 Additional information Supplementary infomation for this article is available. The authors declare no competing interests. The data that support the plots within this paper and other findings of this study are available from the corresponding author upon reasonable request. 13
1906.05291
1
1906
2019-06-12T18:00:01
Halo Meteors
[ "astro-ph.EP" ]
The stellar halo contains some of the oldest stars in the Milky Way galaxy and in the universe. The detections of `Oumuamua, CNEOS 2014-01-08, and interstellar dust serve to calibrate the production rate of interstellar objects. We study the feasibility of a search for interstellar meteors with origins in the stellar halo. We find the mean heliocentric impact speed for halo meteors to be $\sim 270 \mathrm{\;km\;s^{-1}}$, and the standard deviation is $\sim 90 \mathrm{\;km\;s^{-1}}$, making the population kinematically distinct from all other meteors, which are an order-of-magnitude slower. We explore the expected abundance of halo meteors, finding that a network of all-sky cameras covering all land on Earth can take spectra and determine the orbits of a few hundred halo meteors larger than a few mm per year. The compositions of halo meteors would provide information on the characteristics of planetary system formation for the oldest stars. In addition, one could place tight constraints on baryonic dark matter objects of low masses.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 3 (2019) Preprint 14 June 2019 Compiled using MNRAS LATEX style file v3.0 Halo Meteors Amir Siraj,1(cid:63) Abraham Loeb,1† 1Department of Astronomy, Harvard University, 60 Garden Street, Cambridge, MA 02138, USA Accepted XXX. Received YYY; in original form ZZZ ABSTRACT The stellar halo contains some of the oldest stars in the Milky Way galaxy and in the universe. The detections of 'Oumuamua, CNEOS 2014-01-08, and interstellar dust serve to calibrate the production rate of interstellar objects. We study the feasibility of a search for interstellar meteors with origins in the stellar halo. We find the mean heliocentric impact speed for halo meteors to be ∼ 270 km s−1, and the standard de- viation is ∼ 90 km s−1, making the population kinematically distinct from all other meteors, which are an order-of-magnitude slower. We explore the expected abundance of halo meteors, finding that a network of all-sky cameras covering all land on Earth can take spectra and determine the orbits of a few hundred halo meteors larger than a few mm per year. The compositions of halo meteors would provide information on the characteristics of planetary system formation for the oldest stars. In addition, one could place tight constraints on baryonic dark matter objects of low masses. Key words: Galaxy: halo -- meteorites, meteors, meteoroids -- planetary systems -- minor planets, asteroids: general -- comets: general 1 INTRODUCTION 'Oumuamua was the first interstellar object detected in the Solar System by Pan-STAARS (Meech et al. 2017; Micheli et al. 2018). Several follow-up studies of 'Oumuamua were conducted to better understand its origin and composition (Bannister et al. 2017; Gaidos et al. 2017; Jewitt et al. 2017; Mamajek 2017; Ye et al. 2017; Bolin et al. 2017; Fitzsimmons et al. 2018; Trilling et al. 2018; Bialy & Loeb 2018; Hoang et al. 2018; Siraj & Loeb 2019a,b; Seligman et al. 2019). There is significant evidence for previous detections of interstellar meteors (Baggaley et al. 1993; Hajdukova 1994; Taylor et al. 1996; Baggaley 2000; Mathews et al. 1998; Meisel et al. 2002a,b; Weryk & Brown 2004; Afanasiev et al. 2007; Musci et al. 2012; Engelhardt et al. 2017; Hajdukova et al. 2018), including the meter-size CNEOS 2014-01-08 meteor (Siraj & Loeb 2019c). Spectroscopy of gaseous debris from interstellar mete- ors as they burn up in the Earth's atmosphere could reveal their composition, and a worldwide network of all-sky cam- eras would allow for the detection and analysis of interstellar meteors at the centimeter-scale (Siraj & Loeb 2019d). In this Letter, we explore the theoretical population of interstellar meteors originating from the stellar halo, which contains the oldest stars in the Milky Way galaxy and in the universe (Helmi 2008; Frebel et al. 2007). Studying the (cid:63) [email protected][email protected] c(cid:13) 2019 The Authors composition of halo star ejecta could help reveal the nature of primordial planetary system formation. We discuss the kinematics of halo meteors, their abundance and expected detection rates, and what we could learn from analyzing their composition. 2 KINEMATICS φ ≈ 238 km s−1 (Reid & Dame 2016; Bland- We assume that the velocity distribution of halo interstellar objects follows that of their parent stars. The mean rotation speed of the Local Standard of Rest (LSR) with respect to the halo is v(cid:12) Hawthorn & Gerhard 2016). The Sun's velocity components with respect to the LSR, in right-handed Galactic coordi- W) = (10 ± 1, 11 ± 2, 7 ± 5) km s−1 (Bland- nates, are (v(cid:12) Hawthorn & Gerhard 2016). The mean rotation of the local φ ≈ 40 km s−1, halo population with respect to the LSR is ¯vh and the spherical velocity ellipsoid for the local halo pop- ulation is defined by the velocity dispersion, (σh φ) = (141 ± 5, 75 ± 5, 85 ± 5) km s−1 (Bland-Hawthorn & Gerhard 2016). θ, σh U , v(cid:12) V , v(cid:12) r , σh We use the following Monte Carlo method to determine the heliocentric impact velocities of halo interstellar objects. First, we draw randomly from the Gaussian distributions de- scribed by the spherical velocity ellipsoid for the local halo distributions and determine the velocity components of a random halo interstellar object with respect to the LSR. We then subtract the Sun's velocity components relative to the 2 A. Siraj and A. Loeb Figure 1. Distribution of heliocentric impact velocities of halo meteors, with U, V, and W, components shown in blue, green, and red, respectively, and total speed displayed in black. Cutoffs are at ±3σ, where σ is the standard deviation. V , ¯vh U , σh LSR. Finally, assuming an isotropic distribution, we add the kinetic energy from the change in gravitational potential. We find the heliocentric impact velocity distriution to have mean W) = (−10,−210,−7) km s−1, with velocity velocities (¯vh U , ¯vh W) = (140, 90, 80) km s−1, shown in Fig- dispersions (σh V , σh ure 1. The mean heliocentric impact speed is ∼ 270 km s−1, and the standard deviation is ∼ 90 km s−1, making halo me- teors kinematically distinct from all other (Solar System or interstellar) meteors which have characteristic impact speeds that are an order-of-magnitude smaller. 3 ABUNDANCE & DETECTABILITY Approximately 1 in every 103 local stars, and therefore inter- stellar meteors, originate from the Milky Way halo (Helmi 2008). On average, halo meteors travel ∼ 5 times faster than disk meteors. We therefore estimate the impact rates for halo meteors to be a factor of ∼ 200 lower than those for typical interstellar meteors of the same size (Siraj & Loeb 2019c). All-sky camera systems such as AMOS can obtain spec- tra for typical meteors down to a size scale of ∼ 1cm, and considering that halo interstellar meteor impact speeds are approximately ten times higher than typical meteors, we es- timate that, with an adequate frame rate, a system with the sensitivity of AMOS could obtain spectra for halo meteors down to a size scale of ∼ 2mm (Toth et al. 2015). Figure 2 shows expected Earth impact rates as a func- tion of size for halo meteors, assuming a similar size distri- bution for ejected mass as disk stars (Siraj & Loeb 2019c). We expect 2mm halo meteors to impact the Earth at a rate of 104 yr−1. A network of all-sky cameras covering all land on Earth, with the sensitivity of AMOS and an adequate frame rate to determine the orbits of ∼ 300 km s−1 impactors, are expected to detect and take spectra for nearly 103 halo mete- ors per year. We would not expect to obtain physical samples of halo meteors, as meteors larger than 10cm are expected to impact the Earth once every few hundred years. Finally, the Figure 2. Range of expected earth impact rates of halo meteors as a function of size, based on the q = 3.41 ± 0.17 estimate (Siraj & Loeb 2019c). ratio of disk to halo meteors will inform our understanding of primoridal planetary system formation. 4 IMPLICATIONS 4.1 Early Planetary Systems There are three major phases of interstellar object produc- tion during a star's lifetime (Pfalzner & Bannister 2019). The first is the shedding of icy planetesimals from the outer disk by interactions with neighboring stars (Pfalzner et al. 2015; Hands et al. 2019). The second phase involves the ejec- tion of many of the remaining planetesimals (rocky and icy), due to close encounters with giant planets during planet mi- gration (Duncan et al. 1987; Charnoz & Morbidelli 2003; Raymond et al. 2018). The final phase is the ejection of the Oort cloud as a star transitions to a white dwarf (Veras et al. 2011, 2014; Do et al. 2018; Moro-Martin 2019). While icy planetesimals are expected to be ejected across a star's lifetime, rocky planetesimals are thought to be ejected pri- marily during the gas clearing and planet migration phase, at planetary system ages 100 - 700 Myr (Pfalzner & Bannis- ter 2019). The stellar halo is home to the oldest stars in the Milky Way and in the universe (Helmi 2008; Frebel et al. 2007). Studying the composition of their ejecta can therefore re- veal the formation history of some of the first planets, as well as the enrichment history of the early universe. For instance, the observed ratio of icy to rocky planetesimals can help constrain models of planetary system formation, the chemical composition of rocky planetesimals could re- veal unknown details about the gas-clearing and planet mi- gration phases, and the ratios of volatiles in icy planetesi- mals could inform the prospects for life in early planetary systems. Compositional information for halo meteors could also help constrain the origins of exo-Oort clouds, as well as chemical diversity among early planetary systems. Further- more, analysis of halo meteors could test the possibility that carbon-enhanced metal-poor poor stars could create carbon planets (Mashian & Loeb 2016). MNRAS 000, 1 -- 3 (2019) 1E-09 11E+09 1E+18 0.00000010.00010.11001018101019-710101010-9-4-12Earth Impact Rate [yrRadius [m]]-1 4.2 Constraints on Baryonic Dark Matter Mass Hajdukova M., Jr., 1994, Astronomy and Astrophysics 288(1), Halo Meteors 3 330 Hajdukova M., Sterken, V., Wiegert, P., 2018, European Plane- tary Science Congress 12 Hands T. O., et al., 2019, MNRAS Helmi, A., 2018, Astron. Astrophys. Rev., 15, 145 Hoang T., Loeb A., Lazarian A., Cho J., 2018, The Astrophysical Journal, 860(1), 42 Jewitt D., Luu J., Rajagopal J., Kotulla R., Ridgway S., Liu W., Augusteijn T., 2017, The Astrophysical Journal, 850, L36 Mamajek E., 2017, Research Notes of the AAS, 1, 21 Mathews, D. J., Meisel, D. D., Janches, D., Getman, S. V., Zhou, Q.-H., 1998, Meteoroids 1998 (Proc. Int. Conf.), ed. W. J. Baggaley & V. Porubcan (Bratislava: Astronomical Institute of the Slovak Academy of Sciences), 79 Mashian N., Loeb A., 2016, MNRAS, 460(3), 2482 Meech K. J., et al., 2017, Nature, 552, 378 Meisel D. D., Janches D., Mathews J. D., 2002a, The Astrophys- ical Journal, 567, 323 Meisel D. D., Janches D., Mathews J. D., 2002b, The Astrophys- ical Journal, 567, 323 Micheli M., et al., 2018, Nature, 559, 223 Monroy-Rodriguez M. A., Allen C., 2019, The Astrophysical Jour- nal, 790, 159 Moro-Martin A., et al., 2019, The Astrophysical Journal, 157, 86 Musci R., et al., 2012, The Astrophysical Journal, 745, 161 Pfalzner S., et al., 2015, Phys S, 794, 147 Pfalzner S., Bannister M. T., 2019, The Astrophysical Journal, 874, L34 Reid M. J., Dame T. M., 2016, The Astrophysical Journal, 832, 159 Raymond S. N., et al., 2018, MNRAS, 476, 3031 Seligman D., Laughlin G., Batygin K., 2019, The Astrophysical Journal Letters, Siraj A. & Loeb A., 2019, The Astrophysical Journal, 872(1), L10 Siraj A. & Loeb A., 2019, Research Notes of the American As- tronomical Society, 3(1), 15 Siraj A. & Loeb A., 2019, submitted to The Astrophysical Journal Letters Siraj A., Loeb A., 2019, Probing Extrasolar Planetary Systems with Interstellar Meteors. (arXiv:1906.03270) Taylor A. D., Baggaley W. J., Steel D. I., 2018, Nature, 380, 323 Toth J., et al., 2015, Planetary & Space Science, 118, 102 Trilling, D., et al., 2018, The Astronomical Journal, 156, 261. Veras D., et al., 2011, MNRAS, 417, 2104 Veras D., et al., 2014, MNRAS, 437, 1127 Weryk R. J., Brown P., 2004, Icarus, 95, 221. Ye Q.-Z., Zhang Q., Kelley M. S. P., Brown P. G., 2017, The Astrophysical Journal, 851, L5 This paper has been typeset from a TEX/LATEX file prepared by the author. Massive compact halo objects (MACHOs) were theorized to partially explain dark matter in the Galactic halo. Con- straints on MACHOs from microlensing, wide binaries, and disk kinematics constrain abundances of MACHOs at masses m ≥ 10−7M(cid:12) (Alcock et al. 1998; Monroy-Rodriguez & Allen 2014; Brandt 2016). Halo meteors could provide tight con- straints on the baryonic dark matter mass fraction for MA- CHOs at masses m ≤ 10−24M(cid:12). For instance, integrating over the inferred interstellar meteor size distribution (Siraj & Loeb 2019c) implies total mass m ∼ 10−5M(cid:12), a few Earth masses, per star of interstellar baryonic material at masses m ≤ 10−24M(cid:12). Applying a simi- lar approach to the observed size distribution of halo mete- ors will allow for tight constraints on baryonic dark matter masss for MACHOs at asteroidal sizes. 5 DISCUSSION We analyzed the kinematics of halo meteors, showing their mean heliocentric impact speed to be ∼ 270 km s−1, with a standard deviation of ∼ 90 km s−1. We then explored their abundance and detectability, finding that a worldwide net- work of all-sky cameras could take spectra and determine orbits for hundreds of d ≥ 2mm halo meteors per year. Fi- nally, we explored the implications of halo meteors on early planetary systems and on constraining baryonic dark mat- ter. ACKNOWLEDGEMENTS This work was supported in part by a grant from the Break- through Prize Foundation. REFERENCES Afanasiev V. L., Kalenichenko V. V., Karachentsev I. D., 2007, Astrophysical Bulletin, 62(4), 319 Alcock C., et al., 1998, The Astrophysical Journal, 499, L99 Baggaley W. J., Taylor, D.A. & Steel, I.D. 1993, Meteoroids and their Parent Bodies, Proc. Int. Astron. Symp., 53 Baggaley W. J., 2000, Journal of Geophysical Research, 105(A5), 10353 Bannister M. T., et al., 2017, The Astrophysical Journal, 851, L38 Bialy S., Loeb A., 2018, The Astrophysical Journal, 868, L1 Bland-Hawthorn J., Gerhard O., 2016, Annual Review of Astron- omy and Astrophysics, 54, 529 Bolin B. T., et al., 2017, The Astrophysical Journal Letters, Vol- ume 852, Issue 1, article id. L2, 10 pp. (2018)., 852 Brandt T. D., 2016, The Astrophysical Journal, 824, L31 Charnoz S., Morbidelli A., 2003, Icarus, 166, 141 Do, A., Tucker, M. A., Tonry, J. 2018, The Astrophysical Journal, 855, L10 Duncan, M., Quinn, T., Tremaine, S. 1987, The Astronomical Journal, 94, 1330 Engelhardt T., et al., 2017, The Astronomical Journal, 153, 133 Fitzsimmons A., et al., 2018, Nature Astronomy, 2, 133 Frebel A., et al., 2007, The Astrophysical Journal, 660, L117 Gaidos E., Williams J., Kraus A., 2017, Research Notes of the AAS, 1, 13 MNRAS 000, 1 -- 3 (2019)
1112.3305
2
1112
2012-07-04T23:08:47
Planetary Rings
[ "astro-ph.EP", "astro-ph.CO", "astro-ph.GA", "astro-ph.HE", "astro-ph.SR" ]
Planetary rings are the only nearby astrophysical disks, and the only disks that have been investigated by spacecraft. Although there are significant differences between rings and other disks, chiefly the large planet/ring mass ratio that greatly enhances the flatness of rings (aspect ratios as small as 1e-7), understanding of disks in general can be enhanced by understanding the dynamical processes observed at close-range and in real-time in planetary rings. We review the known ring systems of the four giant planets, as well as the prospects for ring systems yet to be discovered. We then review planetary rings by type. The main rings of Saturn comprise our system's only dense broad disk and host many phenomena of general application to disks including spiral waves, gap formation, self-gravity wakes, viscous overstability and normal modes, impact clouds, and orbital evolution of embedded moons. Dense narrow rings are the primary natural laboratory for understanding shepherding and self-stability. Narrow dusty rings, likely generated by embedded source bodies, are surprisingly found to sport azimuthally-confined arcs. Finally, every known ring system includes a substantial component of diffuse dusty rings. Planetary rings have shown themselves to be useful as detectors of planetary processes around them, including the planetary magnetic field and interplanetary impactors as well as the gravity of nearby perturbing moons. Experimental rings science has made great progress in recent decades, especially numerical simulations of self-gravity wakes and other processes but also laboratory investigations of coefficient of restitution and spectroscopic ground truth. The age of self-sustained ring systems is a matter of debate; formation scenarios are most plausible in the context of the early solar system, while signs of youthfulness indicate at least that rings have never been static phenomena.
astro-ph.EP
astro-ph
Planetary Rings Matthew S. Tiscareno Center for Radiophysics and Space Research, Cornell University, Ithaca, NY 14853 [email protected] ABSTRACT Planetary rings are the only nearby astrophysical disks, and the only disks that have been investigated by spacecraft (especially the Cassini spacecraft orbiting Saturn). Although there are significant differences between rings and other disks, chiefly the large planet/ring mass ratio that greatly enhances the flatness of rings (aspect ratios as small as 10−7), understanding of disks in general can be enhanced by understanding the dynamical processes observed at close-range and in real-time in planetary rings. We review the known ring systems of the four giant planets, as well as the prospects for ring systems yet to be discovered. We then review planetary rings by type. The A, B, and C rings of Saturn, plus the Cassini Division, comprise our solar system's only dense broad disk and host many phenomena of general application to disks including spiral waves, gap formation, self-gravity wakes, viscous overstability and normal modes, impact clouds, and orbital evolution of embedded moons. Dense narrow rings are found both at Uranus (where they comprise the main rings entirely) and at Saturn (where they are embedded in the broad disk), and are the primary natural laboratory for understanding shepherding and self-stability. Narrow dusty rings, likely generated by embedded source bodies, are surprisingly found to sport azimuthally-confined arcs at Neptune, Saturn, and Jupiter. Finally, every known ring system includes a substantial component of diffuse dusty rings. Planetary rings have shown themselves to be useful as detectors of planetary processes around them, including the planetary magnetic field and interplanetary impactors as well as the gravity of nearby perturbing moons. Experimental rings science has made great progress in recent decades, especially numerical simulations of self-gravity wakes and other processes but also laboratory in- vestigations of coefficient of restitution and spectroscopic ground truth. The age of self-sustained ring systems is a matter of debate; formation scenarios are most plausible in the context of the early solar system, while signs of youthfulness indicate at least that rings have never been static phenomena. 1. Introduction Planetary rings come in a diverse array of shapes and sizes. They may be broad or narrow, dense or tenuous, dusty or not, and they may contain various kinds of structures including arcs, -- 2 -- wavy edges, embedded moonlets, and radial variations. Rings share the defining characteristic of a swarm of objects orbiting a central planet with vertical motions that are small compared to their motions within a common plane. The latter arises because planets in our solar system (with the exceptions of Mercury and Venus, which have no known natural material in orbit) are fast-enough rotaters that their shapes are dominated by an equatorial bulge that adds a strong quadrupole moment (J2) to their gravity fields (see Section 1.1). This is a major contrast between rings and other astrophysical disks, which are not defined by asymmetry in an external gravity field but by the average angular momentum of the disk itself (in both cases, once a preferred plane is established, collisions among particles damp out the motions perpendicular to it). However, rings do have a number of similarities with other astrophysical disks, which add to the motivation for studying them. Unlike other known disk systems that are either many light-years away or (like the early stages of our solar system) far back in time, planetary rings can be studied up close and in real time. Thus, it is worthwhile to consider the parallels that can be drawn between planetary rings and the study of other disks (see Section 6). In this chapter, after some further introductory notes on important concepts (Sections 1.1 through 1.3), we will give an overview of the known ring systems, as well as systems where rings are unconfirmed but plausible, in Section 2. More detailed descriptions can be found in Section 3, which contains a discussion of various ring structures organized by type, with a focus on finding commonalities among rings in different locations that share certain qualities. In Section 4 we will discuss experimental methods of learning about rings, and in Section 5 we will discuss the age and origin of ring systems. Finally, in Section 6, we will discuss ways that planetary rings can illuminate the study of other astrophysical disks. 1.1. Orbital elements Rings are fundamentally populations of orbiting material. Therefore, to discuss structure within rings, we will occasionally refer to one or more of the six orbital elements that describe the orbit of any object around another object. These six parameters are simply a transformation of the six Cartesian parameters for position and velocity (x, y, z, x, y, and z) under the assumption that the object moves in the gravity field of a point mass (hereafter the "planet"). The derivation can be found in any textbook on orbital mechanics (e.g., Section 2.8 in Murray and Dermott 1999). A diagram of the orbit in space is found in Fig. 1. The size of the orbit, and its gravitational potential energy, is described by the semimajor axis a, which is the mean distance between the orbiting particle and the planet. The shape of the orbit is described by the eccentricity e; for a circular orbit e = 0, but real orbits that remain bound to the planet take elliptical shapes with 0 < e < 1, with most ring particles having e (cid:28) 1. Unbound orbits, either parabolic or hyperbolic, have e ≥ 1. The orbit plane may be inclined with respect to the reference plane (for ring applications, this is often the planet's equatorial plane), by an angle known as the inclination -- 3 -- Fig. 1. -- The geometry of (left) an elliptical orbit within the orbit plane and (right) the orbit plane within 3-D space. I. A non-zero inclination requires an account of the orbit plane's orientation, and thus its line of intersection with the reference plane (the "line of nodes") is described by the longitude of the ascending node Ω, measured with respect to a reference axis. Similarly, a non-zero eccentricity requires an account of the orientation of the ellipse within the orbit plane, and thus the line connecting the planet to the location of the particle's closest approach (its "periapse") is described by the argument of periapse ω. Finally, once the orbit has been defined by the five parameters already mentioned, the particle's position along the orbit can be given by its actual position (the true anomaly f ) or its time-averaged position (the mean anomaly M ) relative to periapse. Also commonly used are the longitude of periapse  = Ω + ω and the mean longitude λ = Ω + ω + M , which are not physical angles since they are the sums of angles not necessarily in the same plane, but they have the virtue of being reckoned from a stationary reference axis rather than a moving line and are useful as long as I is not too large. The osculating orbital elements, which are most simply calculated and most often used, assume that the planet's gravity field is that of a point mass. But for ring applications, the known planets are oblate (or bulged at the equators) due to their fast rotation. This is adequately described by adding to the account of the gravity field a positive quadrupole moment J2 (for details see, e.g., Section 4.5 of Murray and Dermott 1999), though it may be necessary to further include higher moments for applications requiring great precision. The presence of a non-zero J2, in addition to defining the Laplace plane1 for orbits near the planet, causes orbits to precess, in the prograde direction for apses (  > 0) and in the retrograde direction for nodes ( Ω < 0). A non-zero J2 also compromises the physical meaningfulness of the osculating elements, es- pecially for low-eccentricity orbits, introducing fast (i.e., orbit-frequency) variations in all six el- 1The Laplace plane is the plane about which orbits precess. When the vertical motions of objects are damped by mutual collisions, material will settle into a ring centered on the Laplace plane. -- 4 -- ements. The physical meaningfulness of orbital elements can be restored using a revised sys- tem of epicyclic orbital elements (Borderies and Longaretti 1987; Longaretti and Borderies 1991; Borderies-Rappaport and Longaretti 1994), which are based on the geometrical shape of stream- lines. These put the orbit-frequency variations back into an analogue of λ, leaving the other five elements to again describe a static (or at least slowly-varying) orbit. A useful algorithm for con- verting Cartesian coordinates into epicyclic orbital elements was devised by Renner and Sicardy (2006). 1.2. Roche limits, Roche lobes, and Roche critical densities The "Roche limit" is the distance from a planet within which its tides can pull apart a compact object. Simply speaking, a ring would be expected to reside inside its planet's Roche limit, while any disk of material beyond that distance would be expected to accrete into one or more moons. However, the Roche limit does not actually have a single value, but depends particularly on the density and internal material strength of the moon that may or may not get pulled apart (Canup and Esposito 1995). A simple value for the Roche limit can be calculated from a balance between the tidal force (i.e., the difference between the planet's gravitational pull on one side of the moon and its pull on the other side) that would tend to pull a moon apart, and the moon's own gravity that would tend to hold it together. This works out to (e.g., Eq. 4.131 in Murray and Dermott 1999) (cid:18) 4πρp (cid:19)1/3 aRoche = Rp γρ , (1) where R is radius and ρ is internal density, and the subscript "p" denotes the central planet. The dimensionless geometrical parameter γ = 4π/3 ≈ 4.2 for a sphere, but is smaller for an object that takes a non-spherical shape with its long axis pointing toward Saturn, as one would expect for an actively accreting body and as at least several of Saturn's ring-moons appear to do (Porco et al. 2007; Charnoz et al. 2007). Simply distributing the moon's material into the shape of its Roche lobe, with uniform density, yields γ ≈ 1.6 (Porco et al. 2007). However, fully accounting for the feedback between the moon's distorted shape and its (now non-point-mass) gravity field (Chandrasekhar 1969; Murray and Dermott 1999) leads to a rather smaller value, γ ≈ 0.85; on the other hand, some central mass concentration and the failure of a rubble pile to exactly take its equilibrium shape will likely prevent γ from becoming quite this low. Note that the moon's internal density ρ appears in Eq. 1. Thus the Roche limit is variable; a denser object can venture closer to the planet without danger than can an object that is less dense. The moon's diameter, on the other hand, does not appear in Eq. 1. Why then do we commonly imagine a large object getting pulled into smaller pieces when it ventures inside the Roche limit? This is because the Roche limit has been defined here as the distance within which an object can no longer be held together by its own gravity. Size becomes relevant for objects small enough to be held together by their internal material strength (which is not considered in Eq. 1) in spite of the -- 5 -- Fig. 2. -- Roche critical density ρRoche (Eq. 3, with γ = 1.6) plotted against planetary radii for Jupiter (red), Saturn (cyan), Uranus (green), and Neptune (blue). An object must have density higher than ρRoche to be held together by its own gravity; conversely, in the presence of abun- dant disk material, an embedded object will actively accrete as long as its density remains higher than ρRoche. The colored bars along the bottom and along the left-hand side show the extent of each planet's main ring system. For each, a solid circle indicates the outermost extent, and the corresponding minimum ρRoche, of the main rings. tidal forces that enter into the Roche calculation. In fact, it is often more useful in the context of rings to consider the limit from planetary tides not as a critical distance but as a critical density. At any given distance a from the planet, there is a Roche critical density ρRoche at which the moon's size entirely fills its region of gravitational dominance (its "Roche lobe" or "Hill sphere" of characteristic radius rHill). We can rearrange Eq. 1 to obtain (cid:18) m (cid:19)1/3 rHill = a 3Mp and ρRoche = 4πρp γ(a/Rp)3 = 3Mp γa3 , (2) (3) where m and Mp are the masses of the moon and planet, respectively. The first of these two expressions is the most useful for interpreting Fig. 2. Within a ring, where material for accretion is plentiful, any pre-existing solid chunk with internal density greater than ρRoche should accrete a mantle of porous ring material until its density decreases to match ρRoche. This process should govern the size and density of the largest disk- embedded objects (Porco et al. 2007; Charnoz et al. 2007). On the other hand, the density naturally achieved by transient clumps should be compared to ρRoche in order to predict whether disruption (rings) or accretion (discrete moons) will dominate in a particular location, and the persistent -- 6 -- existence of a ring implies that the densities of transient clumps do not exceed ρRoche (that is, we expect ρ (cid:46) ρRoche). As seen in Fig. 2, Saturn's rings extend outward to significantly lower values of ρRoche, approaching 0.4 g cm−3, than are seen in any of the other three known ring systems, probably reflecting their much lower rock fraction (and higher fraction of water ice) as already known from spectroscopy and photometry (Cuzzi et al. 2009). That ρRoche for Saturn's rings reaches values much lower even than the density of solid water ice indicates a high degree of porosity, which is not surprising for a system in balance between disruption and accretion. If the outer edge of a ring system is taken to be the transition between disruption-dominated and accretion-dominated regions, which is probably true at least for Saturn and Uranus given the large number of moons immediately outward of their main rings, and if the porosity of accreting objects is relatively constant among the different systems, then differences in ρRoche at the transition location probably reflect differences in bulk composition. Since Uranus has a transitional ρRoche three times that of Saturn (Fig. 2), we may well infer that its rings are made of material with a higher grain density, i.e. a significantly higher rock/ice ratio. Neptune's transitional ρRoche is intermediate between Saturn's and Uranus', possibly indicating an intermediate rock/ice ratio. Our inference, from the Roche critical density at the ring/moon transition, that the Uranus system is rockier overall than the Saturn system is consistent with the fact that the average density of Saturn's mid-size moons (Matson et al. 2009) is 1.2 g cm−3, while that of Uranus' major moons (Jacobson et al. 1992) is 1.6 g cm−3. We cannot test our inference that Neptune's rock/ice ratio is intermediate in this way, as Neptune has no indigenous major moons due to the cataclysm of Triton's capture (Goldreich et al. 1989). The extent of Jupiter's Main ring, in contrast to the other three ring systems, is clearly limited by the availability of material (which originates at source moons Metis and Adrastea and evolves inward, and which is not abundant) rather than by a disruption/accretion balance. However, its high value of ρRoche ∼ 1.7 g cm−3 places the only known limit on the densities (and thus masses) of the source moons. However, it may not be valid to assume that Metis and Adrastea are held together by gravity, as accreting masses must be, given the large gap in particle size between the ∼ 10-km moons and other Main ring particles, which observationally cannot be larger than 1 km (Showalter et al. 2007). This large gap in particle size might be explained if Metis and Adrastea are solid bodies originating further out, now held together by material strength, while no bodies of similar size are now able to form through in situ accretion. 1.3. Optical depth The amount of material in a system with general disk morphology can be measured in several ways. The most straight-forward is the surface density, the mass per unit surface area of the disk, though it must be borne in mind that a disk with greater vertical thickness will have proportion- ately lower volume density than a vertically-thin disk, even if both have the same surface density. However, surface density can be difficult to measure directly. A much more common observable is -- 7 -- Fig. 3. -- Schematic showing a slanted path through (a) a homogeneous slab and (b) flattened SGWs. The measured optical depth τ is proportional to sin B in the first case, but is relatively insensitive to elevation angle in the second. the optical depth τ , which can be thought of as the attenuation of a beam of light passing through the disk, measured in e-folding terms. That is, τ = − ln ≡ − ln T, (4) (cid:18) I (cid:19) I∗ where I is the observed intensity, I∗ is the unocculted intensity (e.g., from a background star), and the ratio between them is defined as the ring's transparency T . The optical depth is sensitive to both the number density and the size of ring particles, which can be obtained when both the optical depth and the surface density are known (e.g., Colwell et al. 2009a). For a given value of the surface mass density, the optical depth scales inversely with particle size, that is, with the ratio of volume to surface area. For a vertically-thick homogeneous disk, the optical depth is proportional to the path length through the disk, which in turn is proportional to µ ≡ sin B, where the elevation angle B is the angle between the line-of-sight and the ring-plane (Fig. 3a). In order to compare observations taken over a range of elevation angles, the normal optical depth, τ⊥ ≡ µτ , is often used. However, this parameter must be used with caution for disks that lack homogeneity and/or are close to a single layer thick. Colwell et al. (2007) found that τ⊥ varies strongly with µ in the B ring, and that the uncorrected τ is a more robust parameter in that case, indicating that the B ring is composed of vertically-thin nearly-opaque clumps with nearly-transparent gaps between them (Fig. 3b), and that the optical depth is controlled by the relative abundance of clumps and gaps (see Section 3.1.4). In numerical simulations (see Section 4.1), the photometric optical depth τ (Eq. 4) is cum- bersome to calculate, but a useful proxy known as the dynamical optical depth τdyn can be found by summing the total cross-section area of all simulated particles and dividing by the area of the simulation patch. This quantity turns out to be equal to the photometric optical depth as long as particles are randomly distributed, as the Gaussian probability of particle overlap plays the same role when using τdyn to calculate the total transparency that the exponential plays when using Eq. 4. However, for high values of τ , when the distance between particles becomes comparable to  -- 8 -- Fig. 4. -- Galileo image mosaic of Jupiter's rings, seen nearly edge-on at very high phase angle, annotated to show the primary components of the ring system. Image sensitivity increases from left to right, in order to show the increasingly faint structure. Figure from Ockert-Bell et al. (1999). the particle size, particles become constrained as to the locations in space they can occupy and the two quantities diverge. Specifically, τ /τdyn (cid:39) 1 + kD, (5) where the volume filling factor D is calculated from particle radius, disk scale height, and τdyn, and k is a scalar of order unity (Salo and Karjalainen 2003; Tiscareno et al. 2010a). Furthermore, the existence of microstructure such as self-gravity wakes causes the distribution of particles to be strongly non-random, and can cause τdyn to diverge strongly from the photometrically observed τ . 2. Rings by planetary system 2.1. The rings of Jupiter Jupiter is adorned by the simplest of the known ring systems. All of its rings are tenuous and composed of dust-sized2 particles. As the only confirmed ring system without any dense component, and by far the least massive (Burns et al. 2004), Jupiter's is the only ring system to have been discovered by spacecraft without having previously been seen from Earth either directly (as Saturn's) or by stellar occultations (as Uranus' and Neptune's). The Main ring was first clearly described from Voyager 1 images (Owen et al. 1979) after initial hints from charged- particle detectors aboard Pioneer 11 (Fillius et al. 1975; Acuna and Ness 1976; Burns et al. 2004) The basic structure of Jupiter's rings is well understood (see Burns et al. 2004, for a recent comprehensive review). The Main ring and the two Gossamer rings are like three nested "tuna 2Throughout this work, we will use the word "dust" to refer to µm-sized particles regardless of their composition. -- 9 -- Fig. 5. -- New Horizons images of Jupiter's Main ring at low phase (upper panel) and at high phase (lower panel), respectively showing the structures composed of macroscopic particles and the dusty envelope. Credit: NASA. cans" (Fig. 4), with radius set by the semimajor axis (a) of the ring's source moon and vertical height by the moon's vertical excursions relative to Jupiter's equatorial plane (a sin I, for inclination I). Particles enter the ring as ejecta from micro-meteoroid impacts onto the moon (Burns et al. 1999), and begin with orbital parameters a, e, and I (see Section 1.1) similar to the moon's. The tuna-can structure arises as the orientations of particles' orbit planes (Ω) become quickly randomized due to small variations in a, and thus in the precession rate. Particles evolve inward under Poynting-Robertson drag (Burns et al. 1999), thus filling out the cylindrical shape. A double- layered vertical structure arises dynamically because any orbiting particle spends more time at its vertical maxima than it does in the midplane. The increasing vertical thickness of the rings arises because Amalthea's inclination is larger than that of Metis or Adrastea, while Thebe in turn has an even larger inclination. The two Gossamer rings consist entirely of material evolving inward in this way, though the Thebe Gossamer ring has an additional segment extending slightly outward from Thebe's orbit, which has been attributed to charging and discharging of grains as they pass into and out of the planet's shadow (Hamilton and Kruger 2008). Further inward, the ∼1000-km-wide core of the Main ring contains a significant population of cm-size and larger particles lying between the orbits of Adrastea and Metis. This core, which is the only component of the ring system not composed of dust and the only one that appears bright at low phase angles,3 is composed of several ringlets (including one lying just outside Adrastea's orbit), whose cause is not known. New Horizons imaging limits the sizes of objects in this belt to be <1 km (Showalter et al. 2007) -- other than 3The phase angle is formed by the Sun-object-observer lightpath. Dust-sized particles, having size comparable to the wavelength of visible light, tend to diffract light forward and are brightest at high phase angles. Larger objects tend to reflect light and are brightest at low phase angles. -- 10 -- Adrastea and Metis themselves, which have mean radii of 8 and 22 km, respectively. However, Showalter et al. (2007) did find several azimuthal clumps in a ringlet just inward of Adrastea (see Section 3.5). A dusty component of the Main ring extends inward of its core, also evolving under Poynting-Robertson drag (Fig. 5). The dusty component of the Main ring has a vertical scale height that increases monotonically with decreasing distance to Jupiter (Ockert-Bell et al. 1999). When the inward-moving material reaches a radius of 122,800 km from Jupiter's center, it becomes strongly affected by a 3:2 "Lorentz resonance" (Burns et al. 1985) between its orbital period and the rotation period of Jupiter's magnetic field. Inward of this location, the vertical extent of the ring increases dramatically, forming the toroidal Halo ring. Material in the Halo ranges tens of thousands of km above Jupiter's ring plane, though most of its material is concentrated within just a few hundred km (Burns et al. 2004). 2.2. The rings of Saturn Saturn possesses by far the most massive and the most diverse of the known ring systems (Fig. 6). The only ring system known before recent decades, Saturn's rings were among the first objects observed through a telescope, by Galileo Galilei in 1610, explicated as a disk by Christiaan Huygens, proved to consist of individual particles on independent orbits by James Clerk Maxwell, and have in general been the focus of much productive study by astronomers over the past four centuries (Alexander 1962; Van Helden 1984; Miner et al. 2007). The main part of the rings comprises the solar system's only known broad and dense disk (Section 3.1), which was found by G. D. Cassini to be divided into two parts -- now called the A and B rings, separated by what is now called the Cassini Division. Furthermore, the latter is now known to be not an empty gap but simply a region of the disk with more moderate surface density, similar in character to the C ring, which lies inward of the B ring and was discovered in 1850. There are a small number of truly empty radial gaps in the dense disk of the main rings, most of them in the C ring and Cassini Division but two in the outer A ring, all of them sharp-edged. These gaps are named for scientists who have made contributions to the study of Saturn's rings. The two A-ring gaps are held open by moons at their centers, and the C ring's Colombo Gap is known to be held open by a resonance with Titan, but most of the gaps remain unexplained. Recent work by Hedman et al. (2010a) suggested that all the Cassini Division gaps are due to a secondary resonance associated with the Mimas 2:1 resonant mechanism that defines the nearby outer edge of the B ring, though this idea has yet to be successfully worked out in detail (Spitale and Porco 2010). A diverse array of narrow rings and ringlets resides within ring gaps, some of them dense and sharp-edged and others diffuse and/or dusty. They often are given the same name as the gap within which they reside, though several have been given nicknames. These structures, which can be compared with narrow rings around other planets, are discussed in Section 3.2. -- 11 -- Fig. 6. -- This Cassini image mosaic shows Saturn's tenuous D, E and G rings with comparable brightness to the main disk, which occurs because the viewing geometry is at high phase angle (in fact, in eclipse) and also views the unlit face of the main disk. The darkest part of the main disk is actually the densest and most opaque, namely the mid- to outer-B ring. The Cassini Division is difficult to distinguish from the A ring in this view. The markings on the planet do not line up with those on the rings because the latter are due to sunlight filtering directly through the rings while the former are due to light reflected off the rings, then reflected off the planet, and then filtered through the rings again. The Sun, which is actually behind Saturn, can be seen refracted through the planet's atmosphere at 7 o'clock. Enceladus (actually, only its geyser plume is bright in this geometry) can be seen embedded in the E ring at 8 o'clock. The Earth can be seen as a pinpoint of light between the F and G rings at 10 o'clock. Credit: NASA and M. Hedman, annotated by the author. The Saturn system also contains the most diverse retinue of tenuous dusty ring structures known in the solar system, discussed in Section 3.4. The main components, given letters in order of their discovery, are the D ring situated innermost between the main rings and Saturn's atmosphere, the dense F ring (Section 3.3) just off the edge of the A ring, and the G and E rings farther out. The region between the A and F rings, now known as the Roche Division,4 contains a tenuous dusty sheet, and several other rings or ring arcs are named for moons whose orbits they share. The largest known ring in the solar system, the Phoebe ring, was recently discovered by the Spitzer Space Telescope (Verbiscer et al. 2009). This ring is also the only known ring to be tilted from its planet's equatorial plane (it lies in the plane of Saturn's orbit, as solar perturbations are much more important than Saturn's J2 at its distance) and is likely the only known ring whose particles orbit in the retrograde direction (if indeed its material is primarily derived from Phoebe, 4As recently formalized by the IAU, a "division" is defined as a region between two lettered rings that contains a sheet of material, while a "gap" is a clear region within a lettered ring that may or may not contain one or more ringlets (http://planetarynames.wr.usgs.gov/append8.html). Main Rings (D, C, B, Cass. Div., A, F) E Ring G Ring Enceladus Refracted image of the Sun Earth -- 12 -- which orbits retrograde). As particles in the Phoebe ring spiral inward under Poynting-Robertson drag, they preferentially impact the leading hemisphere of Iapetus (Soter 1974; Tamayo et al. 2011), which, together with solar-driven thermal processing, appears likely to explain the strong brightness dichotomy on the surface of that moon (Spencer and Denk 2010; Denk et al. 2010). The end of Cassini 's initial 4-year mission at Saturn (though its extended mission continues) has occasioned several recent reviews of Saturn's rings, including articles by Cuzzi et al. (2010) and Esposito (2010). A recent comprehensive review in five parts discussed the rings' structure (Colwell et al. 2009b), dynamics (Schmidt et al. 2009), particle sizes and composition (Cuzzi et al. 2009), diffuse rings (Hor´anyi et al. 2009), and origins (Charnoz et al. 2009a), in addition to a review of pre-Cassini understanding (Orton et al. 2009). 2.3. The rings of Uranus All of Uranus' main rings are narrow, and many are eccentric and/or inclined, unlike the broad disk of Saturn. On the other hand, many of Uranus' rings are dense and sharp-edged, unlike the diffuse rings of Jupiter. Thus, the Uranian system represents a third paradigm for ring systems, one that truly deserves the label of "rings" in the plural. The main set of 10 narrow rings (including all the named rings except dusty ζ, ν, and µ) occupies a fairly small radial range from 1.64 to 2.00 RU from Uranus' center (Fig. 7), inward of Uranus' 13 small inner moons except that the innermost moon Cordelia is inward of the  ring. A panoply of unnamed dusty rings was seen interspersed with the main rings in the single high-resolution high-phase image taken by Voyager 2 (Fig. 8), and gaps in these have been cited as evidence for additional moons (Murray and Thompson 1990, 1991). The so-called "Portia group" of eight moons packed into an annulus from 59,100 to 76,500 km from Uranus' center (i.e., 2.31 to 2.99 RU) appears from orbital simulations to be dynamically unstable on timescales of 106 to 108 years (Duncan and Lissauer 1997; Showalter and Lissauer 2006). The dusty ν ring, which lies between two of the moons in this group, may well be the detritus of a recent significant collision, perhaps the disruption of a moon. Outward beyond the Portia group, the µ ring is centered on the orbit of Mab (Showalter and Lissauer 2006, see Section 3.4). Given their dynamical instability, the Portia group of Uranian moons are probably constantly evolving by means of occasional collisions followed by the re-accretion of material into new moons (Showalter and Lissauer 2006; Dawson et al. 2010; French and Showalter 2012), and may have looked rather different from its present configuration over most of solar system history, though what effect this might have on the main rings is unknown. The contrast between the main Uranian rings and the Portia group may simply be the difference between a particle population dominated by disruption and one dominated by accretion. The mean surface density of the Portia group region is ∼ 45 g cm−2, calculated by spreading the moons' mass evenly over the annulus containing them. This is comparable to the surface density of dense rings such as Saturn's A ring, though direct comparisons to the Uranian rings are difficult as the masses of the latter are poorly known. The Roche critical density (see Section 1.2) for the boundary between the two regions is 1.2 g cm−3, -- 13 -- Fig. 7. -- Uranus' main rings are situated immediately inward of a retinue of small moons. If one were to spread the mass of Uranus' "Portia group" of moons evenly over the annulus they occupy, the surface density would be similar to that of Saturn's A ring. This moon system may be very similar in origin to the known ring systems, except that the natural density of accreted objects is larger than the Roche critical density (i.e., it is beyond the "Roche limit," see Fig. 2) so that any moon that gets disrupted by a collision (which ought to have happened many times over the age of the solar system) will simply re-accrete. Credit: Wikimedia Foundation, annotated by the author. which possibly indicates a high rock fraction, especially considering the high internal porosity inherent in accretion of small bodies with low central pressures. The composition of the Uranian rings is almost entirely unknown, as Voyager did not carry an infrared spectrometer with enough spatial resolution to detect the rings. However, it is clear from their low albedo that at least the surfaces of the ring particles cannot be primarily water ice. Color imaging indicates that the Uranian rings are dark at all visible wavelengths, indicating a spectrum similar to that of carbon. Most of Uranus' rings have been given Greek letters (α, β, γ, δ, , η, λ, ζ, µ, and ν) in order of their discovery, except for three which are labeled with numbers (4, 5, and 6). This idiosyncratic "Por%a group" Main Rings -- 14 -- Fig. 8. -- This composite image of Uranus' main rings in forward-scattered (left) and back-scattered (right) light shows that a network of dust structures is interleaved with the planet's dense main rings. The disjoint in the  ring is due to its eccentricity. As the left-hand image is the only high-phase image ever successfully taken of Uranus' rings (by the post-encounter Voyager 2 ), the detailed workings of the dust structures remains largely unknown. Credit: NASA and Wikimedia Foundation. system can be traced back to their simultaneous discovery by two different research groups, one of which (Elliot et al. 1977) labeled the rings with Greek letters while the other (Millis et al. 1977) numbered them. The former system was given priority for future use, but three of the rings had not been observed by the former research group and thus retained as their labels the numbers given to them by their discoverers (Miner et al. 2007). A comprehensive review of the Uranian rings, up to and including the Voyager 2 encounter, was published in two parts discussing the rings' structure (French et al. 1991) and particle properties (Esposito et al. 1991). -- 15 -- Fig. 9. -- In Neptune's ring system, uniquely, narrow rings and diffuse rings and moons are all interspersed together. Credit: Wikimedia Foundation. 2.4. The rings of Neptune Neptune's ring system, like that of Uranus, consists primarily of a few narrow rings, though Neptune's are generally less dense, higher in dust, less sharp in their edges, and farther from their planet than those of Uranus. Neptune's rings are named for individuals associated with the 1846 discovery of Neptune. The Le Verrier, Arago, and Adams rings are narrow, while the Galle and Lassell rings are tenuous sheets of dust (Fig. 9). The Adams ring, Neptune's most substantial, is best known for its series of arcs, the first ever discovered, which are discussed in Section 3.5. Unlike those of Saturn and Uranus, Neptune's ring system is thoroughly interwoven with known moons. This is at least partly enabled by the fact that Neptune's rings are the farthest from their planet, in terms of planetary radii, and have the lowest value of ρRoche at the inner edge of the ring system (Fig. 2). Still, the presence of rings, rather than accreted moons, must indicate that the natural density of accreted objects is lower than ρRoche. Thus, it may be that the ring-moons Naiad, Thalassa, Despina, and perhaps Galatea, either originated farther from the planet than their current position, or accreted in a formerly more dense ring. The composition of the Neptunian rings, like that of the Uranian rings, is unknown due to Voyager 's inability to detect them in the infrared. However, again like the Uranian rings, the low -- 16 -- albedo of Neptunian ring particles makes it clear that at least their surfaces cannot be primarily water ice. A comprehensive review of the Neptunian rings, up to and including the Voyager 2 encounter, was given by Porco et al. (1995). 2.5. Unconfirmed ring systems The four giant planets are the only bodies known to have rings, as just described. Here we discuss bodies for which rings have been seriously discussed but not observed. 2.5.1. Mars Mars has been predicted to have a tenuous ring system comprising dust grains ejected from its moons Phobos and Deimos by meteoroid impactors (Hamilton 1996; Krivov and Hamilton 1997, and references therein). Simulations by Burns et al. (2001) indicate that Deimos' ring should be offset away from the Sun and tilted out of Mars' equatorial plane by the Sun's perturbations. However, attempts to observe rings around Mars have been unsuccessful to date (Showalter et al. 2006), and the image quality has progressed to the point that some models can now be observationally excluded. The lack of dust could be due to dust production rates being lower than expected, or the lifetimes of dust particles being shorter than expected. Solar radiation pressure limits the lifetimes of dust particles (especially smaller ones) by driving their orbital eccentricity to values so high that they impact the planet. In situ observations by Mars-orbiting spacecraft of anomalies in the solar wind magnetic field were interpreted in the 1980s as being due to Martian rings, but more extensive measurements by the magnetometer aboard Mars Global Surveyor showed that observable fluctuations are likely due to well-known solar wind or bowshock phenomena (Øieroset et al. 2010) Looking for rings around a solid planet like Mars at low phase is more difficult than similar observations at gas giant planets because of the former's lack of atmospheric methane. Because methane has very strong absorption bands (e.g., at 2.2 µm), images of gas giants taken at selected wavelengths will see a greatly darkened planet, facilitating the detection of faint rings. On the other hand, looking for Martian rings at high phase is less effective than for other dusty rings because particles smaller than ∼ 50 µm are expected to be depleted due to radiation pressure (Hamilton 1996; Showalter et al. 2006) -- 17 -- 2.5.2. Pluto Pluto, like Mars, could harbor a tenuous ring system of dust derived from its small moons Nix and Hydra and P4 (Stern et al. 2006). Charon, as discussed in Section 3.4, is paradoxically less likely to be a major source of dusty rings because its gravitational field will more efficiently retain any dust ejected from its surface. Observations to date (Steffl and Stern 2007) are not sensitive enough even to rule out the conservatively estimated normal optical depth τ⊥ < 10−6 suggested by Stern et al. (2006), though more sensitive observations were very recently obtained (Showalter et al. 2011a). Further Earth-based observations may improve on this sensitivity, and a clearer picture of Pluto's rings (or lack thereof) should come from imaging during the planned New Horizons flyby in 2015, especially during the post-encounter period when the phase angle will be high. However, both of the handicaps discussed above for Mars, the lack of atmospheric methane and the loss of smaller dust grains due to radiation pressure, are also likely to hamper the detection of any Plutonian rings. 2.5.3. Rhea and other moons Rhea, the largest of Saturn's airless moons, is in the opposite situation from Mars and Pluto, with no clear theoretical prediction but a claim that rings have been observed. Jones et al. (2008) reported unusual electron-absorption signatures detected by the Magnetospheric Imaging Instru- ment (MIMI) during Rhea flybys of the Cassini spacecraft, and attributed these signatures to rings. What would be the first known ring system around a moon includes several narrow bands embedded within a broad diffuse cloud. The MIMI instrument detects changes in the miasma of charged particles through which the spacecraft is constantly passing, enabling inferences regarding solid and magnetospheric structures that shape the plasma environment. A good analogy is the way a person driving in a blinding rain can perceive having driven under a bridge by the sudden cessation of raindrops hitting the windshield.5 However, an extensive search for Rhea rings using Cassini ISS images (Tiscareno et al. 2010b) places severe limits on any possible Rhea rings. Using the calculations of Jones et al. (2008), which assume that a particle's ability to block electrons increases linearly with its mass (and thus its volume), the ISS observations limit the narrow rings to a characteristic particle size r > 10 m, and µm-sized dust in any significant abundance is unequivocally excluded. This minimum particle size is unrealistically large given that such large particles must constantly be eroded to smaller sizes, which then would have been detected in the ISS images. Furthermore, Tiscareno et al. (2010b) pointed out that the electron penetration depth for the low-energy electrons used by Jones et al. (2008) is much smaller than the suggested particle sizes, so that a particle's ability to block electrons 5Credit: G. H. Jones in JPL podcast, 6 March 2008 (http://www.jpl.nasa.gov/podcast/content.cfm?content=671). -- 18 -- Fig. 10. -- Comparison of the radius r and number density n of particles, the latter expressed in terms of the extinction coefficient πr2n of putative Rhea rings, inferred from charged-particle absorptions observed by Cassini MIMI (Jones et al. 2008) and imaging non-detection by Cassini ISS (Tiscareno et al. 2010b), shown in red and green, respectively. Dashed lines indicate requirements previously claimed (Jones et al. 2008) for particle sizes larger than the electron penetration depth. For narrow rings (left), even allowing the latter claim, the combined observations require particles larger than 10 m in radius, indicating an unrealistic lack of smaller particles. Furthermore, once allowance is made for the role of the electron absorption length (horizontal lower boundary to red area), the imaging non-detection rules out any form of absorption by solid material as the cause of the observed charged-particle absorptions for both narrow rings and broad cloud. Figure from Tiscareno et al. (2010b). should increase linearly with its surface area rather than its volume. Thus, Tiscareno et al. (2010b) re-calculate the ring optical depth required to explain the Jones et al. (2008) observations to be several orders of magnitude higher than the values excluded by the ISS observations, both for the narrow rings and the broad cloud (Fig. 10). It seems most likely, therefore, that the signatures detected by the Cassini MIMI instrument are due to some magnetospheric phenomenon, and not to rings of solid material around Rhea. Following the MIMI announcement that Rhea might have rings, Schenk and McKinnon (2009) reported a clumpy and discontinuous chain of discrete bluish splotches, up to 10 km wide, aligned along more than half the circumference of a great circle inclined by ∼2◦ to Rhea's equator. Schenk et al. (2011), formally presenting that finding after publication of the ISS non-detection, argued that the band may be a sign of a former Rhean ring, even if none exists now. However, it remains unclear whether a ring system is even plausible at Rhea (whose shape is triaxial rather than oblate), what would be the source of its material, whether a tenuous ring would assume the flatness (generally produced only in dense rings by collisional evolution, while tenuous rings generally are vertically thick) implied by the observed narrow band, and whether the observed color variations are the likely result of such impacting material. -- 19 -- Among other Saturnian moons, Mimas and Tethys also have equatorial color features extending over ±40◦ and ±20◦ of latitude, respectively, much wider than Rhea's and possibly the result of magnetospheric bombardment (Schenk et al. 2011). Iapetus has a 13-km-high equatorial mountain range extending across >110◦ of longitude (Porco et al. 2005a; Giese et al. 2008). Several endogenic hypotheses have been suggested for this structure, but Ip (2006) argued that it was created by ring material falling out of Iapetan orbit, and Levison et al. (2011) further developed the idea by proposing a giant impact to create the ring and sketching a scenario that would simultaneously account for Iapetus' surprisingly oblate shape. While Iapetus may be a more hospitable site for rings than Rhea, given its oblate shape and large Hill sphere, it remains unclear just how or whether orbiting material would be incorporated into a mountain range, nor how such a ring would evolve and fall out. 2.5.4. Exoplanets While hundreds of planets have now been detected in orbit about other stars using methods including radial velocity, transits, astrometry, and direct imaging (Seager 2010), no exoplanet yet has a confirmed ring system. The signature in a transit observation expected from an exoplanet ring system was discussed by Barnes and Fortney (2004), and a few exoplanet detections have been able to set meaningful upper limits on putative ring systems (e.g., Brown et al. 2001), but it is too early to do statistics on the frequency of exoplanet rings, especially since the available observations with sufficient sensitivity are generally for "hot Jupiters" that orbit very close to their stars. There are many factors stacked against the detection of rings around hot Jupiters (Schlichting and Chang 2011), including small Hill spheres, low obliquities6 that would cause rings to be seen edge-on (see also Ohta et al. 2009), loss of ring particles to Poynting-Robertson drag and viscous drag from the planet's exosphere (see also Gaudi et al. 2003), and equilibrium blackbody temperatures too high for even refractory materials to remain solid. However, each of these factors is mitigated for planets (cid:38)0.1 AU from their host stars (Schlichting and Chang 2011). Furthermore, warping of the ring planes (Laplace planes) of "warm Saturns," diagnostic of their planetary oblateness (see Section 1.1) may be discernible in transit lightcurves (Schlichting and Chang 2011). The only known exoplanet for which a ring system has been specifically proposed is Fomal- haut b, which was the first7 exoplanet to be detected by direct imaging (Kalas et al. 2008). The orbit of Fomalhaut b is ∼ 115 AU from its star and maintains the inner edge of an eccentric debris belt that was known before the planet was (see Section 6). The brightness of Fomalhaut b in visible light, together with its dimness in the infrared, has led Kalas et al. (2008) to suggest the planet has a large ring system, which would significantly increase its visible brightness via reflection without affecting its emitted infrared radiation. However, a ring with a surface brightness similar to that 6Here, we refer to the inclination of the planet's equatorial plane with respect to the line of sight from Earth. 7along with the three planets of the HR 8799 system, announced at the same time -- 20 -- of Saturn's A ring while extending to the planet's Roche radius would reflect far too little flux to account for the observations, so it is necessary to invoke a disk extending to ∼ 35 planetary radii, approximately the distance from Jupiter to its outermost large moon Callisto. Such a large disk would not be stable against accretion (Section 1.2) and thus would perhaps be more of a dynamically evolving proto-satellite disk than a stable ring system. A complex two-month-long eclipse observed for the star 1SWASP J140747.93-394542.6 may have been due to a disk surrounding an otherwise-unknown planet, though it is also possible that the occulting disk instead adorns a low-mass stellar companion (Mamajek et al. 2012). 3. Rings by type 3.1. Dense broad disks Saturn: C ring B ring Cassini division A ring Although the idea that Saturn's main ring system is composed of "countless tiny ringlets" continues to appear even in the professional literature, it is inaccurate. It is much better to say that the main ring system is a nearly continuous disk with density that varies radially but only a few true gaps that would separate one "ring" from another. Not only do waves travel radially through the disk (Section 3.1.1), but material likely does as well through ballistic transport (Durisen et al. 1989, 1992) and direct migration (Tiscareno et al. 2010c). Although Saturn's ring system, taken as a whole, is the only dense broad disk known to us, its four main components are quite different from each other and thus still offer room for comparative study. In addition to their wide differences in structure, the components vary in location and density. The B ring is by far the most dense, with measured τ of 5 or more, followed by the A ring at τ ∼ 0.5, while the C ring and Cassini Division both have τ ∼ 0.1 (Fig. 11) The B ring's position astride the synchronous distance, at which the orbital period of particles equals the rotation period of the planet (and thus of its magnetic field) is thought to influence at least some of its properties (e.g., spokes), while the A ring's outermost position causes it to be more susceptible to accretion- related processes. The differences between the C ring and the Cassini Division, despite their apparent similarities in optical depth and composition, may also be primarily due to their different locations with respect to Saturn, its moons, and the B ring. -- 21 -- Fig. 11. -- An optical depth profile (top) and true-color image (bottom) of Saturn's main ring system. Figure from Colwell et al. (2009b). 3.1.1. Spiral waves Spiral density waves and spiral bending waves are the most widespread well-understood phe- nomena in dense rings. First described for the case of galaxies by Lin and Shu (1964), and applied to planetary rings by Goldreich and Tremaine (1978a,b, 1980), they can arise in any disk subject to a periodic perturbation. In Saturn's rings, the predominant mechanism is forcing from a per- turbing moon. At discrete locations in the disk, the orbits of individual particles are resonant with the forcing and become excited. For a resonance with a moon, this happens when the resonance argument ϕ, the quantity that librates about a constant value at the resonant location, is of the form8 ϕ = (m + k)λ(cid:48) − (m − 1)λ − X − kX(cid:48), (6) where m and k are integers, primed quantities refer to the forcing moon, and X is some integer combination of  or Ω (see Section 1.1 for orbital elements). For any given value9 of m and k, and for any particular forcing moon, there is a unique location in the ring-plane at which the resonance 8For brevity, this discussion is limited to inner Lindblad resonances and inner vertical resonances, where the disk is inward of the forcing moon. Nearly all known spiral waves in rings are of this kind, though Tiscareno et al. (2007) detected inwardly- propagating spiral density waves excited by outer Lindblad resonances with Pan. 9The azimuthal parameter m gives the number of spiral arms in the resonant wave pattern, while k + 1 is the "order" of the resonance, with first-order resonances generally being strongest, followed by second-order, etc. -- 22 -- occurs; this location can be found by differentiating Eq. 6 and setting the time-derivative ϕ = 0, since n (= λ), , and Ω are all known functions of radial location in the ring-plane (a), and the orbital frequencies of the moon are known. In fact, since n and n(cid:48) are much larger than the others, the approximate location of the resonance can be found from them alone, and the resonance is generally labeled with the ratio (m+k):(m-1). At a Lindblad resonance (LR), where the identity of X (though not necessarily X(cid:48)) is , the eccentricity of the ring particle is excited, and a spiral density wave (a compression wave) propagates radially outward away from the resonant location. At a vertical resonance (VR), where the identity of X is Ω, the inclination of the ring particle is excited, and a spiral bending wave (a transverse wave) propagates radially inward10 away from the resonant location. For more details see, e.g., Section 10.3 of Murray and Dermott (1999). When the perturbation is not too strong, the radial variation in surface density ∆σ(r) for a spiral density wave has the form ∆σ(r) (cid:39) Aξ cos(ξ2/2)e−(ξ/ξD)3 , (7) where A is an amplitude, ξD is a damping constant, and the dimensionless radial distance from resonance ξ is given by (cid:18) 3(m − 1)n2 LrL (cid:19)1/2 · r − rL rL ξ ≡ 2πGσ0 , (8) where rL and nL are the radius and mean motion at the location of exact resonance, σ0 is the unperturbed surface density, and G is Newton's constant. This simplified equation is valid only for the downstream portion of the wave, ξ (cid:38) 1. We have also removed the phase term, which causes the wave to have a spiral shape. For the linear theory in its full form, see Goldreich and Tremaine (1982), Shu (1984), or Tiscareno et al. (2007). Spiral waves, especially weak ones, are useful structures that can be thought of as in situ scientific instruments placed in the rings. Eq. 7 describes a sinusoid with frequency that increases linearly with distance from the resonance; the rate of increase is inversely proportional to the background surface density σ0, and thus can be used to measure it. The sinusoid oscillates within an envelope whose amplitude begins by increasing linearly, but then turns over and begins to decay as the exponential damping term takes over; the location of that turn-over is governed by the damping constant ξD, which can thus be obtained and used to constrain the ring's dynamic viscosity (Goldreich and Tremaine 1978b; Shu 1984). Additionally, the mass of the perturbing moon can be obtained from the amplitude A, its orbital phase from the wave's phase term, and the absolute distance from Saturn's center from the resonance location rL, though these parameters are often already known too precisely for the density wave to make a significant contribution. An algorithm for extracting these parameters from observed weak density waves (Fig. 12) was described 10The exception is the nodal resonance, labeled -1:0, in which the mean motion of the ring particles is resonant with the forcing moon's nodal precession Ω. This peculiar resonance has a negative pattern speed, and its bending wave propagates outward (Rosen and Lissauer 1988). -- 23 -- Fig. 12. -- A density wave fitting process carried out on a radial brightness profile from a Cassini ISS image. a) Initial radial scan. b) High-pass-filtered radial scan, with the final fitted wave shown in green. c) Wavelet transform of radial scan, with blue line indicating the filter boundary, and the green dashed line indicating the fitted wave's wavenumber. d) Unwrapped wavelet phase, with green dashed line indicating the quadratic fit and green open diamond the zero-derivative point. e) Residual wavelet phase, showing that the interval used for the fit is the interval in which the phase behaves quadratically. f) Wave amplitude, the local maximum of which (vertical dotted line) gives ξD; scale bar indicates the smoothing length of the boxcar filter. Figure from Tiscareno et al. (2007, q.v. for details of the wave-fitting process). -- 24 -- Fig. 13. -- a) Ring surface mass densities σ0 from density wave analysis. b) Local ring viscosities; the associated rms velocity is within the ring plane, and likely enhanced by self-gravity wakes. Open circles indicate Voyager data, filled circles Cassini data. Based on figures from Tiscareno et al. (2007) and Colwell et al. (2009b), q.v. for details of individual measurements. by Tiscareno et al. (2007), making use of the spatially localized frequency spectra provided by a wavelet transform (Daubechies 1992; Torrence and Compo 1998). A profile of surface densities and viscosities in the Cassini Division and A ring is shown in Fig. 13. The Cassini measurements (filled circles) have much less scatter than the Voyager mea- surements (open circles) not only because of greater measurement sensitivity but also because they rely on weaker waves that adhere more closely to the linear theory (Eq. 7) but were not detectable in Voyager data. Non-linear spiral density waves occur when the amplitude of the density perturbation ∆σ becomes comparable to the background density σ0. In this case, instead of the quasi-sinusoidal profile of the linear case (Eq. 7, Fig. 12), the oscillation frequency no longer increases linearly with distance from the resonance, and additionally the wave morphology develops sharp narrow peaks and broad flat troughs. Significant progress has been made on developing an analytical model to describe non-linear density waves (for a description and list of references, see Schmidt et al. 2009). Most recently, Rappaport et al. (2009) extracted model parameters from a series of RSS occultation profiles of the Mimas 5:3 density wave, the strongest (and thus most non-linear) density wave in Saturn's rings (Fig. 14). Spiral bending waves behave similarly to spiral density waves in their wavelength dispersion, but with several differences in other characteristics. Because they are transverse waves, both -- 25 -- Fig. 14. -- Radial optical depth profiles from RSS occultations of the Mimas 5:3 spiral density wave (blue) with model fits (red). Figure from Rappaport et al. (2009). damping and non-linearity arise less readily, which in principle makes them even more useful as ring diagnostics. However, they are significantly less abundant than density waves, because they can only be excited by perturbing moons on inclined orbits. Also, bending waves are harder to observe because they do not directly affect the optical depth. Rather, because they appear as corrugations in the ring-plane, they are most readily seen when illuminated from a light source nearly in the ring-plane and oriented azimuthally so as to shine across waveforms rather than along them, so that the waveforms can maximize the effect of changing the path length and thus the observed τ , even while τ⊥ remains nearly constant (see Section 1.3). In practice, these conditions are met by occultations with low elevation angle and in images taken near in time to Saturn's equinox (which occurs every ∼ 14.5 yr). In Cassini images taken during the 2009 equinox, some of the strongest bending waves (whose peaks rise the highest out of the mean ring-plane) are seen to cast shadows. -- 26 -- 3.1.2. Gap edges and moonlet wakes There are 14 named gaps within Saturn's main rings: 4 in the C ring, 8 in the Cassini Division, and 2 in the A ring (Fig. 11). Both of the A ring's gaps contain known moons that clear them by exerting a torque on nearby ring material, and some gaps in the C ring coincide with known Lindblad resonances, but the other gaps are unexplained. Furthermore, even the two gaps that do contain known moons exhibit a surprising amount of unexpected behavior at their edges. When a ring particle passes through conjunction with a nearby moon, the moon's gravity imparts an eccentricity as well as very slightly pushing the particle's semimajor axis away from its own. On its now-eccentric orbit, the ring particle passes through periapse and apoapse once per orbit; during the same time period, an inward (outward) particle moves forward (backward) by a distance 3π∆a in the moon's frame of reference (Fig. 15). To derive this, consider Kepler's Third Law, which states that an object's orbital rate n decreases with increasing semimajor axis a, specifically n2a3 = constant. Differentiating this with respect to a, we find the equation for keplerian shear, dn da . (9) = − 3 2 n a At relative velocity v = a∆n, over one orbital period P = 2π/n the relative motion of a ring particle with respect to a perturbing moon is vP = 2πa∆n/n. Substituting Eq. 9, this becomes 3π∆a. The characteristic wavelength 3π∆a is ubiquitous in the vicinity of a gap-moon. Not only does the gap edge form waves with that wavelength, but streamlines with that wavelength penetrate into the disk (Fig. 15). Because of the increasing wavelength, the distance between streamlines (which correlates with surface density) forms a pattern called "moonlet wakes"11 that rotates with the gap-moon. It should be emphasized that these structures are not properly called "waves," as they do not propagate or otherwise dynamically evolve; rather, they are kinematic phenomena caused by the gap-moon organizing the orbital properties of the material around it into streamlines. However, some of Pan's moonlet wakes do reach high enough densities that the mutual self-gravity of particles enhances the peaks (Porco et al. 2005b) in a manner similar to non-linear density waves (Section 3.1.1). Before Pan or any other gap-moon had been discovered, the wavy edges of the Encke Gap were tracked by Cuzzi and Scargle (1985), and the nearby moonlet wakes by Showalter et al. (1986). Interpretation of these observed features in terms of the simple theory just described allowed these authors to constrain the position of the moon causing them, which led Showalter (1991) to discover Pan in archival Voyager images. Sharp ring edges can also be maintained at the locations of Lindblad resonances (see Sec- tion 3.1.1). The nature of the perturbation is similar to that in the impulse approximation described 11Moonlet wakes have little in common with self-gravity wakes (Section 3.1.4), despite an unfortunate similarity in terminol- ogy. -- 27 -- Fig. 15. -- (top) Wavy edges and moonlet wakes are seen at the edges of the Keeler Gap, surrounding the position of the gap-moon Daphnis. (bottom) Streamline diagram showing 3π∆a wavelength set up as ring material passes a gap-moon. Different wavelengths at different radial distances ∆a set up the "moonlet wakes" pattern. By Kepler's Third Law, inner material is moving faster than outer material (red arrows); by the same token, in the moon's frame of reference, inner material moves forward and outer material moves backward (white arrows). above, except that now the resulting streamline wavelength is exactly 1/m times the orbital cir- cumference at that location, so that repeated conjunctions combine in a resonant effect. In fact it can be shown that the resonant streamline wavelength 2πa/m reduces to 3π∆a in the case of large m (which is to say, small ∆a/a). For example, the outer edge of the B ring is coincident with the 2:1 LR with Mimas, and the outer edge of the A ring with the 7:6 LR with the co-orbital moons Janus and Epimetheus. Both edges generally exhibit the expected 2-lobed and 7-lobed (respectively) structure, though there are significant deviations that have been attributed to the time-variable orbits of Janus and Epimetheus for the A ring edge (Spitale and Porco 2009) and the excitation of normal modes in the nearby disk for the B ring edge (Spitale and Porco 2010). Of the 8 gaps in the Cassini Division, those not containing ringlets all have circular outer edges, and those not associated with known moon resonances all have freely-precessing elliptical inner edges (Hedman et al. 2010a). The Keeler Gap in the A ring also follows this pattern, with a nearly circular outer edge and an 32-lobed inner edge due to a resonance with Prometheus, though again the expected pattern is superposed with other -- 28 -- Fig. 16. -- Spectrograms of Cassini RSS data showing periodic microstructure in the inner A ring. The central horizontal line in each panel is the direct signal, while the side bands that occur at some locations are coherent diffracted signal from the periodic microstructure acting as a diffraction grating. Figure from Thomson et al. (2007). structure that may be due to additional free or forced modes (Tiscareno et al. 2005; Torrey et al. 2008). Some edges also exhibit vertical structure. The moon Daphnis, at the center of the Keeler Gap, has an inclined orbit that ventures ∼ 9 km above and below the ring plane (Jacobson et al. 2008). The resulting vertical corrugation of nearby portions of the gap edge were predicted and then seen by their shadows cast during the 2009 equinox (Weiss et al. 2009). A region of vertical structure, probably due to embedded moonlets on inclined orbits, has also been detected and tracked in the outer edge of the B ring (Spitale and Porco 2010). 3.1.3. Radial structure Several varieties of azimuthally-symmetric radial structure are found in Saturn's rings. In addition to the gaps discussed above, the two largest in size are a series of sharp-edged annuli in the outer C ring with optical depths several times higher than the surrounding background, which have been dubbed "plateaux", and an undulating variation in optical depth in the inner B ring. Both of these have radial scales of ∼ 100 km and have remained entirely unchanged during the 25 yr between Voyager and Cassini observations (Nicholson et al. 2008), as well as remaining unexplained. -- 29 -- Regions of short-wavelength axisymmetric waves have been found by occultations in the inner A ring and in the B ring (Thomson et al. 2007; Colwell et al. 2007). The wavelengths range from 0.15 to 0.22 km, and are locally monochromatic -- the waveforms interacted with the spectrally coherent Cassini RSS radio occultation signal as if they were a diffraction grating (Thomson et al. 2007). This structure (Fig. 16) has been explained in terms of viscous overstability, which occurs when a perturbation triggers an overly strong restoring force that resulting in continuing oscillations (for details and references, see Schmidt et al. 2009). Viscous overstability can occur when the ring's viscosity increases steeply with density, as naturally occurs in dense rings due to increasingly frequent collisions, and is sensitively affected by the strength of mutual self-gravity, the distribution of particle sizes, and the existence of self-gravity wakes. Work is ongoing to characterize the appearance of overstability waves in Cassini data, as well as their behavior in response to various environmental factors in simulations. Both the A and B rings have sharp drops in optical depth at inner edges, with a "ramp" region of gradually decreasing optical depth (with decreasing distance) inward of them. These ramps have generally been considered to be the outermost portions of the Cassini Division and the C ring, respectively (see Fig. 11). The similarities in morphology, for the inner boundaries of the only two truly dense (τ (cid:38) 0.5) broad disk structures known, are striking. The morphology of the ramp structure as well as the sharp edge at its outward boundary (which, in both known cases, is not correlated with any known resonance strong enough to explain it), as well as the apparent compositional similarity between the ramp material and the denser ring on the other side of the edge, has been explained in terms of ballistic transport, the radial movement of material due to micrometeroid bombardment (Durisen et al. 1989, 1992). However, doubt is cast on that hypothesis by recent measurements of the wavelength dispersion of a spiral wave (Section 3.1.1) stretching across the sharp edge in optical depth between the Cassini Division ramp and the A ring, which indicates that there may not be a corresponding sharp change in surface density at that location (Tiscareno et al. 2009a). 3.1.4. Self-gravity wakes The boundary between disruption-dominated regions (in which disks are stable) and accretion- dominated regions (see Section 1.2) is not a sharp one. In the outer parts of the region of stability for disks, gravitational instabilities drive temporary accretion that is quickly disrupted again by tides, forming a disk micro-structure known as self-gravity wakes (SGWs). This balance is characterized by Toomre's Q parameter (for details and references, see Schmidt et al. 2009), defined as (10) where cr is the radial velocity dispersion, σ is the local surface density, and κ ≡ n −  ≈ n (see Section 1.1). Gravitational instability (i.e., accretion) is generally avoided as long as Q is at least a few times unity, but can become prominent if random velocities are damped (lower cr) 3.36Gσ Q = crκ , -- 30 -- Fig. 17. -- Simulated self gravity wakes (SGWs). This microstructure develops within dense rings as particles clump together under their mutual self-gravity but are ripped apart again by Saturn's tides. Figure courtesy of R. P. Perrine and D. C. Richardson. and/or surface density σ is increased and/or at locations further from the planet (lower κ). Balance between accretion and disruption, leading to SGWs, occurs when Q ∼ 2. Data from the damping of spiral density waves (Section 3.1.1) indicate that cr and σ adjust themselves in order to maintain Q ∼ 2 over a wide region of the A ring (Tiscareno et al. 2007). SGWs generally have a webbed structure that is elongated in a characteristic direction (Fig. 17), usually a few degrees to a few tens of degrees from azimuthal. Perpendicular to the characteristic direction, there is a characteristic spacing given by the Toomre critical wavelength, λcr = 4π2Gσ κ2 . (11) Because of their non-axisymmetric structure and their finite vertical thickness, the brightness of a disk pervaded by SGWs depends on the observer's longitude. An observer looking along the direction of the elongated wake structures will see more of the gaps between the wakes than an observer looking across the wakes (Fig. 18), especially at low elevation angles. Colombo et al. (1976) were the first to suggest the presence of SGWs in Saturn's rings as an explanation for the observed azimuthal brightness asymmetry. Today, observations of ring photometry combined with simulations of SGWs are further refining understanding of ring properties (see Sections 4.1 and 4.2). Repeated stellar occultations with varying geometries are another way to gather information about SGW structure. Such observations from Cassini UVIS (Colwell et al. 2006, 2007) and Cassini -- 31 -- Fig. 18. -- The self-gravity wakes shown here (from the simulations of Salo et al. 2004) are canted at an angle of ∼ 21◦. At low elevation angle B, the transparency is higher when viewed along the wake structures (blue) than when viewed across the wake structures (red), leading to an azimuthal brightness asymmetry that depends on longitude relative to the observer (lower panel). Figure from Schmidt et al. (2009). -- 32 -- Fig. 19. -- Radial profiles of measured parameters for self-gravity wakes in the A ring. Work is ongoing (Hedman et al. 2011a) to understand the "halos" surrounding density waves (dotted lines) but not bending waves (dashed lines), within which the intensity and orientation of SGWs are altered, in addition to changes in spectral absorption by water ice and in the abundance of propellers (Section 3.1.5). Figure from Hedman et al. (2007a). VIMS (Hedman et al. 2007a; Nicholson and Hedman 2010) have been interpreted in terms of simple models with a bimodal distribution of optical depths, treating the SGWs as nearly opaque with optical depth τwake, and the intervening space as characterized by a constant lower optical depth τgap. Both teams have produced resulting data sets of various wake parameters as a function of radial location in the disk (Fig. 19). However, Tiscareno et al. (2010a) investigated the density contrast and the distribution of densities in simulated SGWs and found instead a trimodal distri- bution (Fig. 20). In their histograms (Fig. 20), the high-τ peak corresponds to the τwake assumed in simpler models, while the low-τ peak is practically transparent by contrast. In such a regime, the photometry of SGWs (especially for occultations and for images of the unlit side of the rings) is largely dominated by the mid-τ peak, which simulated movies identify as former wakes in the pro- cess of disruption. The areal average of the mid-τ and low-τ regions can be identified with the τgap values inferred from simpler models, except in the case of very low elevation angle (Tiscareno et al. 2010a), thus preserving the usefulness of the previous UVIS and VIMS studies. Recent preliminary results from a very-high-resolution UVIS occultation appear to give empirical confirmation that the distribution of surface densities in SGW-dominated disk regions is trimodal (Sremcevi´c et al. 2009). -- 33 -- Fig. 20. -- A histogram of local dynamical optical depth τdyn calculated using a local density estimation method for a ring patch containing self-gravity wakes (Tiscareno et al. 2010a). A least- squares fit was made to a sum of three gaussians. The input surface density was 50 g cm−2, and the coefficient of restitution law (see Section 4.2) is that of Borderies et al. (1984) using v∗ = 0.001 cm s−1. Figure modified from Tiscareno et al. (2010a). SGWs may have a dramatic effect on the relationship between ring optical depth and surface density. Simulations by Robbins et al. (2010) found that increased surface density merely added more mass to the already-opaque wakes and only weakly increased the overall optical depth. They estimated that the mass of the B ring may be higher than Voyager -era estimates by a factor of 10 or more, approaching twice the mass of Mimas. See Section 5 for a discussion of the impact of this finding on the age and origin of Saturn's rings. 3.1.5. Propellers A disk-embedded moon that is too small to open a full circumferential gap may still create a local disturbance in the disk. Because of Kepler's Third Law, the radially inward portion of the disturbance is carried forward and leads the moon while the radially outward portion trails the moonlet (Figs. 21 and 22). Due to this characteristic shape, such moonlet-caused disturbances have been named "propellers". A propeller-shaped disturbance can be thought of as a moon's unsuccessful attempt to form a full circumferential gap, which is frustrated by local ring viscous processes that limit the gap's azimuthal extent. Following predictions based on theory and modeling (Spahn and Sremcevi´c 2000; Sremcevi´c et al. 2002; Seiss et al. 2005), propellers in a range of sizes have been observed in Cassini images (Tiscareno et al. 2006a; Sremcevi´c et al. 2007; Tiscareno et al. 2008, 2010c). In all cases, only the propeller-shaped disturbance is directly seen, while the responsible moon at the center is inferred. -- 34 -- Fig. 21. -- Simulated "propeller" disturbance in the A ring due to an embedded moonlet. Note the texture of self-gravity wakes in the unperturbed regions, and the near-horizontal regions of depletion (white) and enhancement (black) in density due to the moonlet. Figure courtesy of M. C. Lewis. Fig. 22. -- Propellers as seen in selected Cassini ISS images. Panel (a) shows a propeller in context of the Encke Gap and several density waves. Panel (b) illustrates the radial offset ∆r between the two azimuthally-aligned lobes, proportional to the size of the unseen central moonlet. Panels (b), (c), and (d) show three views of the same propeller at the same scale, demonstrating how its appearance changes with viewing geometry. Non-equinox views are on the lit (b,e,g) or unlit (a,c,d,f) face of the rings while, panels (h) and (i) show propellers casting shadows near the saturnian equinox. The scale bar in panel (d) also applies to panels (b), (c), (f), and (g). The scale bar in panel (i) also applies to panels (e) and (h). Figure from Tiscareno et al. (2010c). The occurrence of observed propellers is confined to only a few annular regions within the -- 35 -- Fig. 23. -- A histogram of the abundance of propellers, as a function of radius, in the Propeller Belts of the mid-A ring. Figure from Tiscareno et al. (2008). A ring. The three "Propeller Belts" in the mid-A ring are located between 127,000 and 132,000 km from Saturn's center (Fig. 23) and are separated by propeller-poor "halo" regions, centered on large density waves, in which other ring properties are also altered (Fig. 19). It is not known whether the radial variations in observed abundance of propellers are due to variations in their origins, survival, observability, or a combination of the three. The Propeller Belts contain numerous small propellers with the radial offset parameter ∆r (Fig. 22) ranging from 0.3 to 1.4 km and azimuthal extent of up to several km (Tiscareno et al. 2008). A separate class of "giant propellers" has been found in the outermost A ring, outward of the Encke Gap (133,700 km from Saturn's center) and thus called the "trans-Encke" population, with measured ∆r as high as 6 km and azimuthal extent of up to several thousand km (Tiscareno et al. 2010c). Simulated propeller structures have both density-depleted and density-enhanced regions (Fig. 21), and the interpretation of observed propellers in terms of relative density has been a point of con- troversy. In simulated propellers, the radial offset ∆r is consistently close to 4 times the central moon's Hill radius (Eq. 2), and thus is diagnostic of its mass. The first observed propellers, which were seen during Cassini 's Saturn Orbit Insertion (SOI) maneuver and are thus the smallest known but also the best-resolved in the Propeller Belts, are relative-bright with respect to nearby unper- turbed regions of the A ring (Tiscareno et al. 2006a). Although this photometry is consistent with a region of enhanced optical depth, the morphology of the SOI propellers is very similar to that of density-depleted regions in propeller simulations. The photometry of larger propellers in the Propeller Belts also turned out to be generally consistent with enhanced optical depth (Sremcevi´c et al. 2007; Tiscareno et al. 2008). Hypotheses to explain the apparent correlation between high optical depth and depleted density include the temporary liberation of ring-particle regolith within the propeller structure (Sremcevi´c et al. 2007; Halme et al. 2010) and the disruption of self-gravity wakes within the propeller structure (Tiscareno et al. 2010a). In giant propellers, for the first time, relative-dark and relative-bright regions are sometimes seen in the same propeller structure -- 36 -- Fig. 24. -- The feature known as S/2009 S 1, identified by the shadow it cast during equinox, was seen only in this image. Figure from Spitale and Porco (2010). (Fig. 22), enabling density-enhanced and density-depleted regions to be disentangled in some cases (Tiscareno et al. 2010c). However, the structure of giant propellers may be qualitatively different than that of the smaller propellers in the Propeller Belts, so parallels should be drawn cautiously. The giant trans-Encke propellers are larger and more prominent than those in the Propeller Belts, and also much less numerous. Taken together, these factors allow giant propellers to be studied individually and tracked over a period of years. In the Propeller Belts, the swarm of particles is such that, even if the same object were seen on multiple occasions, it would be very difficult to have much certainty that it was the same object due to the many nearby similar objects. This criterion may be useful for drawing a distinction, should one wish to do so, between a "moon" and a "moonlet". Basing such a distinction upon size is arbitrary and lacks any wide agreement, as there is no major physical threshold crossed by objects in the km-size range. We suggest that any object be called a "moon" if it can be singled out for long-term study, while a "moonlet" is a member of a population or swarm that prevents individual tracking. This may still not be a bright line, but at least it's based on a physical property. The question of whether propeller- causing objects deserve to be considered as full-fledged moons is further complicated by the fact that, though their positions have been tracked over long periods, they are not directly seen but hidden within the surrounding propeller-shaped disturbance. The propeller population follows a very steep particle-size distribution, with a power-law index q ∼ 6 (Tiscareno et al. 2008, 2010c) for propeller moonlets of size between 30 m and 300 m. By contrast, the vast majority of ring mass is concentrated in the continuum particles of size between 1 cm and 10 m, with a much shallower power-law index q ∼ 2.75 (Zebker et al. 1985; Cuzzi et al. 2009) -- 37 -- Spitale and Porco (2010) reported a single observation of what appears to be an embedded moon of radius 0.3 km in the outermost parts of the B ring, based on the shadow it cast onto the (vertically much thinner) ring while the 2009 equinox brought nearly edge-on ring illumination (Fig. 24). However, though this object's size is comparable to that inferred for the largest propeller moons, no propeller structure is apparent around this object. Michikoshi and Kokubo (2011) investigated the possibility that the B ring's high surface density, with accompanyingly strong self- gravity wakes, might prevent the formation of a propeller structure. They found that propellers form when the moonlet's Hill radius rHill (Eq. 2) is larger than the Toomre critical wavelength λcr (Eq. 11). While this condition might hold if a moon of the observed size were in the densest parts of the central B ring, the surface densities inferred by Spitale and Porco (2010) for the outer B ring are not high enough. Possible explanations include that the surface densities in the outer B ring are higher than thought, that the observed object is not a moonlet but perhaps an impact cloud or a large SGW clump, that the observed bright feature does indeed include an unresolved propeller structure, or that propellers require a (not well understood) photometric balance in order to be observed and that that balance is not met in the outer B ring. Orbital tracking of trans-Encke propellers has revealed a surprising non-keplerian component to their motion (Tiscareno et al. 2010c). The best-observed example had its semimajor axis increase (as detected by tracking its longitude as a function of time, not its actual radial position) by a rate as high as +0.11 km yr−1 between 2006 and 2007, then decrease by −0.04 km yr−1 from 2007 to 2009 (Fig. 25). Although this profile is similar to a sinusoid, no resonance with a known moon has been found to be capable of producing such behavior. Another mechanism that would produce a sinusoidal residual is the so-called "frog resonance" (Pan and Chiang 2010), in which the propeller moonlet interacts primarily with the mass at either end of the propeller gap. The propeller moonlet plausibly librates with the observed amplitude and period in a simple version of the model, but it remains unclear whether the moon-formed gap responds sluggishly enough to allow the moon to librate within it. Other hypotheses for non-keplerian motion of propellers are based on the concept of "Type I migration." As classically formulated for protoplanetary disks (Ward 1986, 1997; Papaloizou et al. 2007), the angular momentum exchange at inner Lindblad resonances between a disk and an embed- ded mass fails to exactly cancel with that at outer Lindblad resonances, resulting in a differential torque that leads to inward migration of the embedded mass. However, classical Type I migration depends crucially on the gas component of the disk, which causes Lindblad resonance locations to shift asymmetrically. For the case of planetary rings, which are strictly particulate, Crida et al. (2010) re-derived the equations for Type I migration from first principles, using analytical argu- ments and numerical simulations to trace the angular momentum exchange between streamlines of continuum ring particles and the embedded moon. For the case of a homogeneous disk, Crida et al. (2010) found an asymmetric torque that is always inward and is one to two orders of magni- tude too weak to explain the observed non-keplerian motion. Building upon this work, Rein and Papaloizou (2010) considered temporal variations in the disk, proposing that the propeller moon is -- 38 -- Fig. 25. -- Non-keplerian motion of the propeller "Bl´eriot" over 4 years. Panel (a) contains all the data, while panels (b), (c), and (d) contain subsets of the data shown in greater detail. The blue line indicates a linear-plus-sinusoidal fit to all the data, while the red lines indicate piecewise quadratic fits corresponding to a constant drift in semimajor axis and the black lines indicate exponential fits. Figure from Tiscareno (2012). constantly perturbed by stochastic variations in the disk's surface density due to self-gravity wakes, finding that a random walk in the propeller moon's semimajor axis can result. On the other hand, Tiscareno (2012) considered spatial variations in the disk, proposing that an externally-produced radial surface density profile results in an equilibrium semimajor axis to which the propeller moon will return after episodic kicks. Continued observations should distinguish among these models. The leading models inter- pret the existing data as a sinusoid (Pan and Chiang 2010), episodically-initiated exponentials (Tiscareno 2012), and a pure random walk (Rein and Papaloizou 2010), and thus they differ- ently predict the qualitative nature of future data. However, all of the leading models imply that propeller-moons, which are the first objects ever discovered to orbit while embedded in a disk rather than in free space, are directly interacting with the disk, a phenomenon that has long been an integral part of disk models but has never before been directly observed. -- 39 -- 3.1.6. Spokes and impacts "Spokes" are near-radial markings on Saturn's B ring, likely composed of dust levitating above the ring-plane due to electromagnetic forces (for details and references, see Hor´anyi et al. 2009). Spokes usually form on the dawn side of the rings where ring particles are just coming into full sunlight, and have a correlation with periodicities believed to originate with Saturn's magnetic field. Susceptibility to electromagnetic forces surely is connected to the fact that spokes appear at radial locations near to (and often astride) that of synchronous orbit, where the orbital period of a ring particle matches the rotation period of the planet, and thus also of the magnetic field. Spokes appear to be a seasonal phenomenon, correlated with a low elevation angle of the Sun above the ring-plane, probably due to variations in the generation of plasma by photocharging near the rings. The Hubble Space Telescope tracked the decline and disappearance of spokes during "Saturnian October/November" in the late 1990s (McGhee et al. 2005), and they remained absent during the first 1.5 yr of Cassini operations at Saturn before reappearing in "Saturnian January/February" in late 2005 (Mitchell et al. 2006). Since their reappearance, Cassini images have tracked the morphology, photometry, evolution, and temporal variability of spokes (Mitchell et al. 2012). Numerous theories for the formation of spokes have been put forward, but none has gained definitive acceptance. The most popular mechanism is that of Goertz and Morfill (1983), who interpret spokes as dust levitated above the ring plane by an electromagnetic disturbance initiated by a micrometeroid impact. Following the inference of Smith et al. (1982) from Voyager 2 images that spokes with a radial dimension of thousands of km form on a timescale of several minutes, the Goertz and Morfill (1983) model includes a propagation speed for the the plasma cloud (cid:38) 20 km s−1, though this result was criticized by Farmer and Goldreich (2005, see reply by Morfill and Thomas 2005). Other ideas for the rapid appearance of spokes include near-simultaneous impact of a broad assemblage of particles generated elsewhere; the most developed model of this type suggests an electron beam generated by saturnian lightning (Hill and Mendis 1981; Jones et al. 2006), though a dispersed cloud originating from an impact has also been suggested (Hamilton 2006). On the other hand, despite executing a number of high-frequency imaging sequences designed to do so, Cassini has not observed the rapid formation of spokes, but did observe spokes growing from negligible to strong over ∼hr timescales (Mitchell et al. 2012). During the 2009 saturnian equinox, dust clouds evolving under keplerian shear were observed in the A and C rings and attributed to micrometeoroid impacts (Tiscareno et al. 2009b, see also Section 3.6). Ongoing analysis of these impact clouds may lead to new constraints on the inter- planetary micrometeoroid population, as well as better understanding of the relationship between impacts and spokes. -- 40 -- 3.2. Dense narrow rings Saturn: Uranus: Titan ringlet Maxwell ringlet Bond ringlet ("1.470 RS") Huygens ringlet "Strange" ringlet Herschel ringlet ("1.960 RS") Jeffreys ringlet Laplace ringlet ("1.990 RS") 6 ring 5 ring 4 ring α ring β ring η ring γ ring δ ring  ring Dense narrow rings are a unique assemblage of matter, behaving as a coherent self-contained object on planetary lengthscales yet ephemerally thin (1 to 100 km in radial width, compared to ∼ 100, 000 km in diameter) and not in a gravitational ground state (unlike a planet, one would collapse if it stopped moving). The formation and proximate causes of these highly organized dynamical systems are almost entirely unknown, and even their present dynamics are only partly understood. For uranian narrow ringlets, which constitute the majority of the known examples, their highly time-variable properties were only dimly revealed by the single snapshot provided by the Voyager 2 flyby, with temporal resolution provided by Earth-based occultations, while analysis of the more extensive Cassini data set of saturnian narrow ringlets is still in progress. Nearly all known dense narrow rings are either non-circular or inclined to the main ring plane, or both. Their radial widths range from ∼ 1 km (many examples) up to ∼ 100 km for the Maxwell ringlet and the  ring. Furthermore, nearly all dense narrow rings have edges that are quite sharp; as with dense disks, the existence and stability of such sharp edges requires some confinement mechanism to counteract the natural process of radial viscous spreading. Confinement may be due to an external moon or to the ring's own self-gravity, and in some cases to processes yet to be understood. A detailed table of ringlets and their properties was given by Colwell et al. (2009b) for Saturn's rings and by French et al. (1991) for Uranus'. For general details and references on the dynamics of narrow rings, see French et al. (1991) and Schmidt et al. (2009). The "shepherding" mechanism by which a moon opens a gap or maintains an edge in a disk (Section 3.1.2) can also occur with two shepherd moons on either side of a narrow ringlet (Goldreich -- 41 -- and Tremaine 1979a). As previously discussed, this can occur through repeated impulses on nearby material or more distantly through a resonance. The only example of a dense narrow ring with a known shepherd on either side is Uranus'  ring, though Saturn's F ring (see Section 3.3) is a variation on that idea. The Titan ringlet in Saturn's C ring is at the location of an apsidal resonance with Titan, where the ring particle's precession frequency  is commensurate with Titan's orbital motion, so that the eccentric ringlet always keeps its apoapse pointed towards Titan. The Bond ringlet in Saturn's outer C ring is associated with a 3:1 Lindblad resonance with Mimas, and a few edges of uranian rings coincide with resonances, but the details of the interaction are yet to be understood in all cases. Any ringlet of finite width ought to have a different precession rate  at its inner and outer edges, which should smear out the orientation of ring particle orbits and prevent the ringlet from appearing eccentric. However, this effect can be counteracted by the ring's own gravity (Goldreich and Tremaine 1979b, 1981; Borderies et al. 1983), possibly combined with viscous and collisional effects (Dermott and Murray 1980; Chiang and Goldreich 2000; Mosqueira and Estrada 2002). A purely gravitational model requires a positive "eccentricity gradient," which is to say that the eccentricity monotonically increases from the ring's inner edge to its outer edge. A number of ringlets indeed appear to precess about their planet as a rigid body. Observations of some of these, such as the Maxwell ringlet and the α and β rings, are consistent with a pure freely-precessing ellipse (i.e., an m = 1 mode) as described by theory, and the Maxwell ringlet even has a clearly positive eccentricity gradient (Spitale and Porco 2006), while other ringlets appear to have additional components to their motion. The Huygens ringlet appears to have an additional m = 2 mode, possibly influenced by the Mimas 2:1 resonance that governs the nearby outer edge of the B ring, as well as an m = 6 mode of unknown origin (Spitale and Porco 2006). The δ ring also has an m = 2 mode, while the γ ring has an m = 0 mode, which is to say a radial oscillation (French et al. 1991). Higher-m modes, some corresponding to known moons, have also been found in several Uranian rings by Showalter (2011), who also pointed out that non-detections of shepherd moons at Uranus have become significant enough that shepherding is unlikely to be the dominant mechanism of ring confinement. Spiral density waves (Section 3.1.1) can also occur in dense narrow rings, at least those broad enough (generally a few km) for a wave to develop, but not many examples have been identified. A density wave due to the Pandora 9:7 LR appears to propagate through the Laplace ringlet (Colwell et al. 2009b). A possible density wave was detected in the Voyager stellar occultation of Uranus' δ ring, but it cannot be confirmed in the absence of data at multiple longitudes and/or times, and furthermore there is no known moon at the proper place to raise such a wave. In Saturn's rings, gaps are named by the IAU but ringlets are not. In most cases, the main ringlet within a particular gap is given the same name as its gap (though many favor the name "Titan ringlet" for the ringlet in the Colombo Gap, for its strong association with a resonance with Titan), but the existence of multiple ringlets in one gap requires some naming creativity. For -- 42 -- example, the Huygens Gap contains a second dense narrow ringlet outward of the main Huygens ringlet, which has been informally nicknamed the "Strange" ringlet in part because of its unusually high inclination (Spitale et al. 2008) and in part as a complement to the "Charming" ringlet (Section 3.3). Five ringlets with non-circular features were identified in Voyager -era publications (e.g., French et al. 1993) only by their distance from Saturn's center in units of Saturn radii; of these, three can now be easily named for the gap containing them as shown in the accompanying table. However, the former "1.495 RS ringlet" is classified by Colwell et al. (2009b) as a plateau or embedded ringlet because it is adjacent to the continuum C ring rather than fully contained in the Dawes Gap, while the former "1.994 RS ringlet" is now considered part of the continuum Cassini Division between two narrow gaps, rather than a ringlet that nearly fills its gap, judging from the characteristic pattern of empty gaps having circular outer edges and resonant inner edges (Hedman et al. 2010a, see also Section 3.1.2). 3.3. Narrow dusty rings Saturn: Uranus: Neptune: "Charming" ringlet Encke ringlets F ring λ ring Le Verrier ring Arago ring Adams ring The smallest particles in dense rings are usually swept up by larger ones and incorporated into their regolith (Cuzzi et al. 2009), and so such rings are largely dust-free, which is to say that they have few particles smaller than mm- to cm-size and thus are not strongly forward-scattering in their interactions with light. Therefore, the dustiness of a few structures that do occur within Saturn's and Uranus' main rings is a likely indicator of relative dynamism and/or youthfulness that either prevents dust from being swept up or has not allowed time for it to be swept up yet. In Neptune's rings, on the other hand, the lack of any ring component with optical depth exceeding 0.1 may account for the general dustiness of the system. In Saturn's main rings, high dust fractions are found exclusively in a small number of narrow dusty ringlets that occur in the larger empty gaps (Hor´anyi et al. 2009). The only gap with multiple dusty ringlets is the Encke Gap, which three ringlets share with the moon Pan. These are riddled with "clumps" (azimuthal brightness variations) and "kinks" (radial offsets) that drift slowly with respect to each other (Hedman et al. 2005, 2011b; Hor´anyi et al. 2009). The Encke ringlets and the "Charming" ringlet within the Laplace Gap are known to be "heliotropic," which is to say that they have an eccentricity forced by solar radiation pressure that causes their apoapses to always point towards the Sun (Hedman et al. 2007b, 2010b). The "Charming" ringlet is smooth, lacking clumps -- 43 -- Fig. 26. -- Sheared channels in the F ring created as Prometheus (lower right) dips into the ring. The horizontal dimension of this Cassini ISS mosaic covers 60◦ of longitude (∼150,000 km) and the vertical (radial) dimension is 1,500 km. Figure from Murray et al. (2005). or kinks; the best-studied of the heliotropic rings, it also has free eccentricity and free inclination components in addition to its solar-forced eccentricity (Hedman et al. 2010b). The F ring is the granddaddy of narrow dusty ringlets, being by far the best studied as well as the largest. For a review and references, see Colwell et al. (2009b). Located a few thousand km off the outer edge of Saturn's main rings, and with both significant vertical thickness and inclination ((cid:38) 10 km in both cases), the F ring effectively frustrates any attempt to see the rest of the main rings in edge-on viewing geometries. Its high fraction of forward-scattering dust makes it by far the brightest component of Saturn's ring system when viewed at high-phase geometries. The core of the F ring contains a large amount of dust enveloping an unseen belt of km- size moonlets, inferred from their absorption of charged particles (Cuzzi and Burns 1988), from the characteristic "fan" structures they create in surrounding dust (Murray et al. 2008; Beurle et al. 2010), from direct detection by occultations (Esposito et al. 2008; Hedman et al. 2011c), and from shadows cast during the 2009 saturnian equinox (Beurle et al. 2010). Several dusty lanes or "strands" accompany the core on either side, nearly parallel to it though Charnoz et al. (2005) pointed out that the most prominent strands can be laid end-to-end to form a one-armed kinematic spiral. Nascent strands, or "jets," have been associated with collisions between moonlets and the core (Murray et al. 2008). Despite being entirely composed of "clumps" and "kinks" so numerous that they cannot be individually tracked as they are in the Encke ringlets, the F ring core nevertheless maintains over decadal timescales the shape of a freely precessing eccentric inclined ellipse; the orbital solution formulated to account for Voyager and other pre-Cassini data (Bosh et al. 2002) has, somewhat surprisingly, remained a good predictor of the core's position through the Cassini mission (Murray et al. 2008; Albers et al. 2012). However, decade-scale time variations in the core's internal structure, as well as in the surrounding dust, have clearly taken place between the Voyager and Cassini visits (Colwell et al. 2009b; Showalter et al. 2009). The F ring is one of only two narrow rings (along with the  ring) to have known "shepherd" moons orbiting on either side of it. However, it has not been conclusively shown how (or even -- 44 -- that) the moons Prometheus and Pandora actually constrain the F ring in its place, and in fact they appear to stir the ring up at least as much as to maintain it. Simulations by Winter et al. (2007) found the moons to be responsible for both strong chaos and significant radial confine- ment. Both moons create at each closest approach a "streamer-channel" in nearby dust strands that subsequently moves downstream and shears under keplerian motion (Fig. 26, Murray et al. 2005). During the 2009 saturnian equinox, which coincided with the once-in-17-yr alignment12 of Prometheus' apoapse with the F ring's periapse, minimizing their mutual closest approach dis- tance (Chavez 2009), moonlets inferred from their shadows had a clear correlation of abundance with longitude relative to Prometheus (Beurle et al. 2010), indicating that Prometheus is directly triggering accretion within the F ring, the products of which may then be what collides with the core to form new jets and strands (Beurle et al. 2010). All of these interwoven phenomena make the F ring the solar system's foremost natural laboratory for direct observation of accretion and disruption processes. Like the F ring, the λ ring is by far the brightest component of its planetary ring system when viewed at high phase angles, due to its high fraction of forward-scattering dust. But the λ ring appears to be significantly simpler and more sedate than its saturnian cousin, and its low detectability at lower phase and in occultations indicates that it is poor in macroscopic particles. Among dense ring systems, Neptune's has a much higher dust fraction than those of Saturn and Uranus, and is unique in having no significant dust-free regions. The three main rings of Neptune lack sharp edges and are generally more tenuous, with only the Adams ring reaching optical depths even as high as 0.1, and are consequently less well observed. The only post-Voyager observations, other than of the Adams ring and its arcs (see Section 3.5), are a marginal detection of the Le Verrier ring reported by Sicardy et al. (1999), which if confirmed would require it to be brighter than expected from Voyager data, and a clear detection by de Pater et al. (2005) that found the Le Verrier ring's brightness to be consistent with Voyager measurements. Possible explanations include temporal brightening in the Le Verrier ring followed by a return to its previous state, or unexpected spectral properties, or (a possibility they admit) that the Sicardy et al. (1999) detection was affected by image artifacts. 12The periodicity of this alignment was recalcluated by Chavez (2009), using updated orbital data. -- 45 -- 3.4. Diffuse dusty rings Jupiter: Saturn: Uranus: Neptune: Halo ring Main ring Amalthea Gossamer ring Thebe Gossamer ring D ring Roche division Janus/Epimetheus ring G ring Methone ring Pallene ring Anthe ring E ring Phoebe ring ζ ring ν ring µ ring Galle ring Lassell ring Every known ring system has a diffuse component that is optically thin and composed of µm- size dust particles. The dynamics and evolution of diffuse dusty rings are made more complex by the importance of forces other than gravity (Burns et al. 1979). Dust particles are commonly low enough in mass that static electrical charging makes them susceptible to electromagnetic forces comparable to gravity. They also have high ratios of surface area to volume, which makes them susceptible to pressure from solar radiation, including Poynting-Robertson drag. The evolution induced by these additional forces shortens the lifetimes of dust particles so that the stability of diffuse dusty rings is only of a dynamical variety. The ring consists of particles that originated from a source and are on their way to a sink; it may indeed look the same at a different time, if the sources and sinks have not significantly changed, but the particles comprising it will be different. A comprehensive review of dusty ring systems was published by Burns et al. (2001). The primary mechanism for producing orbiting dust particles is micrometeoroid bombardment of larger orbiting bodies. The rate at which a moon of radius R supplies mass to a dusty ring is given by (Burns et al. 1999) dM dt ∼ ΦY FeR2, (12) where Φ is the impactor flux, Y is the yield of ejected mass as a fraction of projectile mass, and Fe is the fraction of ejected mass that achieves planetary orbit. The yield Y depends on the material properties of the moon's regolith, while the fraction Fe of ejecta that can escape the moon's gravity -- 46 -- Fig. 27. -- Supply rate to a dusty ring via impact ejecta (dM/dt) is plotted in black, as a function of source moon radius R. This quantity is the product (Eq. 12) of the total ejecta produced (ΦY R2, plotted in blue in arbitrary units) and the fraction of ejecta that escapes the moon's gravity and achieves planetary orbit (Fe, plotted in red). Moons smaller than Rcrit lose all of their ejecta, while larger moons produce more ejecta but allow a smaller fraction of it to escape. The optimal moon radius for supplying dust to a ringlet varies with surface properties, but is Rcrit ∼ 10 km for icy moons with soft regolith (Burns et al. 1999). depends on the moon's escape velocity and the distribution of ejecta velocities. Empirically, the fraction of ejecta with velocity higher than v goes as (vcrit/v)9/4, where vcrit is the minimum speed at which material is ejected, between 10 and 100 m s−1 where lower values are for softer moon's regoliths (Burns et al. 1984, 1999). The escape velocity of a spherical moon with bulk density ρ goes as vesc ∝ ρ1/2R and also ranges between 10 and 100 m s−1. So for vesc < vcrit we have Fe = 1 and dM/dt ∝ R2, but when vesc > vcrit we have Fe ∝ R−9/4 and dM/dt ∝ R−1/4. That is, there turns out to be an optimal moon radius Rcrit ∼ 10 km for supplying dust to a ringlet, with smaller moons intercepting fewer impactors and thus producing less dust, while larger moons allow a smaller fraction of the produced dust to escape (Fig. 27). However, these trends are only for general estimation. A more detailed treatment must consider the mechanics of ejecta production, including not only surface mechanics but also the changes in the ejecta velocity profile with increasing R as impact formation moves from the strength regime to the gravity regime (e.g., Richardson et al. 2007). Also, larger moons are more efficient at sweeping up escaped particles during subsequent encounters (Agarwal et al. 2008; Kempf et al. 2010). All four known planetary ring systems appear to have an optically thin dusty ring as their inner- most component -- Jupiter's Halo ring, Saturn's D ring, Uranus' ζ ring, and Neptune's Galle ring. -- 47 -- Dust rings often extend inward of their sources because Poynting-Robertson drag in particular enforces an inward evolution of dust particles, as does resonant charge variation inward of syn- chronous orbit (Burns et al. 2004). Both Jupiter's Halo/Main ring and Saturn's D ring have likely source material at their outer edges (namely, Metis and Adrastea, and the C ring, respectively), although at least the D ring has internal radial structure (Hedman et al. 2007c) that may also re- quire embedded sources (i.e., undiscovered moons). The sources of the poorly observed ζ ring and Galle ring are not known, and may also include embedded moons very close to the planet. Some of these dust sheets may extend all the way down to the planet's cloud tops, although Saturn's D ring appears to have a clear gap of 5,000 km between its inner boundary and the cloud tops (Hedman et al. 2007c), through which the Cassini spacecraft is slated to fly in 2017 during its end-of-mission maneuvers (Seal and Buffington 2009). Models of Jupiter's rings also find an empty region above the cloud tops, through which the Juno mission plans to fly in 2016, though this has not been confirmed with definitive observations. Outward movement of dust is possible, and has been invoked to explain an extension of the Gossamer ring beyond the orbit of Thebe (Hamilton and Kruger 2008). The mechanism proposed for outward movement is a "shadow resonance" in which dust particles moving through their planet's shadow temporarily lose their electrical charge and are carried outward by the momentum of their electromagnetically-influenced orbits. In other locations, such as Saturn's G ring, dust evolution is primarily outward because of "plasma drag" from abundant charged particles co- rotating with the planet's magnetic field, which orbit at speeds much faster than the dust's keplerian velocity since the G ring is far outward of synchronous orbit (Hedman et al. 2007d). Not all dusty rings consist of ejecta from a single dominant source moon. Several are tenuous extensions of nearby dense rings, such as the D ring, the dusty sheet in the Roche division, and the Lassell ring, which all lie inward of their likely sources, respectively, the C ring, the F ring, and the Arago ring. The Phoebe ring may also be derived from multiple sources, as it is difficult for models to account for the ring's mass from impacts onto Phoebe alone because Phoebe is so large that it should retain much of its ejecta. However, although Phoebe itself has no known collisional family, disruptive impacts are known to have broken other large irregular satellites into pieces that are separately observed today, and it is quite plausible that km-size pieces have been ejected from Phoebe and continue to share similar orbits. Since, as shown above, km-size moons are actually the most efficient sources of orbiting dust, the observed dust densities can be explained by such a distributed population of source bodies (Verbiscer et al. 2009). The Uranian rings had a very different look (Fig. 28) when the Uranian equinox (an event occurring every 42 years) was observed in 2007 by the Hubble Space Telescope (de Pater et al. 2007). With the Sun and the Earth on opposite sides of the Uranian ring plane, the brightest features in this rare view of the unlit side of the rings had relatively low optical depth but macroscopic particles. The dense  ring practically disappeared because its high optical depth made it opaque, while the dusty λ ring was also dim due to the low phase angle. The brightest feature was at the location of the η ring, and can probably be identified with a low-τ "extension" previously seen -- 48 -- Fig. 28. -- Radial brightness profiles for Uranus rings from Voyager in 1986 and from the Hubble Space Telescope near (2004) and at (2007) the Uranian equinox. Figure from de Pater et al. (2007). adjacent to that ring. Some of the dusty rings seen at high-phase had bright counterparts in HST's unlit-side view, as did a similar "extension" to the δ ring, while others did not. Thus, the latter are likely pure dust rings while the former contain a component of larger particles. A broad innermost ring, which had been given the provisional designation 1986 U2R based on Voyager images, was also seen in HST's unlit-side view and named the ζ ring. However, the core of the ζ ring as seen by HST is several thousand km outward of its position in Voyager images. While acknowledging the possibility that overlapping particle populations with different optical properties would explain both the mid-phase Voyager observations and the HST data, de Pater et al. (2007) argue it is more likely that the ζ ring has changed dramatically in the intervening 20 years. While most dusty rings are either gray or red in their spectral properties, two known rings have a prominent blue color: Saturn's E ring and Uranus' µ ring (Fig. 29). Each of these rings is centered on the orbit of a moon -- Enceladus and Mab, respectively (de Pater et al. 2006). For rings made of µm-sized particles, which are close in size to the wavelengths of visible and infrared light, a blue -- 49 -- Fig. 29. -- A schematic view of the outer rings of Saturn and Uranus, in which each system has been scaled to a common planetary radius. Highlighted are the red coloration of Saturn's G ring and Uranus' ν ring, both of which appear to be typical dusty rings, and the unusual blue coloration of Saturn's E ring and Uranus' µ ring. Figure from de Pater et al. (2006). spectral trend is best explained by a particle-size distribution (regardless of particle composition) that is either narrowly centered on a particular size or characterized by a very steep power law (Showalter et al. 1991; de Pater et al. 2004). For Enceladus and the E ring, the mechanism by which this happens seems well understood: ice particles are spewed from Enceladus south-polar geysers at a variety of speeds, with smaller particles being more easily accelerated by gas in the plume, leading to higher velocities that enable them to join the E ring (Schmidt et al. 2008; Hedman et al. 2009a; Kempf et al. 2010). However, it's harder to imagine this mechanism applying to the µ ring. Enceladus, with a diameter of ∼ 500 km, is already so small that the source of sufficient internal heat to account for its plume is a matter of much debate (e.g., Meyer and Wisdom 2007, 2008a,b), and Mab has approximately 1/10 the diameter (and thus 1/1,000 the mass) of Enceladus. A satellite the size of Mab ought to be a good source of dust liberated by collisions (see above) but that should lead to shallower size distributions and a redder spectral trend like those of most dusty rings. -- 50 -- 3.5. Ring arcs and azimuthal clumps Arcs Saturn: Neptune: G ring Methone ring arc Anthe ring arc Galatea ring arc Adams ring Azimuthal clumps Jupiter: Saturn: Main ring Encke Gap ringlets F ring Neptune's iconic ring arcs, known since Voyager and strongly suspected even earlier from occultation data, have now been joined by several Saturnian structures among the ranks of ring arcs. All share the common characteristic of an azimuthally confined region of enhanced brightness embedded within a fainter circumferential ring.13 All known arcs orbit at the appropriate keplerian rate for their distance from planet center. But there are also significant differences among known arcs, as they cover a wide range of densities and are caused by at least two different mechanisms. Left to itself, any clump of material orbiting a planet should spread out into a ring on a fairly short timescale. This is a direct result of keplerian shear (Eq. 9), by which two objects with semimajor axes a and a + δa will have mean motions n and n + δn, where δn = −(3n/2a)δa, and the time for one to "lap" the other by an entire orbit is Tspread = = − 2π δn 4π 3(GM )1/2 a5/2 δa , (13) where we have used the precise form of Kepler's Third Law, n2a3 = GM . For the giant planets, (GM )1/2 is of order 104 km3/2 s−1, and a typical distance from the planet is a ∼ 105 km. Therefore, even for a compact initial clump δa ∼ 1 km, the spreading time is only a few decades and is inversely proportional to δa. This exercise demonstrates that azimuthal variations in a ring's mass are not intrinsically stable. Therefore, the several examples that exist of observed ring arcs must be dynamically generated or maintained. The two most prominent mechanisms for this are resonant confinement and asymmetric injection of mass into the ring. 13For Saturn's Anthe and Methone arcs, as for Neptune's Galatea arc, the circumferential ring is too faint to have been detected as yet but likely consists of material recently escaped from the arc-confining mechanism. -- 51 -- Corotation resonances have resonant arguments (see Section 3.1.1) of the form ϕ = (m + k)λ(cid:48) − mλ − kX(cid:48), (14) where, as for Eq. 6, m and k are integers that label the resonance as (m+k):m, primed quantities refer to the forcing moon, and X is either  or Ω. For a corotation eccentricity resonance (CER), X is  and the resonance strength is proportional to the perturbing moon's eccentricity, while for a corotation inclination resonance (CIR), X is Ω and the resonance strength is proportional to the perturbing moon's inclination. Corotation resonances differ from their Lindblad and vertical cousins (see Section 3.1.1) in that the lowest-order resonances involve the eccentricity or inclination of the perturbing moon, rather than that of the particle being perturbed. Thus, rather than pumping up the eccentricities and inclinations of ring particles and driving waves, corotation resonances tend to azimuthally confine particles into an orbit commensurate with that of the perturbing moon. This mode of confinement was first proposed for Neptune's ring arcs by Goldreich et al. (1986). 3.5.1. Neptune's Adams ring The first-discovered and best-known set of ring arcs are found in Neptune's Adams ring. These are the densest components of Neptune's ring system (with τ⊥ ∼ 0.1) and were oftentimes the only component detectable in pre-Voyager Earth-based occultations, leading to much confusion as to whether the detected signatures were rings at all until Voyager 2 settled the question. There are five arcs occupying ∼ 20◦ out of a region extending over ∼ 40◦ of longitude (Fig. 9 and 30). The three main arcs were named for the French revolutionary slogans Libert´e, Egalit´e, and Fraternit´e, then closer inspection showed a bifurcation in Egalit´e and a dimmer fourth arc that was named Courage. The first integrated dynamical explanation of the arcs was given by Porco (1991), who com- bined the theoretical framework of Goldreich et al. (1986) with observations showing that the Adams ring lies very close to a 43:42 resonance with the moon Galatea. Porco (1991) suggested that the azimuthal confinement mechanism was due to Galatea's 43:42 corotation inclination res- onance (CIR) while radial spreading due to particle collisions was prevented by Galatea's nearby 43:42 outer Lindblad resonance (OLR). This explanation requires both Libert´e and Egalit´e to stretch over multiple consecutive corotation sites (each of which is only14 360◦/86 = 4.2◦ long) and predicted that the arcs would orbit at the pattern speed of the 43:42 CIR. A synthesis of all Voyager and Earth-based data (Nicholson et al. 1995) found two possible solutions for the arcs' pattern speed, one of which was consistent with the Galatea 43:42 CIR, and detailed dynamical 14Because inclination resonances cannot be first-order (see, e.g., Section 10.3.3 of Murray and Dermott 1999), the CIR actually functions as an 86:84 resonance, which is why the number 86 appears here. For the CER discussed by Namouni and Porco (2002), the corotation sites would be twice as long. -- 52 -- Fig. 30. -- (upper) Reprojected ground-based image of the Adams ring (with arcs) and the Le Ver- rier ring acquired in October 2003. (lower) Azimuthal profiles of the Adams arcs from four separate observations. The leading arc Libert´e was seen by Voyager 2 (black dashed line) to be as bright as the other two main arcs, but subsequently dimmed. Figure from de Pater et al. (2005). See Fig. 9 for schematic view. work (Foryta and Sicardy 1996; Hanninen and Porco 1997) supported the plausibility of resonant confinement by this mechanism. However, the re-acquisition of the arcs with Earth-based imaging (Dumas et al. 1999; Sicardy et al. 1999) called the resonant pattern speed into question, favoring instead Nicholson et al. (1995)'s other solution. An attempt to salvage the Galatea model was made by Namouni and Porco (2002), who proposed an alternative model employing the 43:42 corotation eccentricity resonance (CER) rather than the CIR. They furthermore invoked the mass of the ring arcs themselves to adjust the resonant pattern speed to match the observations, thus deriving a -- 53 -- value for the arcs' mass that is required for their model to work. Further Earth-based observations (de Pater et al. 2005) not only confirmed that the arcs are moving at the "wrong" pattern speed for the original Porco (1991) model, but showed significant changes in the arcs' structure over the 20 years in which they have been observed in detail (Fig. 30). The brightness of Libert´e, already diminished in the 1998 observations, had further declined until it was dimmer than Courage, while Courage had moved forward in longitude relative to the other arcs. Egalit´e has also undergone less dramatic shifts in its morphology and longitude, while Fraternit´e appears largely stable. The Adams arcs remain an enigma. While the model invoking resonant confinement by Galatea has appeared promising, the details have not come together as hoped. The original model by Porco (1991) was appealing because its predicted pattern speed appeared to match very closely with the observations; however, the re-working of the model by Namouni and Porco (2002) is less convincing because it requires the invocation of an additional free parameter in order to fit the data. Furthermore, now that other examples of resonantly confined ring arcs have come to light (see Section 3.5.3), all of which have the morphology predicted by Goldreich et al. (1986), with maxima at the center of a corotation site and brightness falling off long before the site's edges are reached, the objection can be raised anew that Egalit´e and Fraternit´e span multiple corotation sites while showing no clear evidence of internal minima at the expected spatial frequency. On the other hand, the azimuthal structure in Jupiter's Main ring (Section 3.5.2), though far less well-observed, appears similar to the Adams arcs in that the spatial frequencies do not match those of nearby corotation resonances. Perhaps the corotation-resonance model will be vindicated in the end, or perhaps the answer will be more like the shepherding model of Lissauer (1985), which invokes yet- undiscovered moons embedded within the Adams ring, or perhaps the Adams arcs will be found not to be long-term stable structures. 3.5.2. Jupiter's Main ring and other azimuthal clumps In the core of Jupiter's Main ring, New Horizons images found one close pair of azimuthal clumps and another family of three to five clumps (Showalter et al. 2007). The semimajor axes of the α and β clump families, measured from their orbital rates, fall less than 1 km from, respectively, the Metis 115:116 and 114:115 corotation inclination resonances (CIR), the same kind of resonance originally invoked by Porco (1991) for Neptune's Adams arcs. Given the distance between the two semimajor axes and the distance between the resonances, Showalter et al. (2007) estimate the probability of this happening by coincidence to be 4%. Finally, to complete the analogy with Neptune's Adams arcs, the 1.8◦ azimuthal spacing of the clumps does not correspond to the expected 360◦/230 = 1.56◦ for these resonances (Showalter et al. 2007). The Adams ring and the Main ring may turn out to be more like Saturn's F ring and the ringlets in the Encke Gap (Section 3.3), which have rich azimuthal structure that is clearly not associated with a resonant spatial frequency and is most likely due to embedded source moons. -- 54 -- 3.5.3. Saturn's G ring and other moon-embedded arcs Saturn's G ring was, for a long time, the ring that should not be there. Saturn's main disk was clearly a long-term stable structure, the D ring clearly derived from it, the E ring associated with Enceladus, and the F ring confined by Prometheus and Pandora. Yet the G ring, composed of dust grains that cannot persist over long time periods, had no apparent mechanism for its needed continuous generation. The first step towards addressing this question came with the discovery of a relatively bright arc within the G ring (Hedman et al. 2007d). The arc is brighter than the rest of the G ring by a factor of several (but still at τ⊥ ∼ 10−6), and is radially much narrower (∼ 250 km). It is very plausible that the rest of the G ring is composed of grains evolving away from the arc under the influence of non-gravitational forces. Furthermore, the arc's orbital rate matches the 7:6 corotation eccentricity resonance (CER) with Mimas, again invoking the Goldreich et al. (1986) mechanism. The azimuthal profile of the arc is roughly triangle-shaped (Fig. 31), consistent with a source body with only a small-amplitude libration about the exact resonance, but has broad wings that plausibly fill the corotation site's length of 360◦/7 = 51◦ in longitude. A further piece of the puzzle came with the discovery of Aegaeon (Hedman et al. 2010c), a 1-km moon orbiting within the G ring arc. Aegaeon is probably only the largest of a population of source objects within the arc, as electron-absorption measurements indicate massive objects spread over a distance much larger than Aegaeon itself (Hedman et al. 2007d). Two other arcs, even more tenuous than the G ring's, surround the small moons Anthe and Methone, both situated between Mimas and Enceladus. A third moon in this vicinity, Pallene, appears to have a very tenuous ring but not an arc. The Methone arc was first detected by charged- particle absorptions (Roussos et al. 2008), and all three structures were then seen in imaging data (Hedman et al. 2009b). Both Anthe and Methone are in resonance with Mimas, in the 11:10 and 15:14 CERs, respectively, and these resonances continue to confine the arc material once it has left the source moons (Agarwal et al. 2009). The azimuthal length of the Anthe and Methone arcs are both approximately half that of the available corotation site (Hedman et al. 2009b), but that may simply reflect the point at which the faint signal falls below detectability. Because these ring arcs at Saturn are so much more tenuous than Neptune's, they are essentially collisionless. This allows them to be confined by the corotation resonance alone, which is necessary because the associated Lindblad resonances are radially farther away than they are for Neptune's case, both because of Saturn's higher J2 compared to Neptune and because the resonances have lower azimuthal parameter m, and thus are not available to perform the radial confinement that was an essential part of the Goldreich et al. (1986) model for the collisional Adams ring. Finally, we note that Voyager 2 imaged a faint unnamed ring sharing the orbit of the moon Galatea, which may be an arc given its intermittent detectability (Showalter and Cuzzi 1992; Porco et al. 1995). Although little is known about this structure, it can now be placed in context with -- 55 -- Fig. 31. -- (top) Images and (bottom) azimuthal profile of the G ring arc, centered on the Mimas 7:6 corotation site. The peaky shape in the azimuthal profile indicates a significant fraction of arc material is tightly bound to the resonance (i.e., has low libration amplitude). Figure from Hedman et al. (2007d). the Anthe and Methone arcs and may very well be driven by similar mechanisms. -- 56 -- 3.6. Rings as detectors Planetary rings have shown themselves to be useful, in many cases, as detectors of planetary processes around them. Spiral structures in the D ring, and in the similar tenuous dusty sheet in the Roche Division, are driven by Lindblad resonances with the rotation period of the planet's magnetic field. These structures are only seen in these tenuous sheets populated by tiny grains that are easily charged and thus subject to electromagnetic forces. The multiple pattern speeds required to explain the resonant structures are a major source of information for the complex rotation of Saturn's enigmatic magnetic field (Hedman et al. 2009c). Evidence for vertical corrugations in Jupiter's Main ring were first seen in Galileo images (Ockert-Bell et al. 1999; Burns et al. 1999), but were not well-understood. Next, Cassini images of Saturn's D ring showed evidence of a vertical corrugation whose radial wavelength is decreasing with time, a trend easily accounted for with a model of differential precession that begins with the ring as an inclined flat sheet ∼ 20 yr before Cassini 's arrival at Saturn (Hedman et al. 2007c). Images obtained during Saturn's 2009 equinox showed the vertical corrugation pattern extending far into the C ring, and the variation of the wavelength with radius confirms the previous result that the pattern originated in 1983 when an event of some kind caused the ring to be tilted by ∼ 10−7 radians (a few meters) with respect to the Laplace plane (Hedman et al. 2011d). Finally, similar analysis of the corrugations in Jupiter's Main ring yields a superposition of wavelengths implying two tilting events, one in July 1994 and one in 1990 (Showalter et al. 2011b). July 1994 is, of course, the date of the impact of comet Shoemaker-Levy 9 (SL9) into Jupiter, leading to the hypothesis that both corrugation patterns carry records of the impact of a spatially-dispersed cloud, which is one mechanism for spreading the impulse over a large portion of the ring. The cloud of dust surrounding SL9 would likely fill the bill at Jupiter, and either a similar cometary system or a meteor stream could be the cause of the Saturn event. Saturn's rings showed themselves capable of more direct detections of impactors during the 2009 equinox event. Bright markings, canted with respect to the azimuthal direction, were seen on both the A ring and the C ring during the few days before and after the Sun's passage through the ring mid-plane (Tiscareno et al. 2009b). Assuming that these are dust clouds evolving under keplerian shear (Eq. 9), one can calculate the time elapsed since the cloud was radially aligned. In one case, the same cloud was seen 24 hours apart, and the ages derived from keplerian shear differed by the same interval, confirming this interpretation of the observed structures. Also, in very high-phase high-resolution images of the C ring (not during equinox), small streaks appear that are probably transitory dust clouds produced by impacts (Tiscareno et al. 2009b). Both of these discoveries have the potential to use rings observations to constrain the influx of interplanetary -- 57 -- impactors, though the derivation of impactor size from observed dust-cloud parameters has yet to be worked out. Spiral waves and wavy gap edges are both the result of gravitational forcing due to moons. In the case of Janus and Epimetheus, the nature of the forcing changes with time, and the ring maintains a record of that change. Initial steps towards understanding the nature of the rings' record have been made for both density waves (Tiscareno et al. 2006b) and wavy edges (Spitale and Porco 2009). In the case of spiral density waves, the group velocity at which information propagates through the wave is slow enough that information can be gained about the state of the co-orbital moons as much as ten years prior to the time of observation. 4. Experimental rings science 4.1. Numerical simulations Advances in simulation techniques, alongside advances in computing hardware and software, have allowed numerical simulations to become an increasingly important part of the study of plan- etary rings. Modern n-body dynamics exploded after Wisdom and Holman (1991) published a symplectic mapping that uses Hamiltonian methods to calculate perturbations as a deviation from keplerian orbit, rather than as deviations from motion in a straight line, and thus requires fewer correction events per unit of simulated time. This, along with ever-increasing computer power, allowed large n-body simulations to become routine; among the many solar-system applications is the case of dusty rings evolving under gravity and other forces. For dense rings, though, the number of mutually interacting particles is too large for simulations that follow particles along their full orbits about the planet. Instead, for cases that do not focus on large-scale azimuthal structure, a very productive line of simulations follows a relatively small "patch" of the ring with sliding boundary conditions as devised by Wisdom and Tremaine (1988). The patch is surrounded by "mirror" patches (Fig. 32); those on either side in the azimuthal direction are stationary, so that a particle that leaves the patch on one side simply reappears on the opposite side, but the neighboring patches in the radial direction slide past according to keplerian shear (Eq. 9). Even with their greatly reduced spatial dimensions, ring-patch simulations routinely include so many mutually-interacting particles that they push the limits of the available computing power, and several approaches to efficiently accounting for their mutual interactions have been devised. The first great success of ring-patch simulations, and a continuing area of their usefulness, is the characterization of self-gravity wakes (SGWs) (Section 3.1.4). First described by Julian and Toomre (1966) for the case of galaxies, it was understood by the 1980s that this non-axisymmetric structure due to a balance between accretion and disruption was likely present in Saturn's A and B rings and was likely responsible for the observer-centered quadrant azimuthal asymmetry observed in the A ring's brightness (Colombo et al. 1976; Franklin et al. 1987; Dones and Porco 1989). SGWs were -- 58 -- Fig. 32. -- Schematic representation of a ring-patch simulation with sliding boundary conditions. The simulation cell (center, with green particles) is replicated on all sides, with the replicant cells (with yellow particles) positioned according to relative keplerian velocities (Eq. 9). In this representation, increasing radius (+x) is to the right and keplerian orbital motion (+y) is up. Figure from Perrine et al. (2011). successfully produced in ring-patch simulations by Salo (1992, 1995) and by Richardson (1994), and simulations continue to be a vital tool for understanding the mechanics and structure of SGWs and comparing their simulated photometric properties to observational data, as well as other ring properties including the rotational states and thermal properties of ring particles, and the mechanics of sharp edges and propellers (for details and references, see Schmidt et al. 2009). An entirely different class of simulations are the semi-analytical streamline models of Hahn (2007, 2008) and Hahn et al. (2009), which use the theoretical framework of Borderies-Rappaport and Longaretti (1994, and references therein) to probe the distribution of surface density, pressure, and viscosity of a ring in the vicinity of a sharp edge. -- 59 -- 4.2. Physical experiments and the coefficient of restitution What happens when two ring particles collide? Gentle collisions, with velocities of order mm s−1, occur constantly (frequencies of order the orbital frequency) within dense rings. A law describing the coefficient of restitution ε (the ratio between outgoing and incoming kinetic energies for two colliding particles in their center-of-mass reference frame) is an essential input for numerical simulations. A number of physical experiments have been conducted to measure the coefficient of restitution directly, but these must begin with assumptions as to the shape, porosity, and surface friction of ring particles. Consequently, the results of these experiments have been inconsistent. More recently, comparisons between simulations and data have attempted to arrive empirically at a favored coefficient of restitution law, from which ring-particle properties can then be inferred. (cid:18) vn (cid:19)b , Bridges et al. (1984) conducted ground-breaking experiments for determining the collisional properties of icy objects, using a double-pendulum apparatus to achieve the exceedingly low collision velocities seen in ring systems. They found relations for εn, as a function of mutual incoming velocity vn (the subscript denotes normal collisions), for frosty ice spheres at temperatures of ∼ 200 K. Their result, vc εn = (15) with exponent b = −0.234 and critical velocity vc = 0.0077 cm s−1, has been widely used in numerical simulations ever since. Further experiments broadened the range of particle properties tested. Hatzes et al. (1988, 1991) found that a frost layer on the surface can greatly increase the lossiness of a collision (i.e., depress ε), even leading to sticking at low collision velocities. Dilley and Crawford (1996) varied the mass of the incoming particles and found that smaller ice balls led to lossier collisions. Supulver et al. (1995), found less lossy collisions overall (Eq. 15 with b = −0.14 and vc = 0.01 cm s−1) for frost-free spheres at ∼ 100 K and also found that glancing collisions are less lossy than normal collisions. Initial results from experiments in a micro-gravity environment were given by Colwell et al. (2008) and Heisselmann et al. (2010), while Durda et al. (2011) measured the coefficient of restitution for two 1-m granite spheres colliding at low speeds. All collision experiments thus far have represented ring particles as hard spheres of H2O ice. The actual shape of ring particles is not known; they are far too small to be sphericalized by hydrostatic equilibrium, and might be expected to be roughly ellipsoidal if they take the shape of their Roche lobes. Their surfaces may be smoothed by the numerous gentle collisions they experience, after the manner of pebbles in a stream bed, but this also is not known. Finally, ring particles are almost certainly not as hard as solid ice, as their outer layers at least are probably quite porous (Section 1.2). Eventually, the role of physical experiments will ideally shift from determining a preferred restitution law based on an assumption of ring-particle properties, to using an empirically indicated restitution law to infer ring-particle properties. A step towards this goal was taken by Porco et al. (2008), who compared numerical simulations with Voyager imaging data of the azimuthal -- 60 -- Fig. 33. -- (left) Coefficients of restitution, as a function of incoming velocity, according to different input laws. The bold portion of each line indicates the velocity range for most simulated small particles. (right) Velocity dispersion profiles for simulated ring-patches using the same input laws. Figures modified from Porco et al. (2008). asymmetry to arrive at a preliminary conclusion favoring a lower coefficient of restitution (i.e., lossier collisions) than given by the prevailing Bridges et al. (1984) law. They favored instead a law with the form (Borderies et al. 1984) − 2 3 (cid:18) v∗ (cid:19)2 v (cid:34) + 10 3 (cid:18) v∗ v (cid:19)2 − 5 (cid:18) v∗ 9 v (cid:19)4(cid:35)1/21/2 ε = (16) This law is based on the Andrews (1930) theory of colliding spheres, but the more flexible for- mulation of Borderies et al. (1984) allows v∗ to be a free parameter adjusting the law's lossiness. Porco et al. (2008) favored a value of v∗ = 0.001 cm s−1; however, a full treatment of the question, incorporating the substantial Cassini data set, has yet to be completed. A plot of coefficient of restitution distributions derived from a variety of laws is shown in Fig. 33. In addition to the azimuthal photometric asymmetry, both the photometry of propellers (Sec- tion 3.1.5) and the sharpness of gap edges (Section 3.1.2) have the potential of being used to constrain the physical properties of ring particles by combining data with simulations, though work on these lines is just beginning. Propellers are large enough to appear in images with a reason- able amount of detail, but small enough that particle-particle interactions play a significant role in determining their structure (e.g. Lewis and Stewart 2009), although their poorly understood photometry (Tiscareno et al. 2010c), possibly due to vertical structure, may make it difficult to use them as a standard. Occultation data infer edges in Saturn's rings as sharp as 10 to 20 m (Col- -- 61 -- well et al. 2010), significantly sharper than the edges obtained in simulations using the standard Bridges et al. (1984) restitution law (Weiss 2005), while preliminary results from simulations with lossier restitution laws yield sharper edges more in line with observations (J. E. Colwell, personal communication, 2010). 4.3. Spectroscopic ground truth The project of characterizing the optical properties of materials occurring in the outer solar system is ongoing. A large amount of spectroscopic data now exists for the rings of Jupiter and Saturn, in spectral ranges from the infrared to the ultraviolet, that contain numerous bands that are potentially diagnostic of ring-particle composition and/or particle size and state. Laboratory measurements of candidate ring-forming materials are compared to these observations in an effort to constrain the chemical and physical composition of the rings. For details and references, see Cuzzi et al. (2009). 5. Age and origin of ring systems The question of the age of a ring system is actually two separate questions, one being the age of the material and the other the age of the structure currently in place. This distinction is particularly important for diffuse dusty rings (Section 3.4). For example, if Amalthea, Thebe, Adrastea, and Metis were to suddenly disappear, or to suddenly cease emitting dust, Jupiter's rings would dissipate on a timescale on the order of 105 years at the most (Burns et al. 2001). Thus, the age of individual particles in a diffuse ring is on the order of the mean residence time, while the age of the overall structure is related to how long the sources have been in place. The distinction between material age and overall structural age also turns out to be useful for Saturn's main rings. Several aspects of the rings are difficult to reconcile with a ring age comparable to that of Saturn (see Charnoz et al. 2009a, and references therein), including 1) the (cid:38) 95% water- ice composition of the A and B rings is difficult to reconcile with the constant pollution of the rings by infall of interplanetary micrometeoroids; 2) the same infall should significantly erode ring particles, especially in regions of lower optical depth; and 3) exchange of angular momentum with moons currently confining ring edges can only be "rewound" for ∼ 107 yr. On the other hand, the disruption of a sun-orbiting object containing sufficient mass in such a way as to form a ring system is an unlikely event given the recent state of the solar system, and Saturn is actually the least likely of the planets to capture an interloper because the balance of mass and solar distance gives it the smallest Hill sphere among the giant planets, leading to some doubts as to the viability of the young-rings scenario (Charnoz et al. 2009b). One potentially viable theory suggests that the B ring core is ancient, dating from the first -- 62 -- Fig. 34. -- Viscous spreading models predict that disk mass after 5 Gyr of evolution is relatively insensitive to the initial disk mass. Figure from Salmon et al. (2010). Gyr of the solar system in which collisions were much more frequent (Charnoz et al. 2009b), but that much of the specific organization of material in Saturn's rings is only ∼ 107 yr old. Recent indications that the B ring's mass has previously been underestimated (see Section 3.1.4 and Robbins et al. 2010) provide an increased buffer against interplanetary pollution (Elliott and Esposito 2011), as well as making the ring precursor body even larger and thus a recent ring origin even more unlikely. Erosion of ring particles can be counteracted by an accretion-driven recycling process (Durisen et al. 1989; Esposito 2006) and is also slowed if ring particles have spent most of their history in a denser structure than those where they are now found. The current ring-moons may also have formed within the rings and emerged only recently (Charnoz et al. 2010). Many questions still remain, however (Charnoz et al. 2009a). The mass of the B ring needs to be better known, as may well happen as a result of the planned close passes during Cassini 's end-of-mission maneuvers (Seal and Buffington 2009) both through direct sensing of the ring's gravitational pull and through its interaction with charged particles and gamma rays (the latter improving on the Pioneer 11 measurement reported by Cooper et al. 1985). Also needed are improvements in knowledge of the interplanetary impactor flux throughout the history of the rings. The proposed recycling mechanism has yet to be carefully described and modeled. And any theory for the formation of Saturn's rings, whether the age be young or old, requires careful separation of water ice from any silicate and/or metal that must have been part of any body accreting directly from the protoplanetary nebula. Canup (2010) recently suggested that Saturn's rings might have been formed by the disrup- -- 63 -- tion of a Titan-sized (∼2500-km radius) proto-moon near the end of the planetary formation period during which the circum-planetary gas disk is present. Due to its size, the proto-moon is differen- tiated, and as it spirals inward due to interaction with the disk, it first passes the Roche radius for ice (see Section 1.2) and its icy mantle is stripped away. Furthermore, because the incompletely consolidated Saturn would have been ∼50% larger than at present, the proto-moon is engulfed by the planet before reaching the Roche radius for its rocky core, thus explaining the rings' icy composition. The initially resulting ring is ∼1000 times more massive than that of today, but Salmon et al. (2010) showed that viscous spreading of a ring over the age of the solar system can lead to a ring like that of today with relatively little sensitivity to the ring's initial mass (Fig. 34). Super-sizing the argument of Charnoz et al. (2010), Canup (2010) suggest that some of Saturn's mid-size moons (with radii of hundreds of km) may have been spawned by the viscous spreading of this massive ring, accounting for their relatively low densities (cf. Section 1.2), and Charnoz et al. (2011) in turn have explored this scenario in more detail. The details of the Canup (2010) simulations are conducted in the context of that author's theory for circum-planetary disks (e.g., Canup and Ward 2002, 2006), which has been criticized by some (e.g., Mosqueira and Estrada 2003). However, the basic outline of the Canup (2010) story appears to have plausible merit even beyond its specific theoretical context. The rings of Uranus and Neptune, with their lower masses and much darker surfaces, have more plausibly remained little changed over the age of the solar system. The uranian rings, as well as the nearby group of moons, may be part of a system that has oscillated between accretion and disruption for many Gyr (Section 2.3). But why are the rings where they are? Is the dynamical environment such that a uniform supply of material at all distances from the planet would result in the rings as we see them today? Or was the supply of source material somehow confined to the locations at which we now see rings? On the planetary level, why does Saturn have its glorious broad dense disk while Uranus and Neptune have much more modest systems and Jupiter has only moon-generated dust? None of these questions has a clear answer as yet. 6. Rings and other disks Planetary rings are just one variety among many disk-shaped systems known to astronomers, but they are the only variety that is not exceedingly far away in time or space or both and thus that is available for close inspection (Burns and Cuzzi 2006). Other astrophysical disks include proto- planetary, proto-lunar, and proto-satellite disks as theorized for the origins of our solar system and as presently observed as gas disks, dust disks, and debris disks at other stars. They also include accretion disks for binary stars, black holes, active galactic nuclei, etc. Some examples of fruitful cross-pollination between planetary rings studies and other astro- physical disks follow. The interpretation of the inner edge of the Fomalhaut disk in terms of a resonant confinement as seen in planetary rings (Kalas et al. 2005) was validated by the discovery -- 64 -- of Fomalhaut b (Kalas et al. 2008). The dynamics of eccentric rings like Uranus'  ring have been extended and applied to astrophysical eccentric disks such as "superhump" binary star systems (Lubow 2010). Both spiral waves (Section 3.1.1) and self-gravity wakes (Section 3.1.4) were first proposed as physical processes likely to occur in galactic disks, with specific applications to plan- etary rings coming a decade or more later. But the shoe is now on the other foot, with direct observations of spiral waves and of SGWs in planetary rings having become so detailed that they can potentially inform understanding of similar processes occurring in galaxies. A similar pro- cess may soon occur with free unstable normal modes, which have long been seen in numerical simulations of proto-planetary disks (Laughlin et al. 1997) and have now likely been observed di- rectly in Saturn's B ring (Spitale and Porco 2010). Finally, the orbital evolution of disk-embedded "propeller" moons and the nature of their interaction with the disk is only just beginning to be directly observed in Saturn's rings (Section 3.1.5), potentially shedding light on the evolution of planetesimals and other disk-embedded masses. I thank Mark Showalter, Joe Burns, Josh Colwell, Jeff Cuzzi, Jonathan Fortney, Doug Hamil- ton, Matt Hedman, Doug Lin, Phil Nicholson, and John Weiss for helpful conversations. I ad- ditionally thank Robin Canup, John Cooper, Estelle Deau, Larry Esposito, and Rob French for valuable comments on the manuscript. I acknowledge funding from NASA Outer Planets Re- search (NNX10AP94G), NASA Cassini Data Analysis (NNX08AQ72G and NNX10AG67G), and the Cassini Project. REFERENCES Acuna, M. H., and N. F. Ness (1976), The main magnetic field of Jupiter, J. Geophys. Res., 81, 2917 -- 2922, doi:10.1029/JA081i016p02917. Agarwal, M., M. S. Tiscareno, M. M. Hedman, and J. A. Burns (2008), Dynamics of faint rings associated with Methone, Anthe and Pallene, AAS Division for Planetary Sciences Meeting Abstracts, 40, 30.02. Agarwal, M., M. S. Tiscareno, M. M. Hedman, and J. A. Burns (2009), Dynamics of rings and arcs associated with three small moons of Saturn: Methone, Anthe and Pallene, AAS Division on Dynamical Astronomy Meeting Abstracts, 40, 3.05. Albers, N., M. Sremcevi´c, J. E. Colwell, and L. W. Esposito (2012), Saturn's F ring as seen by Cassini UVIS: Kinematics and statistics, Icarus, 217, 367 -- 388, doi:10.1016/j.icarus.2011.11. 016. Alexander, A. F. O. (1962), The Planet Saturn: A History of Observation, Theory and Discovery, Faber and Faber, London. -- 65 -- Andrews, J. P. (1930), Theory of collision of spheres of soft metals, Phil. Mag., Series 7, 9, 593 -- 610. Barnes, J. W., and J. J. Fortney (2004), Transit detectability of ring systems around extrasolar giant planets, Astrophys. J., 616, 1193 -- 1203, doi:10.1086/425067. Beurle, K., C. D. Murray, G. A. Williams, M. W. Evans, N. J. Cooper, and C. B. Agnor (2010), Direct evidence for gravitational instability and moonlet formation in Saturn's rings, Astro- phys. J. Lett., 718, L176 -- L180, doi:10.1088/2041-8205/718/2/L176. Borderies, N., and P. Y. Longaretti (1987), Description and behavior of streamlines in planetary rings, Icarus, 72, 593 -- 603, doi:10.1016/0019-1035(87)90055-8. Borderies, N., P. Goldreich, and S. Tremaine (1983), The dynamics of elliptical rings, Astron. J., 88, 1560 -- 1568, doi:10.1086/113446. Borderies, N., P. Goldreich, and S. Tremaine (1984), Unsolved problems in planetary ring dynamics, in Planetary Rings, edited by R. Greenberg and A. Brahic, pp. 713 -- 734, Univ. Arizona Press, Tucson. Borderies-Rappaport, N., and P.-Y. Longaretti (1994), Test particle motion around an oblate planet, Icarus, 107, 129 -- 141, doi:10.1006/icar.1994.1011. Bosh, A. S., C. B. Olkin, R. G. French, and P. D. Nicholson (2002), Saturn's F ring: Kinematics and particle sizes from stellar occultation studies, Icarus, 157, 57 -- 75, doi:10.1006/icar.2002.6791. Bridges, F. G., A. Hatzes, and D. N. C. Lin (1984), Structure, stability and evolution of Saturn's rings, Nature, 309, 333 -- 335, doi:10.1038/309333a0. Brown, T. M., D. Charbonneau, R. L. Gilliland, R. W. Noyes, and A. Burrows (2001), Hubble Space Telescope time-series photometry of the transiting planet of HD 209458, Astrophys. J., 552, 699 -- 709, doi:10.1086/320580. Burns, J. A., and J. N. Cuzzi (2006), Our local astrophysical laboratory, Science, 312, 1753 -- 1755. Burns, J. A., P. L. Lamy, and S. Soter (1979), Radiation forces on small particles in the solar system, Icarus, 40, 1 -- 48, doi:10.1016/0019-1035(79)90050-2. Burns, J. A., M. R. Showalter, and G. E. Morfill (1984), The ethereal rings of Jupiter and Saturn, in Planetary Rings, edited by R. Greenberg and A. Brahic, pp. 200 -- 272, Univ. Arizona Press, Tucson. Burns, J. A., L. E. Schaffer, R. J. Greenberg, and M. R. Showalter (1985), Lorentz resonances and the structure of the Jovian ring, Nature, 316, 115 -- 119. Burns, J. A., M. R. Showalter, D. P. Hamilton, P. D. Nicholson, I. de Pater, M. E. Ockert-Bell, and P. C. Thomas (1999), The formation of Jupiter's faint rings, Science, 284, 1146 -- 1150, doi:10.1126/science.284.5417.1146. -- 66 -- Burns, J. A., D. P. Hamilton, and M. R. Showalter (2001), Dusty rings and circumplanetary dust: Observations and simple physics, in Interplanetary Dust, edited by E. Grun, B. A. S. Gustafson, S. Dermott, and H. Fechtig, pp. 641 -- 725, Springer, Berlin. Burns, J. A., D. P. Simonelli, M. R. Showalter, D. P. Hamilton, C. D. Porco, H. Throop, and L. W. Esposito (2004), Jupiter's ring-moon system, in Jupiter: The Planet, Satellites and Magnetosphere, edited by Bagenal, F., Dowling, T. E., & McKinnon, W. B., pp. 241 -- 262, Cambridge Univ. Press, Cambridge. Canup, R. M. (2010), Origin of Saturn's rings and inner moons by mass removal from a lost Titan- sized satellite, Nature, 468, 943 -- 926, doi:10.1038/nature09661. Canup, R. M., and L. W. Esposito (1995), Accretion in the Roche zone: Coexistence of rings and ring moons., Icarus, 113, 331 -- 352, doi:10.1006/icar.1995.1026. Canup, R. M., and W. R. Ward (2002), Formation of the Galilean satellites: Conditions of accretion, Astron. J., 124, 3404 -- 3423, doi:10.1086/344684. Canup, R. M., and W. R. Ward (2006), A common mass scaling for satellite systems of gaseous planets, Nature, 441, 834 -- 839, doi:10.1038/nature04860. Chandrasekhar, S. (1969), Ellipsoidal figures of equilibrium, Yale University Press, New Haven. Charnoz, S., C. C. Porco, E. D´eau, A. Brahic, J. N. Spitale, G. Bacques, and K. Baillie (2005), Cassini discovers a kinematic spiral ring a round Saturn, Science, 310, 1300 -- 1304, doi: 10.1126/science.1119387. Charnoz, S., A. Brahic, P. C. Thomas, and C. C. Porco (2007), The equatorial ridges of Pan and Atlas: Terminal accretionary ornaments?, Science, 318, 1622 -- 1624, doi:10.1126/science. 1148631. Charnoz, S., L. Dones, L. W. Esposito, P. R. Estrada, and M. M. Hedman (2009a), Origin and evolution of Saturn's ring system, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 537 -- 575, Springer-Verlag, Dordrecht. Charnoz, S., A. Morbidelli, L. Dones, and J. Salmon (2009b), Did Saturn's rings form during the Late Heavy Bombardment?, Icarus, 199, 413 -- 428, doi:10.1016/j.icarus.2008.10.019. Charnoz, S., J. Salmon, and A. Crida (2010), The recent formation of Saturn's moonlets from viscous spreading of the main rings, Nature, 465, 752 -- 754, doi:10.1038/nature09096. Charnoz, S., et al. (2011), Accretion of Saturn's mid-sized moons during the viscous spreading of young massive rings: Solving the paradox of silicate-poor rings versus silicate-rich moons, Icarus, 216, 535 -- 550, doi:10.1016/j.icarus.2011.09.017. -- 67 -- Chavez, C. E. (2009), Appearance of Saturn's F ring azimuthal channels for the anti-alignment configuration between the ring and Prometheus, Icarus, 203, 233 -- 237, doi:10.1016/j.icarus. 2009.04.033. Chiang, E. I., and P. Goldreich (2000), Apse alignment of narrow eccentric planetary rings, Astro- phys. J., 540, 1084 -- 1090, doi:10.1086/309372. Colombo, G., P. Goldreich, and A. W. Harris (1976), Spiral structure as an explanation for the asymmetric brightness of Saturn's A ring, Nature, 264, 344 -- 345, doi:10.1038/264344a0. Colwell, J. E., L. W. Esposito, and M. Sremcevi´c (2006), Self-gravity wakes in Saturn's A ring measured by stellar occultations from Cassini, Geophys. Res. Lett., 33, L07,201, doi:10. 1029/2005GL025163. Colwell, J. E., L. W. Esposito, M. Sremcevi´c, G. R. Stewart, and W. E. McClintock (2007), Self- gravity wakes and radial structure of Saturn's B ring, Icarus, 190, 127 -- 144. Colwell, J. E., et al. (2008), Ejecta from impacts at 0.2 -- 2.3 m/s in low gravity, Icarus, 195, 908 -- 917, doi:10.1016/j.icarus.2007.12.019. Colwell, J. E., J. H. Cooney, L. W. Esposito, and M. Sremcevi´c (2009a), Density waves in Cassini UVIS stellar occultations. 1. The Cassini Division, Icarus, 200, 574 -- 580, doi:10.1016/j.icarus. 2008.12.031. Colwell, J. E., P. D. Nicholson, M. S. Tiscareno, C. D. Murray, R. G. French, and E. A. Marouf (2009b), The structure of Saturn's rings, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 375 -- 412, Springer-Verlag, Dordrecht. Colwell, J. E., R. G. Jerousek, and L. W. Esposito (2010), Sharp edges in Saturn's rings: Ra- dial structure and longitudinal variability, AAS Division for Planetary Sciences Meeting Abstracts, 42, 50.01. Cooper, J. F., J. H. Eraker, and J. A. Simpson (1985), The secondary radiation under Saturn's A-B-C rings produced by cosmic ray interactions, J. Geophys. Res., 90, 3415 -- 3427, doi: 10.1029/JA090iA04p03415. Crida, A., J. C. B. Papaloizou, H. Rein, S. Charnoz, and J. Salmon (2010), Migration of a moonlet in a ring of solid particles: Theory and application to Saturn's propellers, Astron. J., 140, 944 -- 953, doi:10.1088/0004-6256/140/4/944. Cuzzi, J., R. Clark, G. Filacchione, R. French, R. Johnson, E. Marouf, and L. Spilker (2009), Ring particle composition and size distribution, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 459 -- 509, Springer-Verlag, Dordrecht. Cuzzi, J. N., and J. A. Burns (1988), Charged particle depletion surrounding Saturn's F ring: Evidence for a moonlet belt?, Icarus, 74, 284 -- 324, doi:10.1016/0019-1035(88)90043-7. -- 68 -- Cuzzi, J. N., and J. D. Scargle (1985), Wavy edges suggest moonlet in Encke's gap, Astrophys. J., 292, 276 -- 290, doi:10.1086/163158. Cuzzi, J. N., et al. (2010), An evolving view of Saturn's dynamic rings, Science, 327, 1470 -- 1475, doi:10.1126/science.1179118. Daubechies, I. (1992), Ten Lectures on Wavelets, SIAM, Philadelphia. Dawson, R. I., R. G. French, and M. R. Showalter (2010), Packed perturbers: Short-term in- teractions among Uranus' inner moons, AAS Division on Dynamical Astronomy Meeting Abstracts, 41, 8.07. de Pater, I., S. C. Martin, and M. R. Showalter (2004), Keck near-infrared observations of Saturn's E and G rings during Earth's ring plane crossing in August 1995, Icarus, 172, 446 -- 454, doi:10.1016/j.icarus.2004.07.012. de Pater, I., S. G. Gibbard, E. Chiang, H. B. Hammel, B. Macintosh, F. Marchis, S. C. Martin, H. G. Roe, and M. Showalter (2005), The dynamic neptunian ring arcs: Evidence for a gradual disappearance of Libert´e and resonant jump of Courage, Icarus, 174, 263 -- 272, doi: 10.1016/j.icarus.2004.10.020. de Pater, I., H. B. Hammel, S. G. Gibbard, and M. R. Showalter (2006), New dust belts of Uranus: One ring, two ring, red ring, blue ring, Science, 312, 92 -- 94, doi:10.1126/science.1125110. de Pater, I., H. B. Hammel, M. R. Showalter, and M. A. van Dam (2007), The dark side of the rings of Uranus, Science, 317, 1888 -- 1890, doi:10.1126/science.1148103. Denk, T., et al. (2010), Iapetus: Unique surface properties and a global color dichotomy from Cassini imaging, Science, 327, 435 -- 439, doi:10.1126/science.1177088. Dermott, S. F., and C. D. Murray (1980), Origin of the eccentricity gradient and the apse alignment of the epsilon ring of Uranus, Icarus, 43, 338 -- 349, doi:10.1016/0019-1035(80)90179-7. Dilley, J., and D. Crawford (1996), Mass dependence of energy loss in collisions of icy spheres: An experimental study, J. Geophys. Res., 101, 9267 -- 9270, doi:10.1029/96JE00116. Dones, L., and C. C. Porco (1989), Spiral density wakes in Saturn's A ring?, BAAS, 21, 929. Dumas, C., R. J. Terrile, B. A. Smith, G. Schneider, and E. E. Becklin (1999), Stability of Neptune's ring arcs in question, Nature, 400, 733 -- 735, doi:10.1038/23414. Duncan, M. J., and J. J. Lissauer (1997), Orbital stability of the Uranian satellite system, Icarus, 125, 1 -- 12, doi:10.1006/icar.1996.5568. Durda, D. D., N. Movshovitz, D. C. Richardson, E. Asphaug, A. Morgan, A. R. Rawlings, and C. Vest (2011), Experimental determination of the coefficient of restitution for meter-scale granite spheres, Icarus, 211, 849 -- 855, doi:10.1016/j.icarus.2010.09.003. -- 69 -- Durisen, R. H., N. L. Cramer, B. W. Murphy, J. N. Cuzzi, T. L. Mullikin, and S. E. Cederbloom (1989), Ballistic transport in planetary ring systems due to particle erosion mechanisms I. Theory, numerical methods, and illustrative examples, Icarus, 80, 136 -- 166, doi:10.1016/ 0019-1035(89)90164-4. Durisen, R. H., P. W. Bode, J. N. Cuzzi, S. E. Cederbloom, and B. W. Murphy (1992), Ballistic transport in planetary ring systems due to particle erosion mechanisms II. Theoretical models for Saturn's A- and B-ring inner edges, Icarus, 100, 364 -- 393, doi:10.1016/0019-1035(92) 90106-H. Elliot, J. L., E. Dunham, and D. Mink (1977), The rings of Uranus, Nature, 267, 328 -- 330, doi: 10.1038/267328a0. Elliott, J. P., and L. W. Esposito (2011), Regolith depth growth on an icy body orbiting Saturn and evolution of bidirectional reflectance due to surface composition changes, Icarus, 212, 268 -- 274, doi:10.1016/j.icarus.2010.10.031. Esposito, L. W. (2006), Cassini observations and the history of Saturn's rings, AGU Fall Meeting Abstracts, pp. P23E -- 0110. Esposito, L. W. (2010), Composition, structure, dynamics, and evolution of Saturn's rings, Ann. Rev. Earth Planet. Sci., 38, 383 -- 410, doi:10.1146/annurev-earth-040809-152339. Esposito, L. W., A. Brahic, J. A. Burns, and E. A. Marouf (1991), Particle properties and processes in Uranus' rings, in Uranus, edited by J. T. Bergstralh, E. D. Miner, and M. S. Matthews, pp. 410 -- 465, Univ. Arizona Press, Tucson. Esposito, L. W., B. K. Meinke, J. E. Colwell, P. D. Nicholson, and M. M. Hedman (2008), Moonlets and clumps in Saturn's F ring, Icarus, 194, 278 -- 289, doi:10.1016/j.icarus.2007.10.001. Farmer, A. J., and P. Goldreich (2005), Spoke formation under moving plasma clouds, Icarus, 179, 535 -- 538, doi:10.1016/j.icarus.2005.07.025. Fillius, R. W., C. E. McIlwain, and A. Mogro-Campero (1975), Radiation belts of Jupiter: A second look, Science, 188, 465 -- 467, doi:10.1126/science.188.4187.465. Foryta, D. W., and B. Sicardy (1996), The dynamics of the neptunian Adams ring's arcs, Icarus, 123, 129 -- 167, doi:10.1006/icar.1996.0146. Franklin, F. A., A. F. Cook, R. T. F. Barrey, C. A. Roff, G. E. Hunt, and H. B. de Rueda (1987), Voyager observations of the azimuthal brightness variations in Saturn's rings, Icarus, 69, 280 -- 296, doi:10.1016/0019-1035(87)90106-0. French, R. G., P. D. Nicholson, C. C. Porco, and E. A. Marouf (1991), Dynamics and structure of the Uranian rings, in Uranus, edited by J. T. Bergstralh, E. D. Miner, and M. S. Matthews, pp. 327 -- 409, Univ. Arizona Press, Tucson. -- 70 -- French, R. G., et al. (1993), Geometry of the Saturn system from the 3 July 1989 occultation of 28 SGR and Voyager observations, Icarus, 103, 163 -- 214, doi:10.1006/icar.1993.1066. French, R. S., and M. R. Showalter (2012), Cupid is doomed: An analysis of the stability of the inner Uranian satellites, Icarus, in press, doi:10.1016/j.icarus.2012.06.031. Gaudi, B. S., H. Chang, and C. Han (2003), Probing structures of distant extrasolar planets with microlensing, Astrophys. J., 586, 527 -- 539, doi:10.1086/367539. Giese, B., T. Denk, G. Neukum, T. Roatsch, P. Helfenstein, P. C. Thomas, E. P. Turtle, A. McEwen, and C. C. Porco (2008), The topography of Iapetus' leading side, Icarus, 193, 359 -- 371, doi: 10.1016/j.icarus.2007.06.005. Goertz, C. K., and G. Morfill (1983), A model for the formation of spokes in Saturn's rings, Icarus, 53, 219 -- 229, doi:10.1016/0019-1035(83)90143-4. Goldreich, P., and S. Tremaine (1978a), The velocity dispersion in Saturn's rings, Icarus, 34, 227 -- 239, doi:10.1016/0019-1035(78)90164-1. Goldreich, P., and S. Tremaine (1978b), The formation of the Cassini division in Saturn's rings, Icarus, 34, 240 -- 253, doi:10.1016/0019-1035(78)90165-3. Goldreich, P., and S. Tremaine (1979a), Towards a theory for the Uranian rings, Nature, 277, 97 -- 99, doi:10.1038/277097a0. Goldreich, P., and S. Tremaine (1979b), Precession of the epsilon ring of Uranus, Astron. J., 84, 1638 -- 1641, doi:10.1086/112587. Goldreich, P., and S. Tremaine (1980), Disk-satellite interactions, Astrophys. J., 241, 425 -- 441, doi:10.1086/158356. Goldreich, P., and S. Tremaine (1981), The origin of the eccentricities of the rings of Uranus, Astrophys. J., 243, 1062 -- 1075, doi:10.1086/158671. Goldreich, P., and S. Tremaine (1982), The dynamics of planetary rings, Ann. Rev. Astron. Astro- phys., 20, 249 -- 283, doi:10.1146/annurev.aa.20.090182.001341. Goldreich, P., S. Tremaine, and N. Borderies (1986), Towards a theory for Neptune's arc rings, Astron. J., 92, 490 -- 494, doi:10.1086/114178. Goldreich, P., N. Murray, P. Y. Longaretti, and D. Banfield (1989), Neptune's story, Science, 245, 500 -- 504, doi:10.1126/science.245.4917.500. Hahn, J. M. (2007), The secular evolution of a close ring-satellite system: The excitation of spiral bending waves at a nearby gap edge, Astrophys. J., 665, 856 -- 865, doi:10.1086/519275. -- 71 -- Hahn, J. M. (2008), The secular evolution of a close ring-satellite system: The excitation of spiral density waves at a nearby gap edge, Astrophys. J., 680, 1569 -- 1581, doi:10.1086/588019. Hahn, J. M., J. N. Spitale, and C. C. Porco (2009), Dynamics of the sharp edges of broad planetary rings, Astrophys. J., 699, 686 -- 710, doi:10.1088/0004-637X/699/1/686. Halme, V.-P., H. Salo, M. Sremcevi´c, N. Albers, J. Schmidt, M. Seiss, and F. Spahn (2010), Dy- namical and photometric simulations of propeller features in Saturn's A ring, AAS Division for Planetary Sciences Meeting Abstracts, 42, 50.02. Hamilton, D. P. (1996), The asymmetric time-variable rings of Mars, Icarus, 119, 153 -- 172, doi: 10.1006/icar.1996.0008. Hamilton, D. P. (2006), The collisional cascade model for Saturn's ring spokes, AAS Division for Planetary Sciences Meeting Abstracts, 38, 51.04. Hamilton, D. P., and H. Kruger (2008), The sculpting of Jupiter's gossamer rings by its shadow, Nature, 453, 72 -- 75, doi:10.1038/nature06886. Hanninen, J., and C. Porco (1997), Collisional simulations of Neptune's ring arcs, Icarus, 126, 1 -- 27, doi:10.1006/icar.1996.5613. Hatzes, A. P., F. G. Bridges, and D. N. C. Lin (1988), Collisional properties of ice spheres at low impact velocities, Mon. Not. Roy. Astron. Soc., 231, 1091 -- 1115. Hatzes, A. P., F. Bridges, D. N. C. Lin, and S. Sachtjen (1991), Coagulation of particles in Saturn's rings: Measurements of the cohesive force of water frost, Icarus, 89, 113 -- 121, doi:10.1016/ 0019-1035(91)90091-7. Hedman, M. M., et al. (2005), Morphology, movements and models of ringlets in Saturn's Encke Gap, AAS Division for Planetary Sciences Meeting Abstracts, 37, 64.01. Hedman, M. M., P. D. Nicholson, H. Salo, B. D. Wallis, B. J. Buratti, K. H. Baines, R. H. Brown, and R. N. Clark (2007a), Self-gravity wake structures in Saturn's A ring revealed by Cassini VIMS, Astron. J., 133, 2624 -- 2629, doi:10.1086/516828. Hedman, M. M., J. A. Burns, M. S. Tiscareno, and C. C. Porco (2007b), The heliotropic rings of Saturn, AAS Division for Planetary Sciences Meeting Abstracts, 39, 10.09. Hedman, M. M., et al. (2007c), Saturn's dynamic D ring, Icarus, 188, 89 -- 107, doi:10.1016/j.icarus. 2006.11.017. Hedman, M. M., J. A. Burns, M. S. Tiscareno, C. C. Porco, G. H. Jones, E. Roussos, N. Krupp, C. Paranicas, and S. Kempf (2007d), The source of Saturn's G ring, Science, 317, 653 -- 656, doi:10.1126/science.1143964. -- 72 -- Hedman, M. M., P. D. Nicholson, M. R. Showalter, R. H. Brown, B. J. Buratti, and R. N. Clark (2009a), Spectral observations of the Enceladus plume with Cassini-VIMS, Astrophys. J., 693, 1749 -- 1762, doi:10.1088/0004-637X/693/2/1749. Hedman, M. M., C. D. Murray, N. J. Cooper, M. S. Tiscareno, K. Beurle, M. W. Evans, and J. A. Burns (2009b), Three tenuous rings/arcs for three tiny moons, Icarus, 199, 378 -- 386, doi:10.1016/j.icarus.2008.11.001. Hedman, M. M., J. A. Burns, M. S. Tiscareno, and C. C. Porco (2009c), Organizing some very tenuous things: Resonant structures in Saturn's faint rings, Icarus, 202, 260 -- 279, doi:10. 1016/j.icarus.2009.02.016. Hedman, M. M., P. D. Nicholson, K. H. Baines, B. J. Buratti, C. Sotin, R. N. Clark, R. H. Brown, R. G. French, and E. A. Marouf (2010a), The architecture of the Cassini Division, Astron. J., 139, 228 -- 251, doi:10.1088/0004-6256/139/1/228. Hedman, M. M., J. A. Burt, J. A. Burns, and M. S. Tiscareno (2010b), The shape and dynamics of a heliotropic dusty ringlet in the Cassini Division, Icarus, 210, 284 -- 297, doi:10.1016/j. icarus.2010.06.017. Hedman, M. M., N. J. Cooper, C. D. Murray, K. Beurle, M. W. Evans, M. S. Tiscareno, and J. A. Burns (2010c), Aegaeon (Saturn LIII), a G-ring object, Icarus, 207, 433 -- 447, doi: 10.1016/j.icarus.2009.10.024. Hedman, M. M., P. D. Nicholson, G. Filacchione, F. Capaccioni, M. Ciarnello, and R. N. Clark (2011a), Correlations between the spectra and structure of Saturn's main rings, AAS Divi- sion for Planetary Sciences Meeting Abstracts, 43, 532. Hedman, M. M., J. A. Burns, and M. S. Tiscareno (2011b), Of horseshoes and heliotropes: The dy- namics of dust in the Encke Gap, AAS Division on Dynamical Astronomy Meeting Abstracts, 42, 8.02. Hedman, M. M., P. D. Nicholson, M. R. Showalter, R. H. Brown, B. J. Buratti, R. N. Clark, K. Baines, and C. Sotin (2011c), The Christiansen Effect in Saturn's narrow dusty rings and the spectral identification of clumps in the F ring, Icarus, 215, 695 -- 711, doi:10.1016/j. icarus.2011.02.025. Hedman, M. M., J. A. Burns, M. W. Evans, M. S. Tiscareno, and C. C. Porco (2011d), Saturn's curiously corrugated C ring, Science, 332, 708 -- 711, doi:10.1126/science.1202238. Heisselmann, D., J. Blum, H. J. Fraser, and K. Wolling (2010), Microgravity experiments on the collisional behavior of saturnian ring particles, Icarus, 206, 424 -- 430, doi:10.1016/j.icarus. 2009.08.009. Hill, J. R., and D. A. Mendis (1981), On the braids and spokes in Saturn's ring system, Moon and Planets, 24, 431 -- 436, doi:10.1007/BF00896908. -- 73 -- Hor´anyi, M., J. A. Burns, M. M. Hedman, G. H. Jones, and S. Kempf (2009), Diffuse Rings, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 511 -- 536, Springer-Verlag, Dordrecht. Ip, W.-H. (2006), On a ring origin of the equatorial ridge of Iapetus, Geophys. Res. Lett., 33, L16,203, doi:10.1029/2005GL025386. Jacobson, R. A., J. K. Campbell, A. H. Taylor, and S. P. Synnott (1992), The masses of Uranus and its major satellites from Voyager tracking data and Earth-based Uranian satellite data, Astron. J., 103, 2068 -- 2078, doi:10.1086/116211. Jacobson, R. A., J. Spitale, C. C. Porco, K. Beurle, N. J. Cooper, M. W. Evans, and C. D. Murray (2008), Revised orbits of Saturn's small inner satellites, Astron. J., 135, 261 -- 263, doi:10.1088/0004-6256/135/1/261. Jones, G. H., et al. (2006), Formation of Saturn's ring spokes by lightning-induced electron beams, Geophys. Res. Lett., 33, L21,202, doi:10.1029/2006GL028146. Jones, G. H., et al. (2008), The dust halo of Saturn's largest icy moon, Rhea, Science, 319, 1380 -- 1384, doi:10.1126/science.1151524. Julian, W. H., and A. Toomre (1966), Non-axisymmetric responses of differentially rotating disks of stars, Astrophys. J., 146, 810 -- 830. Kalas, P., J. R. Graham, and M. Clampin (2005), A planetary system as the origin of structure in Fomalhaut's dust belt, Nature, 435, 1067 -- 1070, doi:10.1038/nature03601. Kalas, P., J. R. Graham, E. Chiang, M. P. Fitzgerald, M. Clampin, E. S. Kite, K. Stapelfeldt, C. Marois, and J. Krist (2008), Optical images of an exosolar planet 25 light-years from Earth, Science, 322, 1345 -- 1348, doi:10.1126/science.1166609. Kempf, S., U. Beckmann, and J. Schmidt (2010), How the Enceladus dust plume feeds Saturn's E ring, Icarus, 206, 446 -- 457, doi:10.1016/j.icarus.2009.09.016. Krivov, A. V., and D. P. Hamilton (1997), Martian dust belts: Waiting for discovery, Icarus, 128, 335 -- 353, doi:10.1006/icar.1997.5753. Laughlin, G., V. Korchagin, and F. C. Adams (1997), Spiral mode saturation in self-gravitating disks, Astrophys. J., 477, 410 -- 423, doi:10.1086/303682. Levison, H. F., K. J. Walsh, A. C. Barr, and L. Dones (2011), Ridge formation and de-spinning of Iapetus via an impact-generated satellite, Icarus, 214, 773 -- 778, doi:10.1016/j.icarus.2011. 05.031. Lewis, M. C., and G. R. Stewart (2009), Features around embedded moonlets in Saturn's rings: The role of self-gravity and particle size distributions, Icarus, 199, 387 -- 412, doi:10.1016/j. icarus.2008.09.009. -- 74 -- Lin, C. C., and F. H. Shu (1964), On the spiral structure of disk galaxies., Astrophys. J., 140, 646 -- 655. Lissauer, J. J. (1985), Shepherding model for Neptune's arc ring, Nature, 318, 544 -- 545, doi:10. 1038/318544a0. Longaretti, P.-Y., and N. Borderies (1991), Streamline formalism and ring orbit determination, Icarus, 94, 165 -- 170, doi:10.1016/0019-1035(91)90147-L. Lubow, S. H. (2010), Eccentricity growth rates of tidally distorted discs, Mon. Not. Roy. Astron. Soc., 406, 2777 -- 2786, doi:10.1111/j.1365-2966.2010.16875.x. Mamajek, E. E., A. C. Quillen, M. J. Pecaut, F. Moolekamp, E. L. Scott, M. A. Kenworthy, A. Collier Cameron, and N. R. Parley (2012), Planetary construction zones in occultation: Discovery of an extrasolar ring system transiting a young Sun-like star and future prospects for detecting eclipses by circumsecondary and circumplanetary disks, Astron. J., 143, 72, doi:10.1088/0004-6256/143/3/72. Matson, D. L., J. C. Castillo-Rogez, G. Schubert, C. Sotin, and W. B. McKinnon (2009), The thermal evolution and internal structure of Saturn's mid-sized icy satellites, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 577 -- 612, Springer-Verlag, Dordrecht. McGhee, C. A., R. G. French, L. Dones, J. N. Cuzzi, H. J. Salo, and R. Danos (2005), HST observations of spokes in Saturn's B ring, Icarus, 173, 508 -- 521, doi:10.1016/j.icarus.2004. 09.001. Meyer, J., and J. Wisdom (2007), Tidal heating in Enceladus, Icarus, 188, 535 -- 539, doi:10.1016/j. icarus.2007.03.001. Meyer, J., and J. Wisdom (2008a), Tidal evolution of Mimas, Enceladus, and Dione, Icarus, 193, 213 -- 223, doi:10.1016/j.icarus.2007.09.008. Meyer, J., and J. Wisdom (2008b), Episodic volcanism on Enceladus: Application of the Ojakangas- Stevenson model, Icarus, 198, 178 -- 180, doi:10.1016/j.icarus.2008.06.012. Michikoshi, S., and E. Kokubo (2011), Formation of a propeller structure by a moonlet in a dense planetary ring, Astrophys. J. Lett., 732, L23, doi:10.1088/2041-8205/732/2/L23. Millis, R. L., L. H. Wasserman, and P. V. Birch (1977), Detection of rings around Uranus, Nature, 267, 330 -- 331, doi:10.1038/267330a0. Miner, E. D., R. R. Wessen, and J. N. Cuzzi (2007), Planetary Ring Systems, Springer Praxis, Chichester. -- 75 -- Mitchell, C., C. Porco, L. Dones, and J. Spitale (2012), The behavior of spokes in Saturn's B ring, Icarus, submitted. Mitchell, C. J., M. Hor´anyi, O. Havnes, and C. C. Porco (2006), Saturn's spokes: Lost and found, Science, 311, 1587 -- 1589, doi:10.1126/science.1123783. Morfill, G. E., and H. M. Thomas (2005), Spoke formation under moving plasma clouds: The Goertz-Morfill model revisited, Icarus, 179, 539 -- 542, doi:10.1016/j.icarus.2005.08.008. Mosqueira, I., and P. R. Estrada (2002), Apse alignment of the Uranian rings, Icarus, 158, 545 -- 556, doi:10.1006/icar.2002.6878. Mosqueira, I., and P. R. Estrada (2003), Formation of the regular satellites of giant planets in an extended gaseous nebula I: subnebula model and accretion of satellites, Icarus, 163, 198 -- 231, doi:10.1016/S0019-1035(03)00076-9. Murray, C. D., and S. F. Dermott (1999), Solar System Dynamics, Cambridge Univ. Press, Cam- bridge. Murray, C. D., and R. P. Thompson (1990), Orbits of shepherd satellites deduced from the structure of the rings of Uranus, Nature, 348, 499 -- 502, doi:10.1038/348499a0. Murray, C. D., and R. P. Thompson (1991), Erratum: Orbits of shepherd satellites deduced from the structure of the rings of Uranus, Nature, 350, 90, doi:10.1038/350090a0. Murray, C. D., C. Chavez, K. Beurle, N. Cooper, M. W. Evans, J. A. Burns, and C. C. Porco (2005), How Prometheus creates structure in Saturn's F ring, Nature, 437, 1326 -- 1329, doi: 10.1038/nature04212. Murray, C. D., K. Beurle, N. J. Cooper, M. W. Evans, G. A. Williams, and S. Charnoz (2008), The determination of the structure of Saturn's F ring by nearby moonlets, Nature, 453, 739 -- 744, doi:10.1038/nature06999. Namouni, F., and C. Porco (2002), The confinement of Neptune's ring arcs by the moon Galatea, Nature, 417, 45 -- 47. Nicholson, P. D., and M. M. Hedman (2010), Self-gravity wake parameters in Saturn's A and B rings, Icarus, 206, 410 -- 423, doi:10.1016/j.icarus.2009.07.028. Nicholson, P. D., I. Mosqueira, and K. Matthews (1995), Stellar occultation observations of Nep- tune's rings: 1984-1988., Icarus, 113, 295 -- 330, doi:10.1006/icar.1995.1025. Nicholson, P. D., et al. (2008), A close look at Saturn's rings with Cassini VIMS, Icarus, 193, 182 -- 212, doi:10.1016/j.icarus.2007.08.036. -- 76 -- Ockert-Bell, M. E., J. A. Burns, I. J. Daubar, P. C. Thomas, J. Veverka, M. J. S. Belton, and K. P. Klaasen (1999), The structure of Jupiter's ring system as revealed by the Galileo imaging experiment, Icarus, 138, 188 -- 213, doi:10.1006/icar.1998.6072. Ohta, Y., A. Taruya, and Y. Suto (2009), Predicting photometric and spectroscopic signatures of rings around transiting extrasolar planets, Astrophys. J., 690, 1 -- 12, doi:10.1088/0004-637X/ 690/1/1. Øieroset, M., D. A. Brain, E. Simpson, D. L. Mitchell, T. D. Phan, J. S. Halekas, R. P. Lin, and M. H. Acuna (2010), Search for Phobos and Deimos gas/dust tori using in situ observations from Mars Global Surveyor MAG/ER, Icarus, 206, 189 -- 198, doi:10.1016/j.icarus.2009.07. 017. Orton, G. S., K. H. Baines, D. Cruikshank, J. N. Cuzzi, S. M. Krimigis, S. Miller, and E. Lel- louch (2009), Review of Knowledge Prior to the Cassini-Huygens Mission and Concurrent Research, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 9 -- 54, Springer-Verlag, Dordrecht. Owen, T., G. E. Danielson, A. F. Cook, C. Hansen, V. L. Hall, and T. C. Duxbury (1979), Jupiter's rings, Nature, 281, 442 -- 446, doi:10.1038/281442a0. Pan, M., and E. Chiang (2010), The propeller and the frog, Astrophys. J. Lett., 722, L178 -- L182, doi:10.1088/2041-8205/722/2/L178. Papaloizou, J. C. B., R. P. Nelson, W. Kley, F. S. Masset, and P. Artymowicz (2007), Disk-planet interactions during planet formation, in Protostars and Planets V, edited by B. Reipurth, D. Jewitt, and K. Keil, pp. 655 -- 668, Univ. Arizona Press, Tucson. Perrine, R. P., D. C. Richardson, and D. J. Scheeres (2011), A numerical model of cohesion in planetary rings, Icarus, 212, 719 -- 735, doi:10.1016/j.icarus.2011.01.029. Porco, C. C. (1991), An explanation for Neptune's ring arcs, Science, 253, 995 -- 1001, doi:10.1126/ science.253.5023.995. Porco, C. C., P. D. Nicholson, J. N. Cuzzi, J. J. Lissauer, and L. W. Esposito (1995), Neptune's ring system, in Neptune and Triton, edited by D. P. Cruikshank, pp. 703 -- 804, Univ. Arizona Press, Tucson. Porco, C. C., et al. (2005a), Cassini Imaging Science: Initial results on Phoebe and Iapetus, Science, 307, 1237 -- 1242, doi:10.1126/science.1107981. Porco, C. C., et al. (2005b), Cassini Imaging Science: Initial results on Saturn's rings and small satellites, Science, 307, 1226 -- 1236, doi:10.1126/science.1108056. Porco, C. C., P. C. Thomas, J. W. Weiss, and D. C. Richardson (2007), Saturn's small satellites: Clues to their origins, Science, 318, 1602 -- 1607. -- 77 -- Porco, C. C., J. W. Weiss, D. C. Richardson, L. Dones, T. Quinn, and H. Throop (2008), Simulations of the dynamical and light-scattering behavior of Saturn's rings and the derivation of ring particle and disk properties, Astron. J., 136, 2172 -- 2200, doi:10.1088/0004-6256/136/5/2172. Rappaport, N. J., P. Longaretti, R. G. French, E. A. Marouf, and C. A. McGhee (2009), A procedure to analyze nonlinear density waves in Saturn's rings using several occultation profiles, Icarus, 199, 154 -- 173, doi:10.1016/j.icarus.2008.08.014. Rein, H., and J. C. B. Papaloizou (2010), Stochastic orbital migration of small bodies in Saturn's rings, Astron. Astrophys., 524, A22, doi:10.1051/0004-6361/201015177. Renner, S., and B. Sicardy (2006), Use of the geometric elements in numerical simulations, Cel. Mech. Dyn. Astron., 94, 237 -- 248, doi:10.1007/s10569-005-5533-3. Richardson, D. C. (1994), Tree code simulations of planetary rings, Mon. Not. Roy. Astron. Soc., 269, 493 -- 511. Richardson, J. E., H. J. Melosh, C. M. Lisse, and B. Carcich (2007), A ballistics analysis of the Deep Impact ejecta plume: Determining Comet Tempel 1's gravity, mass, and density, Icarus, 190, 357 -- 390, doi:10.1016/j.icarus.2007.08.001. Robbins, S. J., G. R. Stewart, M. C. Lewis, J. E. Colwell, and M. Sremcevi´c (2010), Estimating the masses of Saturn's A and B rings from high-optical depth N -body simulations and stellar occultations, Icarus, 206, 431 -- 445, doi:10.1016/j.icarus.2009.09.012. Rosen, P. A., and J. J. Lissauer (1988), The Titan -1:0 nodal bending wave in Saturn's Ring C, Science, 241, 690 -- 694. Roussos, E., G. H. Jones, N. Krupp, C. Paranicas, D. G. Mitchell, S. M. Krimigis, J. Woch, A. Lagg, and K. Khurana (2008), Energetic electron signatures of Saturn's smaller moons: Evidence of an arc of material at Methone, Icarus, 193, 455 -- 464, doi:10.1016/j.icarus.2007.03.034. Salmon, J., S. Charnoz, A. Crida, and A. Brahic (2010), Long-term and large-scale viscous evolution of dense planetary rings, Icarus, 209, 771 -- 785, doi:10.1016/j.icarus.2010.05.030. Salo, H. (1992), Gravitational wakes in Saturn's rings, Nature, 359, 619 -- 621, doi:10.1038/359619a0. Salo, H. (1995), Simulations of dense planetary rings III. Self-gravitating identical particles., Icarus, 117, 287 -- 312, doi:10.1006/icar.1995.1157. Salo, H., and R. Karjalainen (2003), Photometric modeling of Saturn's rings I. Monte Carlo method and the effect of nonzero volume filling factor, Icarus, 164, 428 -- 460, doi:10.1016/ S0019-1035(03)00132-5. Salo, H., R. Karjalainen, and R. G. French (2004), Photometric modeling of Saturn's rings. II. Azimuthal asymmetry in reflected and transmitted light, Icarus, 170, 70 -- 90, doi:10.1016/j. icarus.2004.03.012. -- 78 -- Schenk, P., D. P. Hamilton, R. E. Johnson, W. B. McKinnon, C. Paranicas, J. Schmidt, and M. R. Showalter (2011), Plasma, plumes and rings: Saturn system dynamics as recorded in global color patterns on its midsize icy satellites, Icarus, 211, 740 -- 757, doi:10.1016/j.icarus.2010. 08.016. Schenk, P. M., and W. B. McKinnon (2009), Global color variations on Saturn's icy satellites, and new evidence for Rhea's ring, AAS Division for Planetary Sciences Meeting Abstracts, 41, 3.03. Schlichting, H. E., and P. Chang (2011), Warm Saturns: On the nature of rings around extrasolar planets that reside inside the ice line, Astrophys. J., 734, 117. Schmidt, J., N. Brilliantov, F. Spahn, and S. Kempf (2008), Slow dust in Enceladus' plume from condensation and wall collisions in tiger stripe fractures, Nature, 451, 685 -- 688, doi:10.1038/ nature06491. Schmidt, J., K. Ohtsuki, N. Rappaport, H. Salo, and F. Spahn (2009), Dynamics of Saturn's Dense Rings, in Saturn from Cassini-Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 413 -- 458, Springer-Verlag, Dordrecht. Seager, S. (Ed.) (2010), Exoplanets, Univ. Arizona Press, Tucson. Seal, D. A., and B. B. Buffington (2009), The Cassini extended mission, in Saturn from Cassini- Huygens, edited by M. Dougherty, L. Esposito, and S. M. Krimigis, pp. 725 -- 744, Springer- Verlag, Dordrecht. Seiss, M., F. Spahn, M. Sremcevi´c, and H. Salo (2005), Structures induced by small moonlets in Saturn's rings: Implications for the Cassini mission, Geophys. Res. Lett., 32, L11,205, doi:10.1029/2005GL022506. Showalter, M. R. (1991), Visual detection of 1981S13, Saturn's eighteenth satellite, and its role in the Encke gap, Nature, 351, 709 -- 713, doi:10.1038/351709a0. Showalter, M. R. (2011), The rings of Uranus: Shepherded or not?, AAS Division for Planetary Sciences Meeting Abstracts, 43, 1224. Showalter, M. R., and J. N. Cuzzi (1992), Physical properties of Neptune's ring system, BAAS, 24, 1029. Showalter, M. R., and J. J. Lissauer (2006), The second ring-moon system of Uranus: Discovery and dynamics, Science, 311, 973 -- 977, doi:10.1126/science.1122882. Showalter, M. R., J. N. Cuzzi, E. A. Marouf, and L. W. Esposito (1986), Satellite "wakes" and the orbit of the Encke Gap moonlet, Icarus, 66, 297 -- 323, doi:10.1016/0019-1035(86)90160-0. -- 79 -- Showalter, M. R., J. N. Cuzzi, and S. M. Larson (1991), Structure and particle properties of Saturn's E Ring, Icarus, 94, 451 -- 473, doi:10.1016/0019-1035(91)90241-K. Showalter, M. R., D. P. Hamilton, and P. D. Nicholson (2006), A deep search for Martian dust rings and inner moons using the Hubble Space Telescope, Planet. Space Sci., 54, 844 -- 854, doi:10.1016/j.pss.2006.05.009. Showalter, M. R., A. F. Cheng, H. A. Weaver, S. A. Stern, J. R. Spencer, H. B. Throop, E. M. Birath, D. Rose, and J. M. Moore (2007), Clump detections and limits on moons in Jupiter's ring system, Science, 318, 232 -- 234, doi:10.1126/science.1147647. Showalter, M. R., R. French, R. Sfair, C. Arguelles, M. Pajuelo, P. Becerra, M. Hedman, and P. Nicholson (2009), The brightening of Saturn's F ring, AAS Division for Planetary Science Meeting Abstracts, 41, 22.07. Showalter, M. R., D. P. Hamilton, S. A. Stern, H. A. Weaver, A. J. Steffl, and L. A. Young (2011a), New Satellite of (134340) Pluto: S/2011 (134340) 1, Central Bureau Electronic Telegrams, 2769, 1. Showalter, M. R., M. M. Hedman, and J. A. Burns (2011b), The impact of Comet Shoemaker- Levy 9 sends ripples through the rings of Jupiter, Science, 332, 711 -- 713, doi:10.1126/science. 1202241. Shu, F. H. (1984), Waves in planetary rings, in Planetary Rings, edited by R. Greenberg and A. Brahic, pp. 513 -- 561, Univ. Arizona Press, Tucson. Sicardy, B., F. Roddier, C. Roddier, E. Perozzi, J. E. Graves, O. Guyon, and M. J. Northcott (1999), Images of Neptune's ring arcs obtained by a ground-based telescope, Nature, 400, 731 -- 733, doi:10.1038/23410. Smith, B. A., et al. (1982), A new look at the Saturn system - The Voyager 2 images, Science, 215, 504 -- 537, doi:10.1126/science.215.4532.504. Soter, S. (1974), Remarks on the origin of Iapetus' photometric asymmetry, IAU Colloquium, 28. Spahn, F., and M. Sremcevi´c (2000), Density patterns induced by small moonlets in Saturn's rings?, Astron. Astrophys., 358, 368 -- 372. Spencer, J. R., and T. Denk (2010), Formation of Iapetus' extreme albedo dichotomy by exogeni- cally triggered thermal ice migration, Science, 327, 432 -- 435, doi:10.1126/science.1177132. Spitale, J., and C. C. Porco (2006), Shapes and kinematics of eccentric features in Saturn's C ring and Cassini Division, AAS Division on Dynamical Astronomy Meeting Abstracts, 37, 7.02. Spitale, J. N., and C. C. Porco (2009), Time variability in the outer edge of Saturn's A-ring revealed by Cassini imaging, Astron. J., 138, 1520 -- 1528, doi:10.1088/0004-6256/138/5/1520. -- 80 -- Spitale, J. N., and C. C. Porco (2010), Detection of free unstable modes and massive bodies in Saturn's outer B ring, Astron. J., 140, 1747 -- 1757, doi:10.1088/0004-6256/140/6/1747. Spitale, J. N., C. C. Porco, and J. Colwell (2008), An inclined saturnian ringlet at 1.954 Rs, AAS Division for Planetary Sciences Meeting Abstracts, 40, 21.02. Sremcevi´c, M., F. Spahn, and W. J. Duschl (2002), Density structures in perturbed thin cold discs, Mon. Not. Roy. Astron. Soc., 337, 1139 -- 1152, doi:10.1046/j.1365-8711.2002.06011.x. Sremcevi´c, M., J. Schmidt, H. Salo, M. Seiss, F. Spahn, and N. Albers (2007), A belt of moonlets in Saturn's A ring, Nature, 449, 1019 -- 1021, doi:10.1038/nature06224. Sremcevi´c, M., J. E. Colwell, and L. W. Esposito (2009), Small-scale ring structure observed in Cassini UVIS occultations, AGU Fall Meeting Abstracts, pp. P54A -- 05. Steffl, A. J., and S. A. Stern (2007), First constraints on rings in the Pluto system, Astron. J., 133, 1485 -- 1489, doi:10.1086/511770. Stern, S. A., H. A. Weaver, A. J. Steffl, M. J. Mutchler, W. J. Merline, M. W. Buie, E. F. Young, L. A. Young, and J. R. Spencer (2006), A giant impact origin for Pluto's small moons and satellite multiplicity in the Kuiper belt, Nature, 439, 946 -- 948, doi:10.1038/nature04548. Supulver, K. D., F. G. Bridges, and D. N. C. Lin (1995), The coefficient of restitution of ice particles in glancing collisions: Experimental results for unfrosted surfaces, Icarus, 113, 188 -- 199, doi: 10.1006/icar.1995.1015. Tamayo, D., J. A. Burns, D. P. Hamilton, and M. M. Hedman (2011), Finding the trigger to Iapetus' odd global albedo pattern: Dynamics of dust from Saturn's irregular satellites, Icarus, 215, 260 -- 278, doi:10.1016/j.icarus.2011.06.027. Thomson, F. S., E. A. Marouf, G. L. Tyler, R. G. French, and N. J. Rappoport (2007), Periodic microstructure in Saturn's rings A and B, Geophys. Res. Lett., 34, L24,203, doi:10.1029/ 2007GL032526. Tiscareno, M. S. (2012), A modified "Type I migration" model for propeller moons in Saturn's rings, Planet. Space Sci., in press (arXiv:1206.4942). Tiscareno, M. S., J. A. Burns, M. M. Hedman, J. N. Spitale, C. C. Porco, C. D. Murray, and Cassini Imaging Team (2005), Wavy edges and other disturbances in Saturn's Encke and Keeler gaps, AAS Division for Planetary Sciences Meeting Abstracts, 37, 64.02. Tiscareno, M. S., J. A. Burns, M. M. Hedman, C. C. Porco, J. W. Weiss, L. Dones, D. C. Richard- son, and C. D. Murray (2006a), 100-metre-diameter moonlets in Saturn's A Ring from observations of "propeller" structures, Nature, 440, 648 -- 650, doi:10.1038/nature04581. -- 81 -- Tiscareno, M. S., P. D. Nicholson, J. A. Burns, M. M. Hedman, and C. C. Porco (2006b), Unravelling temporal variability in Saturn's spiral density waves: Results and predictions, Astrophys. J. Lett., 651, L65 -- L68, doi:10.1086/509120. Tiscareno, M. S., J. A. Burns, P. D. Nicholson, M. M. Hedman, and C. C. Porco (2007), Cassini imaging of Saturn's rings II. A wavelet technique for analysis of density waves and other radial structure in the rings, Icarus, 189, 14 -- 34. Tiscareno, M. S., J. A. Burns, M. M. Hedman, and C. C. Porco (2008), The population of propellers in Saturn's A ring, Astron. J., 135, 1083 -- 1091, doi:10.1088/0004-6256/135/3/1083. Tiscareno, M. S., M. M. Hedman, J. A. Burns, J. W. Weiss, and C. C. Porco (2009a), Saturn's A Ring has no inner edge, AAS Division for Planetary Sciences Meeting Abstracts, 41, 25.04. Tiscareno, M. S., J. A. Burns, M. M. Hedman, D. DiNino, C. C. Porco, K. Beurle, and M. W. Evans (2009b), Observations of ejecta clouds produced by impacts onto Saturn's rings, AGU Fall Meeting Abstracts, pp. P54A -- 08. Tiscareno, M. S., R. P. Perrine, D. C. Richardson, M. M. Hedman, J. W. Weiss, C. C. Porco, and J. A. Burns (2010a), An analytic parameterization of self-gravity wakes in Saturn's rings, Astron. J., 139, 492 -- 503. Tiscareno, M. S., J. A. Burns, J. N. Cuzzi, and M. M. Hedman (2010b), Cassini imaging search rules out rings around Rhea, Geophys. Res. Lett., 37, L14,205, doi:10.1029/2010GL043663. Tiscareno, M. S., et al. (2010c), Physical characteristics and non-keplerian orbital motion of "propeller" moons embedded in Saturn's rings, Astrophys. J. Lett., 718, L92 -- L96, doi: 10.1088/2041-8205/718/2/L92. Torrence, C., and G. P. Compo (1998), A practical guide to wavelet analysis, Bull. Am. Meteoro- logical Soc., 79, 61 -- 78, (http://atoc.colorado.edu/research/wavelets/). Torrey, P. A., M. S. Tiscareno, J. A. Burns, and C. C. Porco (2008), Mapping Complexity: the wavy edges of the Encke and Keeler Gaps in Saturn's rings, AAS Division on Dynamical Astronomy Meeting Abstracts, 39, 15.19. Van Helden, A. (1984), Saturn through the telescope: A brief historical survey, in Saturn, edited by T. Gehrels and M. S. Matthews, pp. 23 -- 43, Univ. Arizona Press, Tucson. Verbiscer, A. J., M. F. Skrutskie, and D. P. Hamilton (2009), Saturn's largest ring, Nature, 461, 1098 -- 1100, doi:10.1038/nature08515. Ward, W. R. (1986), Density waves in the solar nebula: Differential Lindblad torque, Icarus, 67, 164 -- 180, doi:10.1016/0019-1035(86)90182-X. -- 82 -- Ward, W. R. (1997), Survival of planetary systems, Astrophys. J. Lett., 482, L211 -- L214, doi: 10.1086/310701. Weiss, J. W. (2005), The Physics of Unconstrained Edges in Planetary Rings, Ph.D. thesis, Uni- versity of Colorado. Weiss, J. W., C. C. Porco, and M. S. Tiscareno (2009), Ring edge waves and the masses of nearby satellites, Astron. J., 138, 272 -- 286, doi:10.1088/0004-6256/138/1/272. Winter, O. C., D. C. Mourao, S. M. Giuliatti Winter, F. Spahn, and C. da Cruz (2007), Moonlets wandering on a leash-ring, Mon. Not. Roy. Astron. Soc., 380, L54 -- L57, doi: 10.1111/j.1745-3933.2007.00347.x. Wisdom, J., and M. Holman (1991), Symplectic maps for the n-body problem, Astron. J., 102, 1528 -- 1538. Wisdom, J., and S. Tremaine (1988), Local simulations of planetary rings, Astron. J., 95, 925 -- 940, doi:10.1086/114690. Zebker, H. A., E. A. Marouf, and G. L. Tyler (1985), Saturn's rings: Particle size distributions for thin layer model, Icarus, 64, 531 -- 548, doi:10.1016/0019-1035(85)90074-0. This preprint was prepared with the AAS LATEX macros v5.2.
1211.7027
2
1211
2013-10-18T21:27:35
The influence of orbital dynamics, shape and tides on the obliquity of Mercury
[ "astro-ph.EP" ]
Earth-based radar observations of the rotational dynamics of Mercury (Margot et al. 2012) combined with the determination of its gravity field by MESSENGER (Smith et al. 2012) give clues on the internal structure of Mercury, in particular its polar moment of inertia C, deduced from the obliquity (2.04 +/- 0.08) arcmin. The dynamics of the obliquity of Mercury is a very-long term motion (a few hundreds of kyrs), based on the regressional motion of Mercury's orbital ascending node. This paper, following the study of Noyelles & D'Hoedt (2012), aims at first giving initial conditions at any time and for any values of the internal structure parameters for numerical simulations, and at using them to estimate the influence of usually neglected parameters on the obliquity, like J3, the Love number k2 and the secular variations of the orbital elements. We use for that averaged representations of the orbital and rotational motions of Mercury, suitable for long-term studies. We find that J3 should alter the obliquity by 250 milli-arcsec, the tides by 100 milli-arcsec, and the secular variations of the orbital elements by 10 milli-arcsec over 20 years. The resulting value of C could be at the most changed from 0.346mR^2 to 0.345mR^2.
astro-ph.EP
astro-ph
The influence of orbital dynamics, shape and tides on the obliquity of Mercury Benoıt Noyelles & Christoph Lhotka1 University of Namur, Dpt of Mathematics and Namur Center for Complex Systems (NAXYS), 8 Rempart de la Vierge, B-5000 Namur, Belgium, tel: +3281724940 fax: +3281724914 [email protected],[email protected] ABSTRACT Earth-based radar observations of the rotational dynamics of Mercury (Margot et al. 2012) combined with the determination of its gravity field by MESSENGER (Smith et al. 2012) give clues on the internal structure of Mer- cury, in particular its polar moment of inertia C, deduced from the obliquity (2.04 ± 0.08) arcmin. The dynamics of the obliquity of Mercury is a very-long term motion (a few hundreds of kyrs), based on the regressional motion of Mercury's orbital ascend- ing node. This paper, following the study of Noyelles & D'Hoedt (2012), aims at first giving initial conditions at any time and for any values of the internal structure parameters for numerical simulations, and at using them to estimate the influence of usually neglected parameters on the obliquity, like J3, the Love number k2 and the secular variations of the orbital elements. We use, for that, av- eraged representations of the orbital and rotational motions of Mercury, suitable for long-term studies. We find that J3 should alter the obliquity by 250 milli-arcsec, the tides by 30 milli-arcsec, and the secular variations of the orbital elements by 10 milli- arcsec over 20 years. The resulting value of C could be at the most changed from 0.346mR2 to 0.345mR2. Subject headings: Mercury -- Celestial Mechanics -- Resonances, spin-orbit -- Rotational dynamics 1Now at Dipartimento di Matematica -- Universit`a degli Studi di Roma Tor Vergata -- Via della Ricerca Scientifica, 1, 00133 Roma -- Italy -- 2 -- 1. Introduction The space missions MESSENGER (NASA) and BepiColombo (ESA / JAXA) (Balogh et al. 2007) are opportunities to have a better knowledge of Mercury, in particular its rotational dynamics and its internal structure. The size of an outer molten core can be inverted from the well-known Peale experiment (Peale 1976; Peale et al. 2002) in measuring separately the obliquity of Mercury, yielding the polar momentum of inertia C, and the longitudinal librations, yielding the polar momentum of the mantle Cm. This last assertion is based on the assumption that the longitudinal librations of the mantle are decoupled from the other layers, what can be doubtful if Mercury has a significant inner core (Veasey & Dumberry 2011; Van Hoolst et al. 2012). But this does not affect the determination of C. The determination of C from the obliquity is based on the assumptions that Mercury is at the Cassini State 1, and that it behaves as a rigid body over the timescales relevant for the variations of the obliquity, i.e. a few hundreds of years, due to the regressional motion of Mercury's orbital nodes. The Cassini State 1 is a dynamical equilibrium in which Mercury is in a 3:2 spin-orbit resonance (Pettengill & Dyce 1965; Colombo 1965), and the small free librations around the equilibrium have been damped with enough efficiency so that the obliquity of Mercury can be considered as an equilibrium obliquity. In the following, this obliquity will be denoted as ǫ and is the angle between the normal to the orbit and the angular momentum of Mercury. We will also consider the inertial obliquity K, defined as the angle between the normal to an inertial reference plane and the angular momentum. Peale (2005) estimated the damping timescale to be of the order of 105 years for the free librations and the free precession of the spin, in considering the tidal dissipation and the core-mantle friction. The presence of Mercury near the Cassini State 1 has been confirmed by Earth-based radar observations (Margot et al. 2007, 2012), giving an obliquity of 2.04 ± 0.08 arcminutes. The variations of the obliquity can be considered as small and adiabatic (Bills & Comstock 2005; Peale 2006; Bois & Rambaux 2007; D'Hoedt & Lemaitre 2008). It is commonly accepted that the obliquity ǫ should be inverted using Peale's formula (Peale 1969), that reads as (Yseboodt & Margot 2006): C (cid:19) sin i , (1) ǫ = − C (cid:19) cos i + 2nmR2(cid:0) 7 2e − 123 16 e3(cid:1) C22 − nmR2 (1 − e2)−3/2 C20 where n is the orbital mean motion of Mercury, m its mass, R its mean radius, e its orbital eccentricity, C20 = −J2 and C22 are the most 2 relevant coefficients of the gravity field of Mercury, and i and (cid:19) are the inclination and nodal precession rate of the orbit of Mercury with respect to a Laplace Plane. All these quantities are assumed to be constant. The Laplace Plane is a reference plane, determined so as to minimize the variations of the in- -- 3 -- clination i. There are several ways to define the Laplace Plane (Yseboodt & Margot 2006; Bois & Rambaux 2007; D'Hoedt et al. 2009), and so i and (cid:19) depend on the chosen definition. In order to bypass the uncertainty on the Laplace Plane, Noyelles & D'Hoedt (2012) proposed a Laplace Plane-free study of the obliquity, using averaged equations of the rota- tional motion, averaged variations of the eccentricity, inclination and the associated angles, and a quasi-periodic representation of these quantities. This yields a quasi-periodic repre- sentation of the equilibrium obliquity that is very close to the Cassini State over the domain of validity of the JPL DE 406 ephemerides (Standish 1998), i.e. 6,000 years between JED 0625360.50 (-3000 February 23) and 2816912.50 (+3000 May 06). The inertial reference frame is the ecliptic at J2000. This approach allows to consider the small variations of the orbital elements. The spacecraft MESSENGER currently orbiting around Mercury has recently given us accurate gravity coefficients (Tab.1) (Smith et al. 2012). This is the opportunity to refine our model of the obliquity and to consider usually neglected effects like the tides or higher order gravity coefficients. Table 1: The gravity field coefficients of Mercury derived from MESSENGER data (Smith et al. 2012). These are unnormalized coefficients while Smith et al. give them nor- malized. C20 = −J2 C21 S21 C22 S22 C30 = −J3 C40 = −J4 (−5.031 ± 0.02) × 10−5 (−5.99 ± 6.5) × 10−8 (1.74 ± 6.5) × 10−8 (8.088 ± 0.065) × 10−6 (3.22 ± 6.5) × 10−8 (−1.188 ± 0.08) × 10−5 (−1.95 ± 0.24) × 10−5 The dynamics of the obliquity is a long-term dynamics, since it is ruled by the preces- sional motion of Mercury's orbital node, its period being of the order of 300 kyr. That is the reason why we should average the equations over the short-period perturbations, the period associated being of the order of the year. In previous studies on the subject, only the main perturbative effects were taken into account. In the present work we therefore aim to investigate also the influence of these additional perturbative terms on the obliquity of Mercury: the influence of higher order terms in eccentricity, higher order gravity harmonics, the influence of tides, in both averaged and unaveraged models of the spin-orbit interactions of Mercury. -- 4 -- We first present how we average the equations ruling the rotational dynamics of Mercury (Sect.2). This averaging is more accurate than the one given in (Noyelles & D'Hoedt 2012), i.e. it is done in higher order in eccentricity. Then we explain how we derive 2 new formulae for the obliquity, one analytically (Sect.3), with an approach slightly different than Peale's, and one numerically (Sect.4). Then the reliability of these formulae are tested and additional effects are discussed (Sect.5). The influence on the interpretation of the internal structure of Mercury is finally addressed (Sect.6). 2. The dynamical model This section aims at deriving the equations of the long-term rotation of Mercury. For that we start from the exact equations of the problem, and we expand the potential with respect to the orbital parameters (eccentricity and inclination) and to the spherical harmonics of the gravity field of Mercury (Tab.1). This expansion induces the apparition of fast sinu- soidal perturbations, that disappear after averaging. Such a calculation is already present in (Noyelles & D'Hoedt 2012), but the gravity field of Mercury is there considered only up to the second order, and the expansion in eccentricity / inclination limited to the degree 3. The basic model behind the spin-orbit dynamics of the Sun-Mercury system is given in terms of the Hamiltonian (D'Hoedt & Lemaitre (2008)): H = HK + HR − (GcMm)VG, (2) where Gc is the gravitational constant, M is the mass of the Sun and m ≃ 1.6 · 10−7M is the mass of Mercury. The symbol HK labels the unperturbed Kepler problem: HK = − m3µ2 2L2 o , (3) with the constant µ given by µ = Gc(M + m), and Lo = m√µa is the orbital angular momentum, which depends on the semi-major axis of Mercury a ≃ 0.387AU. HR defines the free rotational motion of Mercury: 1 HR = where A ≤ B < C are the principal moments of inertia, G is the norm of the angular momentum ~G, L = G cos(J) is the projection of ~G onto the polar figure axis of Mercury, 2(cid:0)G2 − L2(cid:1)(cid:18)sin2(l) A + cos2(l) B (cid:19) + L2 2C , (4) -- 5 -- with the angle J usually called the wobble, and l is the conjugated angle to the action L. Physically, l represents the precession of the geometrical polar axis (or figure axis) of Mercury about the angular momentum, and J is the amplitude of this motion. 2.1. General form of the potential The gravitational interaction of the orbital and rotational dynamics is given by the potential VG. It can be expanded into spherical harmonics, and takes the form, after (Cunningham 1970; Bertotti & Farinella 1990): VG = ∞ n Xn=0 Xm=0 Rn rn+1 Pnm(sin ϕ) (Cnm cos(mλ) + Snm sin(mλ)) , (5) where R ≃ 2439.7km (Archinal et al. 2011) is the radius of Mercury, r is the distance of the center of mass of Mercury from the center of mass of the Sun, Pnm = Pnm(u) are the associated Legendre polynomials, which are defined in terms of the standard Legendre polynomials Pn = Pn(u) by: Pnm(u) =(cid:0)1 − u2(cid:1)m/2 d Pn(u) d um . The angles (ϕ, λ) are latitude and longitude with ϕ ∈ (−90◦, 90◦), λ ∈ (0, 360◦), and the Cnm, Snm are the Stokes coefficients, with m ≤ n and n, m ∈ N. The standard convention is to define C00 = 1 and Sn0 = 0. The notation Jn = −Cn0 is also used for the remaining zonal terms, with m = 0. In addition, by the proper choice of the coordinate system through the center of mass of Mercury, the first order coefficients C10, C11 and S11 vanish, the same is true for C21 = S21 = 0 if the axes of figure are aligned with the main axes of inertia. The effect of the perturbation on the rotation is therefore proportional to the size of the remaining zonal (m = 0), tesseral (m < n) and sectorial (m = n) terms. The remaining second order coefficients, C20 = −J2 and C22, can be related to the principal moments of inertia by: J2mR2 = C − (A + B) 2 , C22mR2 = B − A 4 . It is a common practize to introduce the normalized polar moment of inertia c through the additional equation C = cmR2. -- 6 -- The contributions HK,HR and VG are given in different reference frames. Let us denote by e0 the inertial, by e1 the orbital, by e2 the spin and by e3 the figure frame of references, respectively. In the following, the inertial frame will be either a Laplace frame, minimizing the variations of the orbital inclination, or the ecliptic at J2000.0. These choices of course affect the definitions of the inclination i, of the ascending node (cid:19), and of variables of rotation g and h, defined later. To match HK,HR and VG we aim to express them in the inertial frame e0. For this reason we introduce the unit vector (x, y, z), pointing from the center of mass of Mercury to the one of the Sun, which is defined in the body frame e3, in terms of the longitude λ and the latitude ϕ by the relations: x = cos ϕ cos λ, y = cos ϕ sin λ, z = sin ϕ . (6) Together with the definition of Pn in terms of the sum Pn(u) = 1 2n [n/2] Xk=0 (−1)k(2n − 2k)! k!(n − k)!(n − 2k)! un−2k we find for u = sin ϕ the explicit form for Pnm, which are given by: Pnm(sin ϕ) = 1 2n cosm(ϕ) [(n−m)/2] Xk=0 (−1)k(2n − 2k)! k!(n − k)!(n − 2k − m)! sinn−2k−m(ϕ) . (7) Together with the formulae by Vieta (see e.g. Hazewinkel (2001)) sin cos (mλ) = m Xk=0(cid:18) m k (cid:19) cosk λ sinm−k λ sin cos (cid:18) 1 2 (m − k)π(cid:19) we find from Eq.(5), Eq.(6) and Eq.(7): VG = 1 r ∞ Xn=0 n r(cid:19)n Xm=0(cid:18)R (CnmCnm + SnmSnm) , (8) where the Cnm, Snm are now functions of (x, y, z) only. Note, that using the property x2 + y2 + z2 = 1 they can be expressed in different forms. We provide the main terms that we are going to use in the present study in Sect. 2.4. -- 7 -- 2.2. Matching the reference frames To express the unit vector (x, y, z) in e0 we make use of the usual Andoyer angles (l, g, h) together with the angles (J, K); the angle J was already defined above, and the angle K enters the projection of ~G on the inertial z-axis by H = G cos(K). As already stated, we call K the inertial obliquity. The angle h is a node representing the precession of the angular momentum with respect to the orbital plane, and g can be seen as the spin angle. Let us denote by R1, R2, R3 the rotation matrices around the x, y, z - axes, respectively. The unit vector (x, y, z), given in e3, can be expressed in the inertial frame e0 by (x, y, z)e3 = RM · (cos(f ), sin(f ), 0) , where f is the true anomaly, and we introduced the rotation matrix RM being of the form: RM = R3(−l)R1(−J)R3(−g)R1(−K)R3(−h)R3((cid:19))R1(i)R3(ω) . Together with the relations a r = 1 + 2 ∞ Xν=1 Jν(νe) cos(νM) , cos(f ) = 2 1 − e2 e ∞ Xν=1 Jν(νe) cos(νM) − e , sin(f ) = 2√1 − e2 ∞ Xν=1 dJν(νe) de sin(νM) ν , where Jν are the Bessel functions of the first kind and M is the mean anomaly, we are able to express the potential Eq.(8) in terms of the rotational and orbital elements only: VG = VG(l, g, h, J, K, a, e, i, ω,M). (9) It can also be expressed in terms of suitable action angle variables (li, Li), with i = 1, . . . , 6, using the set of modified Andoyer variables -- 8 -- (l1, l2, l3) = (l + g + h, −l, −h) , (L1, L2, L3) = (G, G − L, G − H) = (G, G(1 − cos(J)), G(1 − cos(K))) for the rotational motion, and by making use of the classical Delaunay variables (l4, l5, l6) = (M, ω, (cid:19)) , (L4, L5, L6) =(cid:16)Lo, L4√1 − e2, L5 cos(i)(cid:17) (10) (11) for the orbital dynamics. In this setting the potential can be written in the form V = V (l, L) with l = (l1, . . . , l6) and L = (L1, . . . , L6). 2.3. Simplifications and assumptions The rotation period of Mercury, about Tr = 58.6d, and the orbital period around the Sun, To ≃ 87.9d lie close to the 3 : 2 resonance (2To ≃ 3Tr). It is thus desirable to find a much simpler dynamical model, which reproduces the qualitative dynamics close to the resonance. We first introduce the change of coordinates S3:2 : (l, L)7→(σ, Σ) with σ = (σ1, . . . , σ6), Σ = (Σ1, . . . , Σ6) defined by the generating function S3:2 of the second kind S3:2 = Σ1(cid:18)l1 − 3 2 l4 − l5 − l6(cid:19) + Σ2l2 + Σ3 (l3 + l6) + Σ4l4 + Σ5l5 + Σ6l6 . In this setting the relevant resonant dynamics can be easily described in terms of the variables 3 2 σ1 = l1 − Σ1 = L1 , Σ2 = L2 , Σ3 = L3 , l4 − l5 − l6 , σ2 = l2 , σ3 = l3 + l6 , (12) while the remaining variables become: -- 9 -- 2Σ4 = 2L4 + 3Σ1 , Σ5 = L5 + Σ1 , Σ6 = L6 + Σ1 − Σ3 , and σi = li with i = 4, 5, 6. In our first approach we aim to construct a simple resonant model, valid only close to exact resonance, by making use of the following assumptions, that we also justify briefly: i) we neglect the wobble motion of Mercury, i.e. we assume that the figure polar axis is the rotation axis; we therefore set J = 0, which implies L2 = 0 and reduces Eq.(4) to: HR = 2 Σ1 2C (13) This motion should in fact induce a deviation of about only 1 meter of the spin pole from the geometrical North pole (Noyelles et al. 2010). ii) we neglect the effect of the rotation on the orbital dynamics, i.e. we investigate the dynamics on the reduced phase space dσi dt = ∂H ∂Σi , dΣi dt = − ∂H ∂σi , with i = 1, 3, therefore assume that the orbital parameters a, e, i, ω, (cid:19),M are known quantities, and Σi = Σi(t), σi = σi(t) with i = 4, 5, 6 act as external time-dependent parameters on the dynamics of the reduced phase space. The assumption is valid since the effect of the rotation of Mercury on its orbit is much smaller compared to the perturbations due to the other planets. The reason is that the energy associated with the rotational dynamics is negligible with respect to the orbital energy. iii) we neglect short periodic effects2 and replace the potential VG by its average over the mean anomaly of Mercury M = l4 = σ4, which we denote by hV i, in short: hV i ≡ hVGiσ4 = 1 2π Z 2π 0 VG(σ, Σ)dσ4 . 2 The effects within time scales, which are smaller than the revolution period of Mercury around the Sun. -- 10 -- As a result the averaged potential hV i becomes independent of σ4, or the mean anomaly M, and thus the conjugated action Σ4 becomes a constant of motion, and as a conse- quence the semi-major axis a is assumed to be constant too. The assumption preserves the qualitative aspects of the dynamics since the orbital and rotational periods of Mercury are short compared to the periods of revolution of the remaining nodes. By averaging theory we maintain the qualitative aspects of the dynamics also in the averaged model. Dufey et al. (2009) have estimated the influence of the short-period oscillations to be smaller than 20 milli-arcsecond on the obliquity. iv) we assume the presence of a perturbation leading to an additional precession of the In this setting we find for the nodes with constant precession rates, say ω, (cid:19) 6= 0. remaining phase state variables, connected to the orbital motion, σ5 and σ6: dσ5 dt = ω , dσ6 dt = (cid:19) . These are the only remaining time dependent variables, their frequencies being con- stant. This assumption has to be discussed: the main influence of the perturbations of the other planets that are important for the long-term dynamics of the rotational motion of Mercury are the secular perturbations of the nodes. In a first order approximation we therefore implement this cumulative effect by a linear precession of the nodes. Our first assumption (cid:19) = const is also in agreement with one of the definitions of If we therefore define (cid:19) with respect to the Laplace plane we the Laplace plane. optimize our results. We also consider a constant precession rate ω. This is a pretty good approximation if the reference plane is the ecliptic at J2000, as given by the JPL HORIZONS website over 6,000 years. If we use another reference plane like the Laplace plane, then the argument of the pericenter ωl is defined from a different origin. The two definitions of the pericentre result in very similar precession rates, the difference being due to the differential precession of the orbital plane with respect to these 2 references planes, this is a small effect that we can safely neglect. Note that since σ5 = σ5(t) = ωt + ω0 and σ6 = σ6(t) = (cid:19)t + (cid:19)0 the time dependent generating function S3:2 leads to ∂S3:2 ∂t = −Σ1 ω + (Σ3 − Σ1) (cid:19) which we have to add to our new Hamiltonian. The final resonant model takes the form -- 11 -- HI = h0 − GcMm hV i , with the new fundamental part: h0 = HK + HR − Σ1 ω + (Σ3 − Σ1) (cid:19) (14) (15) and where HK transforms into the new expression: 2m3µ2 HK = − (2Σ4 − 3Σ1)2 . We derive from Eq.(15) the unperturbed angular frequencies: σ1 = σ3 = ∂h0 ∂Σ1 ∂h0 ∂Σ3 = Σ1 C − 3 2 = (cid:19) since m3µ2 (cid:0)Σ4 − 3 2Σ1(cid:1) 3 − ω − (cid:19) = l1 − 3 2 n − ω − (cid:19) , σ4 = ∂h0 ∂Σ4 = m3µ2 (cid:0)Σ4 − 3 2Σ1(cid:1) 3 ≡ n . Here n is the mean motion of Mercury. For ω = (cid:19) = 0 the spin-orbit resonance of Mercury translates into the commensurability: 2 σ1 = 2 l1 − 3 σ4 = 0 , σ3 − σ6 = 0 , while for ω, (cid:19) 6= 0 small frequency corrections due to the potential hV i have to be taken into account. 2.4. Workout of the time dependent resonant model In the following discussion we limit our investigation to the contributions of the averaged potential, in which the Stokes coefficients in Table 1 are bigger than the threshold 10−7, which -- 12 -- turn out to be C20, C22, C30, C40. According to this simplification the averaged potential hV i can be split into the form: hV i = C20 hV20i + C22 hV22i + C30 hV30i + C40 hV40i , where we used the notation hVnmi to indicate the terms proportional to the coefficient Cnm. For the ongoing investigation we also need to separate the individual terms into purely resonant terms, just depending on σ1, σ3, and time dependent resonant terms through the presence of the additional angles σ5 = l5 = ω, σ6 = l6 = (cid:19), with ω = ω(t), (cid:19) = (cid:19)(t). These terms allow to see the variations of the obliquity because of the variations of the precessional motion, while the time independent terms contain only the mean precession rate of the node (cid:19). We use the short-hand notation: hVnmi = hVnmi (σ1, σ3, l5, l6) = fnm(cid:16)hvnmi (σ1, σ3) + hunmi (σ1, σ3, l5, l6)(cid:17) , (16) where fnm denotes a common factor, vnm labels the time-independent resonant and constant terms, while unm labels the time-dependent resonant terms, only. The expression proportional to C20 in the unaveraged potential VG takes the form: The term hV20i: R2 r3 C20 = 1 2 R2 r3 (cid:0)3z2 − 1(cid:1) , which becomes after the average over the mean anomaly, and expanded up to 4th order in the orbital eccentricity e: hV20i = − R2 (8 + 12e2 + 15e4) 256a3 ([1]20 + [2]20 cos (σ3) + [3]20 cos (2σ3)) + O(e5) . The coefficients [j]20, j = 1, 2, 3 depend on Σ1, Σ3 through the norm of the angular mo- mentum G and the inertial obliquity K (see Eq.(10), Eq.(12)), as well as on the orbital inclination i, and can also be found in the Appendix. Note, that hV20i = f20 hv20i (σ3) does not depend on σ1 and is free of the angles l5, l6 up to O(e5) (i.e. hu20i = 0). -- 13 -- The term hV22i: The expression proportional to C22 in VG is of the form: R2 r3 C22 = 3 R2 r3 (cid:0)x2 − y2(cid:1) . The averaged term up to O(e5) becomes: hV22i = − R2 256a3 (hv22i (σ1, σ3) + hu22i (σ1, σ3, l5)) + O(e5) , with hv22i = E1 ([1]22 cos (2σ1) + [2]22 cos (2σ1 + σ3) + [3]22 cos (2σ1 + 2σ3) + [4]22 cos (2σ1 + 3σ3) + [5]22 cos (2σ1 + 4σ3)) , where we collect the terms depending on the orbital eccentricity e by: The time dependent part of the expression is E1 = e(cid:0)−56 + 123e2(cid:1) . hu22i = 318e3 ([6]22 cos (2σ1 + 2l5) + [7]22 cos (2σ1 + σ3 + 2l5) + [8]22 cos (2σ1 + 2σ3 + 2l5) + [9]22 cos (2σ1 + 3σ3 + 2l5) + [10]22 cos (2σ1 + 4σ3 + 2l5)) , where the coefficients [1]22 . . . [10]22, depending on the resonant actions, are given in the Appendix. Note, that hv22i enters a term with the only resonant argument 2σ1, while the other Fourier modes are of the form 2σ1 + kσ3, k = 1, 2, 3, . . . up to order O(e4). The terms hu22i = hu22i (σ1, σ3, l5) do not depend on l6 and contain terms of the form 2σ1 + kσ3 + 2l5, k = 1, 2, 3, . . . with the additional argument 2l5. -- 14 -- The term hV30i: The third order term proportional to C30 turns out to be: R3 r4 C30 = 1 2 R3 r4 z(cid:0)5z2 − 3(cid:1) . After the averaging no pure resonant terms survive (hv30i = 0) and hV30i takes the form hV30i = R3 128a4 hu30i (σ3, l5) with hu30i = E2 ([1]30 sin (l5 − 3σ3) + [2]30 sin (l5 − 2σ3) + [3]30 sin (l5 − σ3) + [4]30 sin (l5) + [5]30 sin (l5 + σ3) + [6]30 sin (l5 + 2σ3) + [7]30 sin (l5 + 3σ3)) . In the above expression the [j]30, with j = 1, . . . 7, label the third order coefficients, depending on the resonant actions, which are collected together in the Appendix. Moreover, we define E2 = e(cid:0)2 + 5e2(cid:1) to collect the contributions, which depend on the eccentricity. Note, that the time dependent terms hu30i are again independent of l6 (they only depend implicitly on it through the definition of σ1, σ3). We also conclude, that C30 does not contribute to the time independent resonant model, at the first order of masses approximation. The expression proportional to C40 of hVGiσ4 originates from the term The term hV40i: R4 r5 C40 = 1 8 R4 r5 (cid:0)3 − 30z2 + 35z4(cid:1) , and can again be split into -- 15 -- hV40i = R4 4096a5 (hv40i (σ3) + hu40i (σ3, l5)) . Here the time-independent part is given by the expression: hv40i = E3 ([1]40 + [2]40 cos (σ3) + [3]40 cos (2σ3) + [4]40 cos (3σ3) + [5]40 cos (4σ3)) , where and the time dependent contributions are: E3 =(cid:0)8 + 40e2 + 105e4(cid:1) , hu40i = E4 ([6]40 cos (2l5) + [7]40 cos (2l5 − 4σ3) + [8]40 cos (2l5 − 3σ3) + [9]40 cos (2l5 − 2σ3) + [10]40 cos (2l5 − σ3) + [11]40 cos (2l5 + σ3) + [12]40 cos (2l5 + 2σ3) + [13]40 cos (2l5 + 3σ3) + [14]40 cos (2l5 + 4σ3)) , where E4 = e2(cid:0)2 + 7e2(cid:1) (the coefficients [. . .]40 can be found again in the Appendix). Note, that hv40i just depends on σ3, while hu40i depends on σ3, 2l5 but not on l6. To summarize, we find: (i) The effect proportional to C20 (or J2) is of order R2/a3, is time independent, and depending on the resonant angle σ3 only. (ii) The effect proportional to C22 can be split into time dependent as well as time inde- pendent terms: the time independent terms are proportional to eR2/a3, while the time dependent contributions are of order e3R2/a3. The former contains a Fourier term depending just on the resonant argument 2σ1, while the latter is depending on integer combinations of σ1, σ3 and l5 only. (iii) There is no time independent effect proportional to C30 (or J3) on the averaged dy- namics. Only time dependent terms, which are of order eR3/a4, and depending on integer combinations of σ3 and l5 contribute to it. -- 16 -- (iv) There is an important, time independent contribution, of order R4/a5, which is pro- portional to C40 and just depending on the resonant angle σ3. The time dependent part of the potential is of order e2R4/a5 only. The preceeding list shows, that the time dependent effects are smaller than the time inde- pendent. For the numerical integration of the averaged system (see Sec.4) we are going to use the full potential of the form Eq.(16), while for the analytical study we are going to use the time independent part of the potential only. 3. Analytical treatment of the Cassini State 1 One of our goals is to find a formula similar to Peale's (Eq.1). For that, we start from our averaged Hamiltonian and make the assumption that the orbital quantities are constant. These quantities are the mean motion n, the eccentricity e, the inclination i, and the re- gression rate of the ascending node (cid:19). Here the quantities i and (cid:19) are defined with respect to a Laplace Plane, that minimizes the variation of the orbital inclination i. This means in particular that i and (cid:19) are different from the ones given by ephemerides, usually using the ecliptic at J2000.0. To obtain a simple analytical formula giving the obliquity, we neglect the resonant terms in hV i, which depend on time through l5 = ω(t), l6 = (cid:19)(t), and investigate the long-term dynamics close to σ1 = σ3 = 0. The potential hV i = hV i (σ1, σ3) reduces to hV i (σ1, σ3) = C20f20 hv20i (σ3) + C22f22 hv22i (σ1, σ3) + C40f40 hv40i (σ3) , the integrable parts reduce to h0 = HK + HR − Σ1 ω + (Σ3 − Σ1) (cid:19) , and the new Hamiltonian model becomes: HI0 = h0 − GcMm(cid:16)C20f20 hv20i + C22f22 hv22i + C40f40 hv40i(cid:17) . (17) The requirement to remain at the equilibrium is that Σ1 = Σ3 = 0 and translates into the set of equations: -- 17 -- f1 (Σ1, Σ3) ≡(cid:18)∂HI0 f2 (Σ1, Σ3) =(cid:18) ∂HI0 ∂Σ1 (cid:19)σ1,σ3=0 ∂Σ3 (cid:19)σ1,σ3=0 = 0 , = 0 . (18) The system can be solved for Σ1 = Σ1∗, and Σ3 = Σ3∗, which implies the 'equilibrium norm' of the angular momentum G∗ and the 'equilibrium obliquity' K∗, which comes from the equations Σ1∗ = G∗, Σ3∗ = Σ1∗ (1 − cos (K∗)). The physical interpretation of the equilibrium solution is the following: σ1 = 0 means that the axis of smallest inertia points (on average) towards the Sun, σ3 = 0 ensures that the node of the equator of Mercury is locked with the node of its orbit. While G∗ defines a small correction of the unperturbed spin frequency, the angle K∗ defines a specific value of the inertial obliquity. It translates to the usual called obliquity ǫ by the relation cos(ǫ) = cos(i) cos(K) + sin(i) sin(K) cos (σ3) , (19) which reduces to ǫ∗ = i − K∗ for σ3 = 0. This corresponds to the third Cassini Law (Cassini 1693; Colombo 1966; Noyelles 2009) stating that the Laplace normal, spin and orbit normal are coplanar. The contributions hvnmi depend on (Σ1, Σ3) through sin K = sK = sK (Σ1, Σ3) , cos K = cK = cK (Σ1, Σ3), since from Σ1 = L1, Σ3 = L3 we find Σ1 = G, Σ3 = Σ1(1 − cos(K)). We aim to express Eq.(18) in terms of (G, K) and thus have also to express the derivatives ∂ /∂Σ1 , ∂ /∂Σ3 in terms of (G, K) too. A simple calculation shows for sK and cK and its derivatives: ∂cK ∂Σ1 = 1 − cK G , ∂cK ∂Σ3 1 G , = − ∂sK ∂Σ1 = cK − 1 GtK , ∂sK ∂Σ3 = 1 GtK , 2 = 1 and 0 ≤ K ≤ where we have introduced tK = tan(K), and used the relation sK π/2 to simplify the expressions. A long but straightforward calculation shows that Eq.(18) can be written as: 3n 2 GCMm 2 + cK R4 + a5 [3]f C40(cid:19) = 0 , f1(G, K) = − f2(G, K) = (cid:19) + C + G − C(cid:0) ω + (cid:19)(cid:1) G (cid:18)R2 G (cid:18) R2 a3 ([4]f C20 + [5]f C22) + GCMm a3 ([1]f C20 + [2]f C22) + a5 [6]f C40(cid:19) = 0 , R4 (20) -- 18 -- 3 [1]f = − 3 [2]f = − 8 15 [3]f = 32(cid:0)8 + 12e2 + 15e4(cid:1) s2(i−K)tK/2 , e(cid:0)−56 + 123e2(cid:1) c3 1024(cid:0)8 + 40e2 + 105e4(cid:1)(cid:0)2s2(i−K) + 7s4(i−K)(cid:1) tK/2 (i−K)/2s(i−K)/2tK/2 , with and [4]f = 3 [5]f = 3 64 [6]f = − 32(cid:0)8 + 12e2 + 15e4(cid:1) s2(i−K)/sK = −[1]f /(2s2 e(cid:0)−56 + 123e2(cid:1) s3 1024(cid:0)8 + 40e2 + 105e4(cid:1)(cid:0)2s2(i−K) + 7s4(i−K)(cid:1) /sK = −[3]f /(2s2 (i−K)/2/sK = −[2]f /(2s2 i−K/s2 K/2) , K/2) , 15 K/2) . Note, that we can use the second part of Eq.(20) to eliminate G from the remaining equation to get: F (K) = −(cid:18)3n 2 + ω + (cid:19)cK(cid:19) + GCMm(cid:18) R2 a3 ([1]F C20 + [2]F C22) + R4 a5 [3]F C40(cid:19) = 0 , (21) with [1]F = − [2]F = − [3]F = − [4]f C Ω [5]f C Ω [6]f C Ω 3 3 = − 32C (cid:19)sK (cid:0)8 + 12e2 + 15e4(cid:1) s2(i−K) , e(cid:0)−56 + 123e2(cid:1) si−K = − 1024C (cid:19)sK (cid:0)8 + 40e2 + 105e4(cid:1)(cid:0)2s2(i−K) + 7s4(i−K)(cid:1) . 64C (cid:19)sKs(i−K)/2 3 , 15 = 2 The relation Eq. 21 can be solved for the angle K in an implicit way. We now substitute Eq.(19) for σ3 = 0 in Eq.(21) and expand around ǫ = 0 up to first order. Moreover we make use of the relations C = cmR2 and n2a3 ≃ GCM to get rid of some constants. Within these approximations we arrive at the simple formula: (cid:19) n 2 3 ǫ =(cid:18)1 + n 2C22(cid:0) 7 cos(i) + 2 ω 3n(cid:19) × c (cid:19) sin(i) 2e − 123 16 e3(cid:1) − C20(cid:0)1 + 3 2e2 + 15 8 e4(cid:1) + C40(cid:0) R a(cid:1)2(cid:0) 5 2 + 25 2 e2 + 525 n (cid:19)2 3(cid:18) (cid:19) 16 e4(cid:1) − 2 (22) . c sin(i)2! -- 19 -- Notice, that here the parameter (cid:19) is assumed to be negative. This yields a negative obliquity, in the following we use a positive value for ǫ, in fact ǫ. The formula coincides for C40 = 0 and ω = 0 with Peale's formula up to O( (cid:19)/n)2 with the exception of the term c (cid:19) cos(i) in the denominator (Yseboodt & Margot 2006) 3. A comparison of Eq.(22) with Eq.(1) and with Eq.(4) of Peale (1981) shows that the difference between those formulas is less than 1 arcsecond within the interval 0.3 ≤ c ≤ 0.4, with an offset of about 10ǫ2 ≃ 600mas in the case of Mercury, compared with the formulae given in (Yseboodt & Margot 2006) and (Peale 1981). Note that the parameter C30 is absent in this simple averaged model (there is however a time dependent effect as we will see below), the influence of C40 on the denominator is of the order of O(R/a)2 ≃ 4. · 10−5, and O( (cid:19)/n)2 ≃ 8. · 10−7 does not modify the results. The quantity 2/3(cid:0) (cid:19) cos(i) + ω(cid:1) /n ≃ 7. · 10−7 has also a negligible influence, this supports the omission of ω in Peale's formula. We conclude the section with a short summary of the assumptions, which were made to obtain Eq.(22). i) zero wobble, ii) only the coupling of the spin on the orbit was taken into account, iii) short periodic effects are neglected (average over mean orbital motion and time), iv) the orbital parameters i, e, n, (cid:19) and ω are kept constant. The resulting accuracy of the formula can be quantified as follows (errors related to angles are given in radians): 1) for the average a 4th order expansion in eccentricity e was taken into account (error O(e5) ≃ 3. · 10−4), 2) time dependent resonant terms are neglected (with this we set dǫ/dt = 0), 3) a first order expansion in obliquity ǫ allowed to make the formula explicit and induced an additional error of O(ǫ2) ≃ 1. · 10−3. 3For small ( (cid:19)/n) it is possible to simplify the calculations by setting G = nsC ≃ 3 2 nC in the second part of Eq.(20), expanding up to 1st order in ǫ close to ǫ = 0. The resulting formula for ǫ coincides with Eq.(22) up to O( (cid:19)/n2). The additional terms are therefore stemming from the small spin frequency correction, not taken into account in (Peale 1981). -- 20 -- This formula and Peale's formula as well are lacking of the fact, that the obliquity changes with time not only because (cid:19) is not a constant but because the 'equilibrium' conditions σ1 = σ3 = 0 cannot be fulfilled for all times. As we have seen, even for constant precession rates, the Hamiltonian is not a conserved quantity, for a Hamiltonian of the form H = H(Σ1, Σ3, σ1, σ3, t) we cannot deduce that σ1 = σ3 = 0 for all times, since i.e. σ1,3 = ∂H(Σ1, Σ3, σ1, σ3, t) ∂Σ1,3 → σ1,3 6= 0 , and therefore σ1,3 are subject to change too. A second (and probably more visible) concern of formulas of the type Eq.(22) is the fact, that the resulting ǫ strongly depends on the choice of the reference plane to which the orbital inclination i and in which the average precession rate (cid:19) are defined. To optimize the result, the reference plane should be the plane to which the variations in i become minimal, which can be seen as a generalization to the so-called Laplace plane (not to be confused with the invariant Laplace plane, the plane normal to the angular momentum of the complete system). In the next Section we present a possible solution to these kinds of problems. 4. Including the secular variations of the orbital elements Up to now we have considered, as Peale did, that the orbital quantities n, e, i, ω and (cid:19) were constant. In (Noyelles & D'Hoedt 2012) a numerical method considering the variations of the orbital elements is proposed. We will here use the same method, with the refinements that more spherical harmonics are considered, the second-order spherical harmonics of Mercury J2 and C22 are known with a much better accuracy, and the averaged equations are expanded up to a higher degree in eccentricity / inclination. 4.1. Influence on one example In a reference frame based on the ecliptic at J2000, we define averaged eccentricities and inclinations (Eq. 6 to 10 of (Noyelles & D'Hoedt 2012)) as: h(t) = −7.76651 × 10−11t2 + 1.43999 × 10−6t + 0.200722, k(t) = −2.31417 × 10−11t2 − 5.52628 × 10−6t + 0.0446629, p(t) = 2.38036 × 10−16t3 − 9.03918 × 10−12t2 − 1.27635 × 10−6t + 0.0456355, (23) (24) (25) -- 21 -- p(t) = −1.04673 × 10−11t2 − 1.27792 × 10−6t + 0.0456362, q(t) = 2.52407 × 10−16t3 − 1.06586 × 10−11t2 + 6.54322 × 10−7t + 0.0406156, q(t) = −1.21729 × 10−11t2 + 6.52656 × 10−7t + 0.0406163, (26) (27) (28) with h(t) = e(t) sin (t), k(t) = e(t) cos (t), p(t) = sin(cid:16) i(t) the time origin begin J2000 and the time unit being the year. The inclination of Mercury i is here defined with respect to the ecliptic at J2000, and = ω + (cid:19) = l5 + l6. The Eq.23 to 28 are fits over the duration of the JPL DE406 ephemerides, i.e. 6,000 years. For the inclination variables, 2 fits are given: the third-order fit (Eq.25 and Eq.27) is more accurate, but the second-order one is easier to extrapolate into a trigonometric decomposition. 2 (cid:17) sin (cid:19)(t) and q(t) = sin(cid:16) i(t) 2 (cid:17) cos (cid:19)(t), From these formulae we extract trigonometric expressions (Eq.24 to 27 in (Noyelles & D'Hoedt 2012)): with h(t) = 0.1990903983 sin 1(t) + 0.01094807206 sin 2(t), k(t) = 0.1990903983 cos 1(t) + 0.01094807206 cos 2(t), p(t) = 0.06094690052 sin Ω1(t) + 0.01442538649 sin Ω2(t), q(t) = 0.06094690052 cos Ω1(t) + 0.01442538649 cos Ω2(t), 1(t) = 2.852011398 × 10−5t + 1.30845314198, 2(t) = 4.767836272 × 10−6t + 2.26085090227, Ω1(t) = −2.298222197 × 10−5t + 0.60658814513, Ω2(t) = 1.340719884 × 10−5t + 2.28580288184. (29) (30) (31) (32) (33) (34) (35) (36) Using trigonometric series make the orbital solutions easy to extrapolate without di- vergence. We extrapolate them over several millions of years to optimize their numerical identification, and we use it to perform numerical integrations of the Hamilton equations derived from the averaged Hamiltonian HI0 (Eq.16). We assume that the resulting obliquity is close to the real one over the validity of the DE406 ephemerides. This assumption will be checked in Sect.5.1. We can now plug these new orbital elements in the equations of the averaged rotational dynamics of Mercury. Using Laskar's frequency analysis (Laskar 1999, 2005), we express -- 22 -- the ecliptic obliquity K (the ecliptic at J2000 being our inertial reference plane) and the resonant argument σ3 with a quasiperiodic decomposition such as K(t) = i(t) +X ai cos (ωit + φi) , σ3(t) = X bj cos (ωjt + φj) , (37) (38) ai and bj being real amplitudes, ωi,j frequencies, and φi,j phases at t = 0. The frequencies can either come from the forced motion, and so are combinations of the ones present in the orbital motion (Eq.33 to Eq.36), or are due to free librations, that are expected to be damped. Their presence in the numerical outputs comes from a non optimal choice of the initial conditions. Our first run with C20 = −5.031 × 10−5, C22 = 8 × 10−6, C30 = −1.188 × 10−5, C40 = −1.95 × 10−5 and C = 0.35mR2 yields the Tab.2. The frequency analysis giving this table has been performed over 7.95 Myr with 4,096 points equally spaced by 1,942.5 years. The identification of the oscillating arguments has been made in comparing the frequencies and initial phases of these arguments with integer combinations of the proper modes (Eq.33 to 36). The phases are useful to discriminate be- tween sines and cosines in the decompositions of K(t) and σ3(t). The free librations should have periods of ≈ 15 years in longitude4 and ≈ 1, 000 years in obliquity (D'Hoedt & Lemaitre 2004; Rambaux & Bois 2004), these periods should appear aliased, while the forced perturba- tions should not since their periods should be bigger than 10,000 years. We then optimize the initial conditions iteratively in removing the free librations, as described in (Noyelles et al. 2013). These free oscillations act as a noise in the determination of the forced ones, that explains for instance small discrepancies in the periods. Refining these initial conditions improves the accuracy of the determination of the forced oscillations. 4.2. Fitting the initial conditions The goal is to find relevant initial conditions for any set of interior parameters (C20, C22, C30, C40, C) consistent with the observations and the theory. For that we considered 56 differents cases, in letting the interior parameters vary one by one between the uncertainties due to MES- SENGER data (Tab.1). The range of variations we considered is: 4The period of the expected longitudinal librations is in fact close to 12 years because of the molten outer core. In this averaged model we consider that Mercury is a rigid, homogeneous body since it is appropriate to estimate the obliquity. -- 23 -- Table 2: An example of frequency analysis of the numerical outputs of our system. N Amplitude (arcmin) Period (y) Identification K − i 1 2 3 4 5 6 7 8 9 σ3 1 2 3 4 5 6 7 8 9 10 1.9755 0.2682 0.0592 0.0254 0.0132 0.0121 0.0026 0.0025 0.0025 6.2509 1.4683 0.3462 0.1112 0.0816 0.0400 0.0396 0.0261 0.0214 0.0209 ∞ 172,665.2 86,332.6 264,530.5 60,999.1 57,555.3 4,167.1 43,166.2 4,302.6 < K − i > Ω2 − Ω1 2Ω2 − 2Ω1 ω1 − ω2 2ω1 − 2Ω1 3Ω2 − 3Ω1 free 4Ω2 − 4Ω1 free Ω2 − Ω1 172,665.2 2Ω2 − 2Ω1 86,332.6 3Ω2 − 3Ω1 57,555.1 2ω1 − 2Ω1 60,999.0 4Ω2 − 4Ω1 43,166.3 497,291.8 2 − 1 + Ω2 − Ω1 104,473.0 1 − 2 + Ω2 − Ω1 45,075.0 21 − 3Ω1 + Ω2 free 4,167.1 free 4,302.6 • C20 ∈ [−5.1 × 10−5;−4.9 × 10−5] (nominal value: −5.031 × 10−5), • C22 ∈ [8 × 10−6; 8.2 × 10−6] (nominal value: 8.088 × 10−6), • C30 ∈ [−1.3 × 10−5,−1.1 × 10−5] (nominal value: −1.188 × 10−5), • C40 ∈ [−2.19 × 10−5;−1.71 × 10−5] (nominal value: −1.95 × 10−5), • C/(mR2) ∈ [0.32; 0.38] (nominal value: 0.35). For each of them, a numerical study has been performed as explained above. After refinement of the initial conditions that gave us an accurate frequency decomposition of the signals, we identified 34 amplitudes, all given by the frequency analysis. We can now write: -- 24 -- K = i + a1 − 2a2 cos (Ω2 − Ω1) + 2a3 cos (2Ω2 − 2Ω1) − 2a4 cos (1 − 2) + 2a5 cos (21 − 2Ω1) − 2a6 cos (3Ω2 − 3Ω1) + 2a7 cos (4Ω2 − 4Ω1) + 2a8 cos (2 − 1 + Ω2 − Ω1) + 2a9 cos (1 − 2 + Ω2 − Ω1) + 2a10 cos (1 + 2 − 2Ω1) − 2a11 cos (21 − 3Ω1 + Ω2) − 2a12 cos (5Ω2 − 5Ω1) + 2a13 cos (21 − 22) − 2a14 cos (1 − 2 − 2Ω1 + 2Ω2) − 2a15 cos (2 − 1 − 2Ω1 + 2Ω2) − 2a16 cos (21 − 2Ω2) , (39) σ3 = 2a17 sin (Ω2 − Ω1) − 2a18 sin (2Ω2 − 2Ω1) + 2a19 sin (3Ω2 − 3Ω1) − 2a20 sin (21 − 2Ω1) − 2a21 sin (4Ω2 − 4Ω1) − 2a22 sin (2 − 1 + Ω2 − Ω1) − 2a23 sin (1 − 2 + Ω2 − Ω1) + 2a24 sin (21 − 3Ω1 + Ω2) + 2a25 sin (5Ω2 − 5Ω1) + 2a26 sin (21 − Ω1 − Ω2) − 2a27 sin (1 + 2 − 2Ω1) + 2a28 sin (−1 + 2 − 2Ω1 + 2Ω2) + 2a29 sin (1 − 2 − 2Ω1 + 2Ω2) − 2a30 sin (21 − 4Ω1 + 2Ω2) − 2a31 sin (6Ω2 − 6Ω1) + 2a32 sin (1 + 2 − 3Ω1 + Ω2) − 2a33 sin (1 − 2 − 3Ω1 + 3Ω2) − 2a34 sin (2 − 1 − 3Ω1 + 3Ω2) , (40) with ai = C/ (mR2) αiC/ (mR2) + βiC20 + γiC22 + δi for i = 1, 2, 5, 6, 17, 18, 19, 20, 21, 22, 23, 24, 26, 27, for i = 3, 4 ai = C/ (mR2) αi + βiC20 + γiC22 ai = C/ (mR2) αi + βiC20 for i = 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 34, and ai = C/ (mR2) αiC/ (mR2) + βiC20 + γi for i = 25, 28, 29, 30, 31, 32, 33. (41) (42) (43) (44) The form of these formulae comes from Peale's formula (Eq.1 & Eq.22). We expected to get amplitudes alike -- 25 -- C/(mR2) αiC/(mR2) + βiC20 + γiC22 + ζiC30 + φiC40 + ηi ai = . (45) For that, we tried to fit amplitudes with respect to one parameter, i.e. when possible and when not, and also f(cid:18) C mR2(cid:19) = aC/ (mR2) 1 + bC/ (mR2) f(cid:18) C mR2(cid:19) = aC/(cid:0)mR2(cid:1) f (C20) = f (C22) = f (C30) = f (C40) = 1 a + bC20 1 a + bC22 1 a + bC30 1 a + bC40 , , , . (46) (47) (48) (49) (50) (51) When 2 numbers are present (a and b), they are fitted simultaneously. It turned out that it was impossible to estimate the influence of C30 and C40 with enough reliability, that is the reason why they do not appear in the final formulae (Eq.41 to 44). The coefficients used are gathered in Tab.3. 5. The influence of the different effects The derivation of these new formulae for the obliquity of Mercury allows us to estimate the influence of usually neglected effects, like the secular variations of the orbital elements, the tides and the higher order harmonics. For that we first need to test the reliability of our initial conditions on a real, non-averaged simulation of the rotation of Mercury. -- 26 -- 5.1. Differences between the averaged and the unaveraged system We proceed as in (Noyelles & D'Hoedt 2012), Sect.4. To simulate the non-averaged rotation of Mercury, we integrate numerically the equations related to the following Hamil- tonian: (P 2 − 3δP ) H = − n 2(1 − δ) 3 GCM 2 (52) (3 − 30z2 + 35z4)! , nd3 ǫ1(cid:0)x2 + y2(cid:1) + ǫ2(cid:0)x2 − y2(cid:1) + ǫ3(cid:18)R d(cid:19)2 d(cid:19) z(5z2 − 3) + ǫ4(cid:18)R with ǫ1 = −C20mR2/C, ǫ2 = 2C22mR2/C, ǫ3 = C30mR2/(2C), ǫ4 = C40mR2/(8C), δ = 1 − Cm/C, Cm being the polar inertial momentum of the mantle of Mercury, P is the norm of the angular momentum normalized by nC, d is the Sun-Mercury distance, and x and y are the first two coordinates of the unit vector pointing to the Sun in a reference frame defined by the principal axes of inertia of Mercury. The canonical variables associated with this Hamiltonian are l1, l3, P = L1 nC , R = L3 nC , the Hamilton equations associated being dl1 dt = ∂H ∂P , dt = ∂H ∂R , dl3 dP dR ∂l1 dt = − ∂H dt = − ∂H ∂l3 , . The position of the Sun with respect to Mercury is computed using JPL DE406 ephemerides, so it contains every perturbation, including the planetary ones. Inappropriate initial condi- tions in the numerical integration of the equations would yield unexpected free oscillations. The free longitudinal oscillations, their period being ≈ 12 years, can be easily damped adi- abatically over 5,000 years, the DE406 ephemerides starting at -3,000. However, the free oscillations in obliquity, whose period is about 1,000 years, cannot be damped over such a timescale without a significant and artificial impact on the equilibrium. So, we add a damp- ing only on the longitudinal motion, and we get free oscillations in obliquity as in Fig.1. This obliquity is the actual obliquity ǫ, it is equal to i− K only if σ3 = 0. We obtain it with cos ǫ = ~G · ~n ~G , (53) -- 27 -- 2.09 2.08 2.07 2.06 2.05 2.04 2.03 i ) n m c r a ( ε y t i u q i l b O 2.02 -5000 -4000 -3000 -2000 -1000 0 1000 Time from J2000 (y) Fig. 1. -- Obliquity ǫ of Mercury given by the unaveraged system, with C20 = −5.031× 10−5, C22 = 8.088 × 10−6, C30 = C40 = 0 and C = 0.35mR2. The ≈ 1,000 years oscillations are free librations that would not be present if the initial conditions were ideal. where ~n is the instantaneous normal to the orbit. The amplitude of the free oscillations can be seen as an estimation of the error due to the initial conditions. This error can come from all the approximations made in the averaging process, in particular the limitation to a first order averaging, the expansions in eccentricity, or the exclusion of the planetary perturbations. The Fig.2 shows the amplitude of these oscillations for 2 orbital theories, JPL DE406 and INPOP10a (Fienga et al. 2011), and different values of the interior parameters C20, C22 and C, the other ones not affecting our initial conditions. These amplitudes have been obtained thanks to a frequency analysis. We can see that the amplitude of the free oscillations is always smaller than 750 milli-arcsec. For INPOP10a, many points are missing. The reason is that this theory gives the orbital motion of Mercury over 2 kyr, while DE406 gives it over 6 kyr. The free librations that the frequency analysis is expected to detect have a period of the order of 1 kyr, so in some of the numerical simulations, 2 kyr are not long enough to represent 2 free periods. For this reason, they are sometimes not detected. Usually ephemerides are designed to be very accurate over a quite limited timespan, so 2 kyr should be long enough. But the specific case of a planetary obliquity is a long-term dynamics, for -- 28 -- that an orbital theory over several thousands of years is required. Anyway, we can see that no free oscillation with an amplitude bigger than 660 milli-arcsec is detected, this is smaller than the worst case with DE406. Margot et al. (2012) derived from observations a moment of inertia C = 0.346mR2, so the amplitude of the free oscillations should be ≈ 650 milli-arcsec with DE406. This is the theoretical error induced by the Eq.39 and Eq.40. ) c e s c r a ( s n o i t a l l i c s o e e r F 0.66 0.64 0.62 0.6 0.58 0.56 0.54 0.52 0.5 0.48 C=0.31mR2 C=0.35mR2 C=0.39mR2 8 8.02 8.04 8.06 8.08 8.1 8.12 8.14 8.16 8.18 8.2 C22 (x1e6) with DE406 with INPOP10a Fig. 2. -- Amplitude of the free oscillations, induced by our initial conditions. Each point is the amplitude of the free oscillations given by a frequency analysis, after a numerical integration of the Hamilton equations derived from Eq.(52) with our initial conditions. We considered differents sets of interior parameters (C20, C22, and C). The orbital motion of Mercury is given by DE406 (left) and INPOP10a (right). 5.2. Influence of J3 We failed to find an influence of C30 = −J3 in our numerical and analytical formulae of the equilibrium obliquity. Anyway, this parameter is present in the full equations of the rotational dynamics of the system, and we checked its influence in plotting the average value of the obliquity ǫ with respect to J3 (Fig.3). In Fig.3 we clearly see the influence of C30 on the obliquity ǫ, which we did not see in the simple analytical nor the single averaged but still time dependent model. As we learned from (iii) of Sect.2.4, there is a time-dependent contribution proportional to C30 of the order of eR3/a4, which is affecting the action Σ3 - thus the inertial obliquity K as well as ǫ through the presence of the resonant argument σ3 and the argument of the perihelion ω. However, the effect is very small due to the presence of the a4 in the denominator and can safely be -- 29 -- 2.0658 2.0657 2.0656 2.0655 2.0654 2.0653 2.0652 2.0651 i ) n m c r a ( ε y t i u q i l b O 2.065 -1.3 -1.28 -1.26 -1.24 -1.22 -1.2 -1.18 -1.16 -1.14 -1.12 -1.1 C30 (x1e5) Fig. 3. -- Influence of C30 on the mean obliquity, obtained after numerical integration of the non-averaged equations. The other interior parameters are taken in Tab.1. neglected for the same reasoning we did to explain why J4 does not influence the results. A linear fit gives (54) so neglecting J3 should induce an error of ≈ 253 mas for J3 = 1.188 × 10−5. We did the same job for J4 without finding any reliable influence. ǫ = (−355.197C30 + 2.06115) arcmin, From a theoretical point of view the difference between the averaged (but still time dependent) and original systems of equations of motion can be seen in the following way: with the average over the mean orbital longitude we neglect not only the fast periodic effects (the relevant timescale being the orbital period of Mercury), but we also set the semi-major axis of Mercury to a constant value. Thus, in the averaged system, we are unable to include all the perturbations, which affect the time evolution of the semi-major axis of the planet. In addition we needed to expand the potential into a truncated powerseries to be able to introduce the resonant argument and perform the averaging over the fast angle. Since we have limited all our expansions to the 4th order in eccentricity, an additional difference of the order O(e5) is expected. Last but not least, we use the simple average rule, which is -- 30 -- equivalent to a first order averaging in the ratio of the masses m/M (the mass of Mercury over the mass of the Sun). 5.3. The tides The shape of Mercury, if hydrostatic, is due to the influence of a centrifugal potential, due to Mercury's spin, and a tidal potential. This tidal potential can be split into a static and an oscillating potential. As a consequence the gravity field parameters C20 and C22 and the polar momentum of inertia C should experience periodic variations alike (Giampieri 2004; Rappaport et al. 2008; Van Hoolst et al. 2008): C20(t) = C static 20 + C22(t) = C static 22 − C(t) = C static − k2 2 k2 24 k2 3 qte cos l4, qt (2 cos l4 − e cos 2l4) , qteMR2 cos l4, (55) (56) (57) where k2 is the classical Love number, e the eccentricity of Mercury, and l4 = M is the mean anomaly of Mercury. These deformations take into account the variations of the distance Sun-Mercury (proportional to the eccentricity e) and the 3:2 spin-orbit resonance, inducing non-synchronous rotation. This is the reason why the variation of C22 is not proportional to the eccentricity (Giampieri 2004). We also have qt = −3 , (58) M a(cid:19)3 m (cid:18)R where M is the mass of the Sun, m the mass of Mercury, R its mean radius and a its semi- major axis. With e = 0.2056, a = 57, 909, 226.5415 km (JPL HORIZONS), and M/m = 6.0239249 × 106, we have qte = −2.778369 × 10−7 and qt = −1.351347 × 10−6. The Love number k2 should be between 0 (fully inelastic rigid Mercury) and 1.5 (fully fluid Mercury). We think, from the detection of the longitudinal librations of Mercury, that it should be closer to a rigid body than to a fluid one, so we consider k2 = 0.5. We have in fact very few data that would help to estimate k2. Spohn et al. (2001) estimate it between 0.3 and 0.45 if Mercury has no rigid inner core, and between 0.1 and 0.4 if it has one. With different assumptions on the composition, Rivoldini et al. (2009) estimate it between 0.2 and 0.8. The results are given in Fig.4. We can see a peak-to-peak variation of ≈ 30 mas. We can also see the secular drift due to the variations of the orbital elements related to eccentricity and inclination, ≈ 10 mas over 20 years. -- 31 -- Fig. 4. -- Variations of the obliquity of Mercury due to tides, with the numerical formula (left), and the analytical one (right). The period of variation is the orbital period, i.e. 88 days. 5.4. Summary The influence of the different effects is gathered in the Tab.4. The free librations that are mentioned are generated by the lack of accuracy of our initial conditions (Eq.39 & 40), they do not have any physical relevance. We have made the assumptions that they have been damped to a negligible amplitude, as suggested by Peale (2005). The other effects are smaller than that, they all have been estimated in this study except the polar motion, coming from (Noyelles et al. 2010), and the amplitude of the short-period librations, from (Dufey et al. 2009). The polar motion is an oscillation of the rotation axis of Mercury about the geometrical figure axis, with a period of 175.9 days. This motion is also plotted in (Rambaux et al. (2007), Fig.7). These numbers are very small compared to the accuracy of the determination of the obliquity, i.e. ≈ 5 arcsec. 6. Inverting the obliquity of Mercury Recently, Margot et al. (2012) measured an obliquity of (2.04 ± 0.08) arcmin, and they used Peale's formula (Eq.1) to get C/(mR2) = 0.346 ± 0.014. In this work, we present alternative formulae that we propose to use to invert the obliquity of Margot et al. (2012). We can consider that we propose 4 new formulae. From an analytical (Eq.22) and a numerical studies (Eq.53, from Eq.39 & 40) we get 2 formulae. The influence of J3, that we detect only numerically with the full equations of the system (Fig.3 & Eq.54) leads us to write down 2 new equations, in which the 2 obliquities given by the Eq.22 & 53 are corrected by 355.197 × J3. -- 32 -- In considering the values of the spherical harmonics given in Tab.1, we get the theoretical obliquity of Mercury 7 years after the date J2000.0 to be close to the mid-date of the radar observations. We considered that C was the only unknown parameter. In the analytical formula, we used the same dynamical parameters as Yseboodt & Margot (2006), i.e. i = 8.6◦ and a regressional period of the ascending node set to 328 kyr. We also consider a precessional period of the pericenter ω of 128 kyr, as suggested by the JPL HORIZONS website. This number is adviced over 6,000 years. And we get • Margot et al. (2012): C/(mR2) = 0.346 ± 0.014, • Analytical formula (Eq.22): C/(mR2) = 0.34712 ± 0.01361, • Numerical formula (Eq.53): C/(mR2) = 0.34576 ± 0.01349, • Analytical formula (Eq.22) + J3: C/(mR2) = 0.34640 ± 0.01361, • Numerical formula (Eq.53) + J3: C/(mR2) = 0.34506 ± 0.01348. All our numbers are consistent with the ones coming from Peale's formula. The polar moment of inertia of Mercury could be 0.345mR2 instead of 0.346mR2, this difference is very small with respect to the accuracy of the observations. Giving so many digits lacks of physical relevance, but is necessary to stress the tiny differences between our 4 formulae. 7. Conclusion This study tackles the influence of usually neglected effects like the secular variations of the orbital elements, the tides and the higher order harmonics on the instantaneous obliquity of Mercury. The main goal is to invert it to get clues on the internal structure, in particular the polar inertial momentum C. Moreover, it gives optimized initial conditions of the orientation of the angular momentum of Mercury, at any time and for any values of the internal structure parameters, that can be directly used in numerical simulations. This is a refinement of (Noyelles & D'Hoedt 2012). These initial conditions have the advantage to be Laplace plane free, they are instead based on the ecliptic, whose definition is robust. They have been obtained thanks to averaged equations of the rotational motion, suitable for long-term studies. We hope that they will help fitting the rotation of Mercury by future experiments, like the radioscience experiment MORE in BepiColombo (Milani et al. 2001). A C-code implementing our formulae can be downloaded with the electronic version of this manuscript. -- 33 -- We have shown that the usually neglected effects have an influence smaller than 1 arcsec, while the observations have an accuracy of ≈ 5 arcsec. The determination of C can be at the most altered from 0.346mR2 to 0.345mR2, what does not fundamentally change our understanding of the internal structure of Mercury. The size of the molten core can be obtained from the longitudinal librations of Mercury, but depends on whether we consider a rigid inner core or not (Van Hoolst et al. 2012). Acknowledgements This research used resources of the Interuniversity Scientific Computing Facility located at the University of Namur, Belgium, which is supported by the F.R.S.-FNRS under con- vention No. 2.4617.07. It also benefited from the financial support of the contract Prodex C90253 "ROMEO" from BELSPO. Benoıt Noyelles is F.R.S.-FNRS post-doctoral research fellow. The authors are indebted to the 2 reviewers, Alain Vienne and Marie Yseboodt, who pointed out some typos in formulae and whose comments significantly improved the manuscript, and to Nicolas Rambaux, who indicated them the reference to Giampieri, to properly include the tidal effects. REFERENCES Archinal B.A., A'Hearn M.F., Bowell E., Conrad A., Consolmagno G.J., Courtin R., Fukushima T., Hestroffer D., Hilton J.L., Krasinsky G.A., Neumann G., Oberst J., Seidelmann P.K., Stooke P., Tholen D.J., Thomas P.C. & Williams I.P., 2011, Report of the IAU Working Group on cartographic coordinates and Rotational Elements: 2009, Celestial Mechanics and Dynamical Astronomy, 109, 101-135 Balogh A., Grard R., Solomon S.C., Schulz R., Langevin Y., Kasaba Y. & Fujimoto M., 2007, Missions to Mercury, Space Sci. Rev., 132, 611-645 Bertotti B. & Farinella P., 1990, Physics of the Earth and the Solar System, Kluwer Bills B.G. & Comstock R.L., 2005, Forced obliquity variations of Mercury, J. Geophys. Res., 110, E04006 Bois E. & Rambaux N., 2007, On the oscillations in Mercury's obliquity, Icarus, 192, 308-317 Cassini G.D., 1693, Trait´e de l'origine et du progr`es de l'astronomie, Paris Colombo G., 1965, Rotational period of the planet Mercury, Nature, 208, 575 -- 34 -- Colombo G., 1966, Cassini's second and third laws, AJ, 71, 891-896 Cunningham L.E., 1970, On the computation of the spherical harmonic terms needed during numerical integration of the orbital motion of an artificial satellite, Celestial Mechan- ics, 2, 207-216 D'Hoedt S. & Lemaitre A., 2004, The spin-orbit resonant rotation of Mercury: a two degree of freedom Hamiltonian model, Cel. Mech. Dyn. Astr., 89, 267-283 D'Hoedt S. & Lemaitre A., 2008, Planetary long periodic terms in Mercury's rotation: a two dimensional adiabatic approach, Cel. Mech. Dyn. Astr., 101, 127-139 D'Hoedt S., Noyelles B., Dufey J. & Lemaitre A., 2009, Determination of an instantaneous Laplace plane for Mercury's rotation, Advances in Space Research, 44, 597-603 Dufey J., Noyelles B., Rambaux N. & Lemaitre A., 2009, Latitudinal librations of Mercury with a fluid core, Icarus, 203, 1-12 Fienga A., Laskar J., Kuchynka P., Manche H., Desvignes G., Gastineau M., Cognard I. & Theureau G., 2011, The INPOP10a planetary ephemeris and its applications in fundamental physics, Celestial Mechanics and Dynamical Astronomy, 111, 363-385 Giampieri G., 2004, A note on the tidally induced potential of a satellite in eccentric orbit, Icarus, 167, 228-230 Hazewinkel M., 2001, Vi`ete theorem, Encyclopedia of Mathematics, Springer Laskar J., 1999, Introduction to frequency analysis, In: Proceedings of 3DHAM95 NATO Advanced Institute, vol.553, S'Agaro, pp. 134-150 Laskar J., 2005, Frequency Map Analysis and quasiperiodic decompositions, In:Benest et al. (Eds.), Hamiltonian Systems and Fourier Analysis: New Prospects for Gravitational Dynamics, Cambridge Sci. Publ., pp. 99-129 Margot J.-L., Peale S.J., Jurgens R.F., Slade M.A. & Holin I.V., 2007, Large longitude libration of Mercury reveals a molten core, Science, 316, 710-714 Margot J.-L., Peale S.J., Solomon S.C., Hauck II S.A., Ghigo F.D., Jurgens R.F., Yseboodt M., Giorgini J.D., Padovan S. & Campbell D.B., 2012, Mercury's moment of inertia from spin and gravity data, J. Geophys. Res., 117, E00L09 Milani A., Rossi A., Vokrouhlick´y D., Villani A. & Bonanno C., 2001, Gravity field and rotation state of Mercury from the BepiColombo Radio Science Experiments, Planet. Space Sci., 29, 1579-1596 -- 35 -- Noyelles B., 2009, Expression of Cassini's third law for Callisto, and theory of its rotation, Icarus, 202, 225-239 Noyelles B., Dufey J. & Lemaitre A., 2010, Core-mantle interactions for Mercury, MNRAS, 407, 479-496 Noyelles B., Delsate N. & Carletti T., 2013, Equilibrium search algorithm of a perturbed quasi-integrable system, submitted, arXiv:1101.2138 Noyelles B. & D'Hoedt S., 2012, Modeling the obliquity of Mercury, Planet. Space Sci., 60, 274-286 Peale S.J., 1969, Generalized Cassini's laws, AJ, 74, 483-489 Peale S.J., 1976, Does Mercury have a molten core?, Nature, 262, 765-766 Peale S.J., 1981, Measurement accuracies required for the determination of a mercurian liquid core, Icarus, 48, 143-145 Peale S.J., Phillips R.J., Solomon S.C., Smith D.E. & Zuber M.T., 2002, A procedure for determining the nature of Mercury's core, Meteoritics and Planetary Science, 37, 1269-1283 Peale S.J., 2005, The free precession and libration of Mercury, Icarus, 178, 4-18 Peale S.J., 2006, The proximity of Mercury's spin to Cassini state 1 from adiabatic invariance, Icarus, 181, 338-347 Pettengill G.H. & Dyce R.B., 1965, A radar determination of the rotation of the planet Mercury, Nature, 206, 1240 Rappaport N.J., Iess L., Wahr J., Lunine J.I., Armstrong J.W., Asmar S.W., Tortora P., Di Benedetto M. & Racioppa P., 2008, Can Cassini detect a subsurface ocean in Titan from gravity measurements?, Icarus, 194, 711-720 Rambaux N. & Bois E., 2004, Theory of the Mercury's spin-orbit motion and analysis of its main librations, Astronomy and Astrophysics, 413, 381-393 Rambaux N., Van Hoolst T., Dehant V. & Bois E., 2007, Inertial core-mantle coupling and libration of Mercury, Astronomy and Astrophysics, 468, 711-719 Rivoldini A., Van Hoolst T. & Verhoeven O., 2009, The interior structure of Mercury and its core sulfur content, Icarus, 201, 12-30 -- 36 -- Smith D.E., Zuber M.T., Phillips R.J, Solomon S.C., Hauck II S.A., Lemoine F.G., Mazarico E., Neumann G.A., Peale S.J., Margot J.-L., Johnson C.L., Torrence M.H., Perry M.E., Rowlands D.D., Goossens S., Head J.W. & Taylor A.H., 2012, Gravity field and internal structure of Mercury from MESSENGER, Science, 336, 214-217 Spohn T., Sohl F., Wieczerkowski K. & Conzelmann V., 2001, The interior structure of Mercury: what we know, what we expect from BepiColombo, Planet. Space Sci., 49, 1561-1570 Standish E.M., 1998, JPL planetary and lunar ephemeris, DE405/LE405. JPL Interoffice Memorandum IOM 312.D-98-048 Van Hoolst T., Rambaux N., Karatekin O., Dehant V. & Rivoldini A., 2008, The librations, shape, and icy shell of Europa, Icarus, 195, 386-399 Van Hoolst T., Rivoldini A., Baland R.-M. & Yseboodt M., 2012, The effect of tides and an inner core on the forced longitudinal libration of Mercury, Earth and Planetary Science Letters, 333-334, 83-90 Veasey M. & Dumberry M., 2011, The influence of Mercury's inner core on its physical libration, Icarus, 214, 265-274 Yseboodt M. & Margot J.-L., 2006, Evolution of Mercury's obliquity, Icarus, 181, 327-337 Appendix A We collect the various coefficients of the form [j]nm, with j, n, m ∈ N, from Section 2.4. If we introduce the notations cx = cos(x), sx = sin(x) they can be written in the form: At second order in hV20i: [1]20 = 3 (1 + 3c2i) c2K , [2]20 = 48cicKsisK , [3]20 = −6c2Ks2 i . At second order in hV22i: [1]22 = 48c4 i/2c4 [4]22 = 48s2 i/2sis2 K/2sK , [5]22 = 48s4 i/2s4 K/2 , K/2 , [2]22 = 12 (1 + ci) (1 + cK) sisK , [3]22 = 18s2 i s2 K , This preprint was prepared with the AAS LATEX macros v5.2. -- 37 -- in hv22i and K/2s2 [6]22 = −4c4 i s4 [9]22 = 4ci (cK − 1) sisK , [10]22 = −4s2 i , [7]22 = 4 (1 + cK) sicisK , [8]22 = − (1 + 3c2i) s2 K , K/2 in hu22i. At order three we have in hV30i: i s3 K , [2]30 = 30 (ci − 1) (1 + 3ci) cKsis2 K , (13 + 20ci + 15c2i) (3 + 5c2K) s2 i/2sK , i/2s2 [1]30 = −30s2 [3]30 = − [4]30 = − [6]30 = 30 (1 + ci) (3ci − 1) cKsis2 3 2 3 4 (3cK + 5c3K) (si + 5s3i) , [5]30 = 3 4 c2 i/2 (13 − 20ci + 15c2i) (sK + 5s3K) , K , [7]30 = −15 (ci − 1) (1 + ci) 2s3 K . The fourth order coefficients, being part of hv40i, turn out to be: 15 8 (9 + 20c2i + 35c4i) (9 + 20c2K + 35c4K) , [2]40 = [1]40 = 3 64 (2s2i + 7s4i) (2s2K + 7s4K) , [3]40 = 15 (5 + 7c2i) (5 + 7c2K) s2 i s2 K , [4]40 = 840cicKs3 i s3 K , [5]40 = 105s4 i s4 K . The fourth order coefficients in hu40i are: [6]40 = 15 4 (5 + 7c2i) (9 + 20c2K + 35c4K) s2 i , [7]40 = 840s4 i s2 i s4 K , 2 K , i/2s3 [9]40 = 120 (9 + 14ci + 7c2i) (5 + 7c2K) s4 [8]40 = 6720(cid:0)2ci/2 + c3i/2(cid:1) cKs5 [10]40 = 30(cid:0)19ci/2 + 7(cid:0)2c3i/2 + c5i/2(cid:1)(cid:1) s3 [11]40 = −30c3 [12]40 = 120c4 [13]40 = 6720c5 i/2(cid:0)19si/2 + 7(cid:0)s5i/2 − 2s3i/2(cid:1)(cid:1) (2s2K + 7s4K) , i/2 (9 − 14ci + 7c2i) (5 + 7c2K) s2 K , i/2cK(cid:0)s3i/2 − 2si/2(cid:1) s3 K , i/2s2 i/2 (2s2K + 7s4K) , K , [14]40 = −210 (ci − 1) (1 + ci) 3s4 K . -- 38 -- Table 3: Coefficients involved in the Eq.41 to 44. γi δi αi βi i 1.224118 × 107 −0.041284174514 −1.022791 × 107 1 −9.6916394157 1.8067315 × 108 −1.495053 × 108 2 −15.290923597 57.319667133 8.648745 × 108 −6.75794 × 108 -- −288.588048 3 109.5839605 −2.3327045 × 109 −2.8315945 × 109 4 -- −5.967955 × 109 −1.4620515 × 1010 5 −4270.5575456 9618.2490875 4.59872 × 109 −3.29021 × 109 5690.3083952 −5791.2841683 6 160882.75 −1.563842 × 1010 7 -- -- −330146.25 −3.3956545 × 1010 -- -- 8 −307545.35 −3.441592 × 1010 9 -- -- −4.59746 × 1010 -- -- −879441.5 10 −5.12722 × 1010 11 −1025209.5 -- -- 760424 −7.318535 × 1010 -- -- 12 −3583265 −1.7097535 × 1011 13 -- -- −1352596 −1.522164 × 1011 -- -- 14 −1475953.5 −1.554497 × 1011 15 -- -- −4788455 −2.4530905 × 1011 -- -- 16 17 0.00068014043987 133840.35 −111936.3 0.0459703076 568708 −0.0063487867442 −474698 0.5097512707 18 19 2431534 −0.32882074509 −2005542 3.6758306831 1.6246447377 −12322835 −31382330 20.470309592 20 21 −4.3502813922 10602865 −8470000 22.351343806 −28807205 −25710160 −14.991580755 1.8968370715 22 23 0.75276141087 −30518425 −25704595 28.007891612 −52410050 −54.254151937 −127924300 151.69560968 24 25 384.87733645 −35729050 -- 11.320467298 275.67522095 −91723450 −262171000 60.949766585 26 27 −265414800 −96422550 257.09346578 93.149257619 -- −1098.8078842 −108893750 256.01109782 28 29 -- −109386550 −1017.8654389 93.896968203 30 −223772150 −4722.8636039 -- 880.39601125 -- 2105.3653480 −145710250 31 −390.02956574 32 -- −8327.8625575 −404092500 1929.3930214 2613.8902647 −456302000 −4855.3715926 -- 33 -- -- −438245500 −2905.1715 34 -- 39 -- Table 4: Influence on the mean obliquity of usually neglected effects. The amplitude of the free librations can be seen as an estimation of the error due to theory. Effect Free librations C30 Polar motion Tides Short-period librations < 20 mas (Dufey et al. 2009) Secular drift C40 Influence on obliquity < 750 mas ≈ 250 mas ≈ 80 mas (Noyelles et al. 2010) ≈ 30 mas ≈ 10 mas over 20 years negligible
1805.09352
1
1805
2018-05-23T18:01:38
HAT-P-11: Discovery of a Second Planet and a Clue to Understanding Exoplanet Obliquities
[ "astro-ph.EP" ]
HAT-P-11 is a mid-K dwarf that hosts one of the first Neptune-sized planets found outside the solar system. The orbit of HAT-P-11b is misaligned with the star's spin --- one of the few known cases of a misaligned planet orbiting a star less massive than the Sun. We find an additional planet in the system based on a decade of precision radial velocity (RV) measurements from Keck/HIRES. HAT-P-11c is similar to Jupiter in its mass ($M_P \sin{i} = 1.6\pm0.1$ $M_J$) and orbital period ($P = 9.3^{+1.0}_{-0.5}$ year), but has a much more eccentric orbit ($e=0.60\pm0.03$). In our joint modeling of RV and stellar activity, we found an activity-induced RV signal of $\sim$7 m s$^{-1}$, consistent with other active K dwarfs, but significantly smaller than the 31 m s$^{-1}$ reflex motion due to HAT-P-11c. We investigated the dynamical coupling between HAT-P-11b and c as a possible explanation for HAT-P-11b's misaligned orbit, finding that planet-planet Kozai interactions cannot tilt planet b's orbit due to general relativistic precession; however, nodal precession operating on million year timescales is a viable mechanism to explain HAT-P-11b's high obliquity. This leaves open the question of why HAT-P-11c may have such a tilted orbit. At a distance of 38 pc, the HAT-P-11 system offers rich opportunities for further exoplanet characterization through astrometry and direct imaging.
astro-ph.EP
astro-ph
Draft version May 25, 2018 Preprint typeset using LATEX style AASTeX6 v. 1.0 HAT-P-11: DISCOVERY OF A SECOND PLANET AND A CLUE TO UNDERSTANDING EXOPLANET OBLIQUITIES Samuel W. Yee1,12, Erik A. Petigura1,7, Benjamin J. Fulton1,8, Heather A. Knutson1, Konstantin Batygin1, Gáspár Á. Bakos2,9, Joel D. Hartman2, Lea A. Hirsch3, Andrew W. Howard1, Howard Isaacson3, Molly R. Kosiarek4,10, Evan Sinukoff5,1, and Lauren M. Weiss6,11 8 1 0 2 y a M 3 2 . ] P E h p - o r t s a [ 1 v 2 5 3 9 0 . 5 0 8 1 : v i X r a 1California Institute of Technology, Pasadena, CA, 91125, USA 2Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA 3University of California, Berkeley, Berkeley, CA 94720, USA 4University of California, Santa Cruz, Santa Cruz, CA, 95064, USA 5Institute for Astronomy, University of Hawai'i, Honolulu, HI 96822, USA 6University of Montréal, Montréal, QC H3T 1J4, Canada 7Hubble Fellow 8Texaco Postdoctoral Fellow 9Packard Fellow 10NSF Graduate Research Fellow 11Trottier Fellow [email protected] ABSTRACT HAT-P-11 is a mid-K dwarf that hosts one of the first Neptune-sized planets found outside the solar system. The orbit of HAT-P-11b is misaligned with the star's spin - one of the few known cases of a misaligned planet orbiting a star less massive than the Sun. We find an additional planet in the system based on a decade of precision radial velocity (RV) measurements from Keck/HIRES. HAT-P-11c is similar to Jupiter in its mass (MP sin i = 1.6±0.1 MJ) and orbital period (P = 9.3+1.0−0.5 year), but has a much more eccentric orbit (e = 0.60± 0.03). In our joint modeling of RV and stellar activity, we found an activity-induced RV signal of ∼7 m s−1, consistent with other active K dwarfs, but significantly smaller than the 31 m s−1 reflex motion due to HAT-P-11c. We investigated the dynamical coupling between HAT-P-11b and c as a possible explanation for HAT-P-11b's misaligned orbit, finding that planet-planet Kozai interactions cannot tilt planet b's orbit due to general relativistic precession; however, nodal precession operating on million year timescales is a viable mechanism to explain HAT- P-11b's high obliquity. This leaves open the question of why HAT-P-11c may have such a tilted orbit. At a distance of 38 pc, the HAT-P-11 system offers rich opportunities for further exoplanet characterization through astrometry and direct imaging. Keywords: planetary systems – planets and satellites: detection – planets and satellites: dynamical evolution and stability – stars: individual (HAT-P-11) 1. INTRODUCTION HAT-P-11 is a mid-K dwarf known to host HAT-P- 11b, a super-Neptune on a P = 4.88 day orbit, with MP = 23.4 ± 1.5 M⊕ and RP = 4.36 ± 0.06 R⊕. The planet was first discovered by Bakos et al. (2010) us- ing ground-based photometry and confirmed by radial velocities (RVs), which constrained its mass and eccen- tricity. Bakos et al. (2010) found a moderate eccentricity of e = 0.198 ± 0.046, the first clue that the HAT-P-11 system is dynamically hot. At the time, HAT-P-11b was the smallest planet discovered by ground-based transit photometry. HAT-P-11 was observed by the Kepler Space Tele- scope (Borucki et al. 2010) during its prime mission (2009–2013). Deming et al. (2011) and Sanchis-Ojeda & Winn (2011) analyzed this data and found spot-crossing anomalies at particular phases of the transit of HAT-P- 11b, which are consistent with a nearly polar orbit cross- ing two active latitudes on the host star. This was in agreement with the results from two independent RV campaigns by Winn et al. (2010b) and Hirano et al. (2011), who used the Rossiter-McLaughlin (RM) effect to measure the planet's orbital obliquity to be λ ≈ 100◦. 2 Using the Kepler photometry, Huber et al. (2017) also reported a tentative detection of HAT-P-11b's secondary eclipse. Here, we present an extended RV timeseries spanning 10 years (Section 2), which show a long-period Keplerian signal with P ≈ 9 years. While HAT-P-11 is chromo- spherically active, we show in Section 3 that the RV signal cannot be explained by activity alone. In Section 4, we model the RV time series including the effects of planet b, planet c, and stellar activity. We investigate the dynamical connection between the two planets in Section 5 and find that HAT-P-11c can explain the high obliquity of HAT-P-11b. Finally, we place the HAT-P- 11 system in context of other exoplanet systems (Sec- tion 6) and discuss prospects for future characterization (Section 7). 2. SPECTROSCOPIC OBSERVATIONS The California Planet Search (CPS; Howard et al. 2010) has observed the HAT-P-11 system since 2007 August with the High Resolution Echelle Spectrometer (HIRES; Vogt et al. 1994) at the Keck I 10m telescope on Maunakea. We collected a total of 253 spectra with an iodine cell in front of the spectrometer, which im- prints iodine absorption lines to serve as a wavelength reference against which RVs can be measured precisely. The spectra have signal-to-noise ratios (S/N) between 100 and 130 per pixel on blaze near 5500 Å. 2.1. Radial Velocities We used the standard CPS pipeline described in Howard et al. (2010) to determine the RVs. This in- volves forward modeling the stellar and iodine spectra convolved with the instrumental point spread function for different spectral segments (Marcy & Butler 1992; Valenti et al. 1995). The complete set of RV data is presented in Table 1, with a median uncertainty of 1.4 m s−1. In the subsequent analysis, we have excluded two sets of very high cadence observations taken within 4 hr of the transit of HAT-P-11b, which are affected by the RM effect. This leaves us with 144 remaining RV measurements, which are plotted in Figure 1a. In their original dis- covery, Bakos et al. (2010) reported a significant long- term drift over two years of RV observations, which they interpreted as a possible second planet. With our ex- tended observational baseline of ten years, we see that this long-period trend has reversed, suggesting that we have now viewed a complete orbit of this outer compan- ion. A generalized Lomb-Scargle periodogram (Zech- meister & Kürster 2009) of the raw RVs shows a peak at ∼ 3463 days (Figure 1c), just over 9 years. Table 1. Radial Velocity and Activity Measurements Time BJDTBD σ(RV) SHK Index σ(SHK) Hα Index RV m s−1 m s−1 σ(Hα) Flag 6.50 6.75 2454335.891030 2454335.897680 2454336.746470 2454336.859340 2454336.947330 0.27 2454337.729220 −12.86 8.03 4.30 1.03 1.09 0.94 1.03 1.00 1.14 0.5599 0.0056 0.04539 0.00026 0.5614 0.0056 0.04537 0.00026 0.5748 0.0057 0.04533 0.00025 0.5751 0.0058 0.04531 0.00026 0.5765 0.0058 0.04543 0.00027 0.5886 0.0059 0.04602 0.00028 1 1 1 1 1 1 Note-Radial velocity (RV) and activity measurements calculated from HIRES observations. A 1 in the Flag column indicates that the data point was used in our analysis. Table 1 is published in its entirety in machine-readable format. A portion is shown here for guidance regarding its form and content. 2.2. Stellar Activity Indicators HAT-P-11 is known to be a spotted, chromospher- ically active star (Deming et al. 2011; Morris et al. 2017b). Stellar activity can produce spurious RV sig- nals that may be mistaken for a planet (see e.g. Robert- son et al. 2014; Haywood et al. 2014). To investigate whether stellar activity could account for the 9 year RV signal, we extract two activity indices from our spectro- scopic observations. The Mount Wilson SHK index traces the chromo- spheric emission in the cores of the Ca II H&K lines (Vaughan et al. 1978) and is a standard activity tracer for main-sequence stars. We extract SHK from our spec- tra following the procedure of Isaacson & Fischer (2010), 3 Figure 1. Panel (a): Radial velocity (RV) time series, showing a long-period signal. Panel (b): SHK index time series. Error bars for both panels (a) and (b) are shown but are comparable to the size of the points. Panel (c): A Lomb-Scargle periodogram of the RV data shows strong power at P ≈ 3500 days. This signal and its harmonics dominate the periodogram and overwhelm the 4.88 day signal from the known inner planet, HAT-P-11b. Panel (d): The SHK periodogram has a peak at a similar period. We note the strong signal in both periodograms at 29 days, the inferred rotational period of HAT-P-11. and our measurements are precise to 1%. We also measured the Hα index, which has been found to be a good activity tracer for late-type stars (Gomes da Silva et al. 2011; Robertson et al. 2014). While Hα tracked SHK closely, the size of the variations were on the 1-2% level, comparable to the measurement uncertainty. Therefore, we henceforth use SHK as the activity tracer. Both activity indices for each observation are provided in Table 1. The periodogram of SHK (Figure 1d) has a peak at ∼ 3800 days, close to peak found in the RV periodogram. Morris et al. (2017a,b) also observed this activity signal and interpreted it as a solar-like dynamo. Given the comparable timescales of the RV and activity cycles, we consider whether activity could be responsible for the RV variability in the following section. 3. IS THE LONG-PERIOD RV SIGNAL DUE TO STELLAR ACTIVITY? Our decade of RV observations of HAT-P-11 have re- vealed a long-period signal, suggestive of a planet. Here, we assess whether this signal could be caused by stellar activity. We show that activity is incompatible with the observed 9 year RV signal for three reasons: (1) the ob- served amplitude is much larger than activity-induced RV variability seen in similar stars, (2) there is a signif- icant phase offset between the activity and RV cycles, and (3) the RV-activity correlation is too weak to ac- count for the RV signal. 3.1. Amplitude of RV Signal We first subtracted the effect of HAT-P-11b from the RV time series, using a model generated from the orbital parameters derived by Bakos et al. (2010). The residual RVs are shown in Figure 2 and the remaining long-period signal has a semi-amplitude of ∼ 35 m s−1. Typical activity-induced RV signals are significantly smaller. Isaacson & Fischer (2010) measured the chro- mospheric activity and RV jitter of ∼ 2600 main- sequence and subgiant stars. For the ∼ 300 stars in the sample similar to HAT-P-11 (1.0 < B − V < 1.3), the typical RMS in measured RVs was around 4 − 8 m s−1. In particular, there was no increase in jitter as a func- tion of SHK index, suggesting that K dwarfs do not have significant activity-induced jitter. These findings were corroborated in a similar study by Lovis et al. (2011), who observed 304 FGK stars with HARPS over seven years, finding a maximum activity- induced RV signal of 11 m s−1. This study also found that RV correlation with magnetic activity is minimized in stars with Teff ≈ 4800 K, where even strong magnetic cycles induced RV signals of only several m s−1. Thus, it is unlikely that the ∼ 35 m s−1 RV signal in the HAT-P-11 data could be attributed to stellar ac- 45005000550060006500700075008000604020020RV [m s1](a)45005000550060006500700075008000JD - 24500000.450.500.550.600.65SHK Index(b)1001011021031041050.00.20.40.6RV Periodogram PowerP = 3463d(c)100101102103104105Period (days)0.00.10.20.30.4SHK Periodogram PowerP = 3827d(d) 4 tivity alone, as it is more than three times larger than previously known activity-induced signals, particularly when we consider the reduced sensitivity of RVs to chro- mospheric activity in K dwarfs. 3.2. RV-Activity Phase Offset Another line of reasoning favoring the planet interpre- tation is the phase offset between the RV and SHK cycles. Activity-induced RV signals arise due to suppression of convective blueshift, primarily by plages (e.g. Haywood et al. 2014; Dumusque et al. 2014). Because the SHK in- dex measures chromospheric Ca II H&K emission, it is a direct measure of plage activity (Shine & Linsky 1974). Hence, any activity-induced RV signal should move in lockstep with the SHK activity indicator, without any phase offset. The presence of spots may cause a phase shift be- tween the SHK and induced RV signal, due to masking of parts of the star that are rotationally blue- or red- shifted (Haywood et al. 2014). The maximum offset be- tween the two signals due to rotation is only a fraction of the stellar rotation period (29 days) and is therefore negligible when compared to the 9 year period of the RV signal. Inspection of Figure 1 shows that the RV and SHK time series reach their respective minima at times that differ by more than a year. To measure the significance of this offset, we used the publicly available RadVel soft- ware package (Fulton et al. 2018)1 to fit a Keplerian model to the residual RVs described in Section 3.1, and to the SHK indices (Figure 2). For the RVs, we measured an eccentricity of e = 0.565 ± 0.035, period of 3334 ± 220 days, and a peri- astron passage of JD = 2456859+22−31. If this signal were in fact due to stellar activity, we would expect the shape and period of the SHK cycle to be similar. We thus fit the SHK time series with another Keplerian using priors on eccentricity and period corresponding to the RV fit. For the SHK indices, we measure a "periastron passage" of JD = 2457271+28−34, more than 400 days after tp of the RV signal. This corresponds to a phase offset of ∼ 12%, a difference of > 10σ. There is no physical basis to expect such a 400 day offset between the long-period SHK and RV cycles. This suggests that their apparent similarity is no more than a coincidence, rather than a causative relationship be- tween stellar activity and measured RVs. 3.3. RV-SHK Correlation Finally, if the long-period RV variation was indeed due to stellar activity, they should be correlated across the 1 https://radvel.readthedocs.io Figure 2. Top: The maximum-likelihood Keplerian model fit to the residual RVs, after removing the effect of the inner planet. Vertical dashed lines mark the 1-σ confidence inter- val for the time of periastron passage. Bottom: Using the model parameters and uncertainties derived from the RV fit as priors, we fit a Keplerian to the SHK indices. The time of periastron passage for this model is 412 days later, demon- strating a significant phase offset between the two signals. Figure 3. Residual RVs as a function of SHK index. While there is some correlation between the RVs and stellar activity, the low Pearson's r statistic suggests that only ∼ 34% of the total variation could be part of an activity-induced signal. entire dataset. Figure 3 shows the residual RVs after removing the effect of planet b as a function of SHK index. While there exists a weak linear correlation, the Pearson's r statistic is only 0.34, indicating that up to a third of the total RMS variation in RVs can be accounted for by stellar activity. The remaining variation, reflected in the large scatter around the fitted line, must be due to another mechanism. 3.4. Summary Stellar activity alone is insufficient to explain the RV variability of HAT-P-11. The amplitude of the RV vari- 45005000550060006500700075008000604020020Residual RV [m/s]Planet b removed45005000550060006500700075008000JD - 24500000.150.100.050.000.05SHK Index (Arbitrary Offset)0.450.500.550.600.65SHK Index604020020Residual RV (m/s)Planet b removed 5 Figure 4. Panel (a): Residual RVs, after removing the effect of the inner planet. We identified four different intervals of up to 180 days each with relatively high observational cadence, over which the RV signal of any putative outer planet can be neglected. These four intervals are marked with colored points. Panel (b): SHK indices, with the same observation periods marked. Panel (c): For each identified period, a strong correlation between residual RV and SHK can be observed, with the slope consistent over all four seasons. Panel (d): Same as panel (c) but with the effects of both planets b and c removed. The offsets between the four observing seasons have vanished, such that the correlations from each season are now fully consistent with each other. For clarity, the errors in SHK index are not shown in panels (c) and (d). ation is too large to be caused solely by stellar activity (Section 3.1), and the activity cycle is offset from the RV signal by more than a year (Section 3.2). The correla- tion between the residual RVs and SHK indices also show that most of the RV variation cannot be attributed to stellar activity (Section 3.3). We therefore subsequently adopt a two-planet interpretation for the data. 4. RV MODELING Here, we describe our modeling of the HAT-P-11 RVs that includes contributions from two planets. The weak RV-SHK correlation, described in Section 3.3, motivated an analysis that simultaneously includes the effects of stellar activity. To better understand the connection between the residual RVs and stellar activity, we identify four seasons of < 180 days with at least 15 observations, over which long-period variations can be neglected (Figure 4). For each of these four intervals, we find much stronger lin- ear correlations between the SHK index and residual RVs than the correlation present in the full set of observa- tions. The Pearson's r statistics were > 0.5 for all in- tervals, and the p-values were less than 5%. We also observe that the high-cadence segments have different 45005000550060006500700075008000604020020Residual RV (m/s)Planet b removed(a)45005000550060006500700075008000JD - 24500000.450.500.550.600.65SHK Index(b)0.450.500.550.600.65SHK Index604020020Residual RV (m/s)(c)Planet b removed0.450.500.550.600.65SHK Index(d)Planets b and c removed 6 mean RVs, but the correlations have consistent slopes as determined by a bootstrap resampling. This suggests that stellar activity has a small but consistent effect on the RV measurements, but it cannot explain the offsets between observing seasons, since they occur at the same SHK values but have mean RVs that differ by more than 50 m s−1. Given this short-timescale RV-activity correlation, we modeled the data using a two-planet Keplerian model as well as a linear correlation between the RVs and SHK. We used the RadVel package (Fulton et al. 2018) to per- form maximum-likelihood fitting and MCMC parameter estimation. We fixed the period and time of conjunction for HAT- P-11b according to the values derived by Huber et al. (2017) from four years of Kepler data. The remaining orbital parameters for planets b and c, as well as an average RV offset, γ, were allowed to float. We parame- terized e and ω of each planet as e sin ω to guard against a bias toward non-zero eccentricities as recommended by Eastman et al. (2013). We also imposed a beta distribution prior for the eccentricities recommended in Kipping (2013). e cos ω and √ √ The slope of the activity-RV correlation is a new free parameter, cS, such that the induced RV signal is cS∆SHK,i. Here, ∆SHK,i ≡ SHK,i − SHK is the mean- centered SHK index at time ti. Any further constant offset is absorbed into γ. The likelihood is (cid:34) (cid:88) lnL = − 1 2 (vi − vm,i − cS∆SHK,i)2 i σ2 i + σ2 jit + ln 2π(σ2 i + σ2 jit) where vi and vm,i are the measured and model RVs at time ti, σi is the corresponding uncertainty on the mea- sured RV, and σjit is the jitter. The results of our RV fit are shown in Figure 5 and the derived planetary parameters are given in Table 2. We also provide the posterior distributions from our MCMC analysis in Appendix C. We find that HAT-P-11c is a MP sin i = 1.60+0.09−0.08 MJ giant planet with semima- jor axis of a = 4.13+0.29−0.16 AU. Its high-eccentricity or- bit (e = 0.601+0.032 −0.031) gives it a periastron distance of 1.67+0.14−0.13 AU and an apoastron distance of 6.61+0.52−0.30 AU. This large separation reached at apoastron will have a positive effect on any future attempts to detect the planet via direct imaging, as we discuss in Section 7. Once the effect of both planets is removed (Figure 4d), residual RVs show a strong linear correlation with the SHK values (Pearson's r = 0.479, p-value = 8 × 10−9), where the offsets between the four observing seasons are eliminated. The total semi-amplitude of the activity- induced RV is ∼ 7 m s−1, consistent with that observed in stars of similar spectral types (see Section 3.1). We also investigated models with higher and lower complexity. We first examined a single-planet model with activity as well as a two-planet model without activity correction. These models were not favored when compared using the Bayesian Information Crite- rion (BIC; Schwarz 1978). We also considered the possi- bility of additional planets in the system, but these were not found by a two-dimensional Keplerian Lomb-Scargle (2DKLS; O'Toole et al. 2009) periodogram search. We describe these model comparisons in detail in Appendix A. Table 2. System Parameters (cid:35) A B B B B B B B C 0.683 ± 0.009 0.809+0.02−0.03 4780 ± 50 +0.31 ± 0.05 6.57 ± 0.09 1.5 ± 1.5 29.2 6.5+5.9−4.1 37.89 ± 0.33 ≡ 4.887802443 ≡ 2454957.8132067 0.218+0.034 −0.031 19+14−16 23.4 ± 1.5 0.05254+0.00064 −0.00066 4.36 ± 0.06 0.0413+0.0018 −0.0019 2454957.15+0.17−0.20 0.0637+0.0020 −0.0019 2454959.60+0.17−0.20 Stellar Parameters R(cid:63) (R(cid:12)) M(cid:63) (M(cid:12)) Teff (K) [Fe/H] V (mag) v sin i (km s−1) Prot (d) Age (Gyr) Distance (pc) Planetary Parameters Planet b P (days) Tconj (JD) e ω (◦) MP sin i (M⊕) a (AU) RP (R⊕) rperi (AU) Tperi (JD) rapo (AU) Tapo (JD) Planet c P (days) E Tconj (JD) E E e ω (◦) E MP sin i (M⊕) E a (AU) E rperi (AU) E Tperi (JD) E rapo (AU) E Tapo (JD) E A: Deming et al. (2011) B: Bakos et al. (2010) C: Gaia Collaboration et al. (2016a) D: Huber et al. (2017) E: This work 3407+360−190 2456746+24−32 0.601+0.032 −0.031 143.7+4.8−4.9 507+30−27 4.13+0.29−0.16 1.67+0.14−0.13 2456862+20−26 6.61+0.52−0.30 2458565+166−87 D D E E E E D E E E E 7 Figure 5. Two-planet Keplerian fit to the RVs, including a linear correlation between SHK and radial velocities. Panel (a): The most probable model and full radial velocity time series. Panel (b): Residuals from the most probable model, after removing the effect of both planets and the SHK decorrelation. Panels (c) and (d): Phase-folded RVs and the most probable model for planets b and c respectively, with contributions of the other planet and SHK decorrelation removed. The large red circles show phase-binned RVs. Finally, to ensure our methodology does not always favor planets over activity, we applied an identical anal- ysis to the HD99492 system, another active mid-K dwarf with long-period activity and RV signals (Appendix B). In this case, the BIC rejects a planetary explanation for the RVs and prefers a pure stellar activity model, in agreement with the findings of Kane et al. (2016). 5. SYSTEM DYNAMICS AND SPIN-ORBIT MISALIGNMENT The orbit of HAT-P-11b is known to be misaligned with its host star's spin axis, with an obliquity of λ ≈ 100◦, corresponding to a nearly polar orbit (see Section 1). There are a number of other planets with misaligned orbits (see Albrecht et al. 2012; Dai & Winn 604020020406080RV [m s1]a)HIRES20082010201220142016Year45005000550060006500700075008000JD - 24500001260612Residualsb)0.40.20.00.20.4Phase201001020RV [m s1]c)Pb = 4.89 daysKb = 10.42 ± 0.65 m s1eb = 0.218 ± 0.033 0.40.20.00.20.4Phase402002040RV [m s1]d)Pc = 3407 ± 270 daysKc = 30.9 ± 1.2 m s1ec = 0.601 ± 0.032 8 2017). Many explanations have been proposed for such misalignments, including Kozai-Lidov cycles (e.g. Fab- rycky & Tremaine 2007), planet-planet scattering (e.g. Nagasawa et al. 2008), primordial tilting of the proto- planetary disk (e.g. Batygin 2012), or angular momen- tum transport by internal gravity waves (e.g. Rogers et al. 2012). Here, we examine the dynamical coupling between HAT-P-11b and c and assess if it can explain the observed misalignment. The orbital angular momentum of HAT-P-11c is much greater than that of HAT-P-11b and the star's spin an- gular momentum, allowing us to make the approxima- tion that the orbital plane of planet c is invariant. As a matter of convenience, we define angles that describe the orientation of planet b's orbit, inclination ib and argument of periastron ωb, with respect to the orbital plane of planet c. Note that this reference plane is not the sky plane, which is often used to describe the or- bits of transiting planets. In this coordinate system, ib is therefore the relative inclination between the two planets. Following Mardling (2010), we write down the orbit-averaged Hamiltonian for the interaction of the two planets, expanded to quadrupole order in semimajor axis ratio (Kaula 1964):2 (cid:19)2(cid:32) (cid:18) ab (cid:19)(cid:18) 3 ac (cid:33)3 1(cid:112)1 − e2 (cid:19) c Gmbmc ac H = 1 4 (cid:20)(cid:18) 3 2 15 4 cos2 ib − 1 2 2 + e2 b 1 + b sin2 ib cos 2ωb e2 (1) The second term, containing e2 b sin2 ib, gives rise to the Kozai-Lidov mechanism, in which the inner planet un- dergoes cycles trading large inclinations for large eccen- tricities. However, because HAT-P-11b is very close in to its star, general relativistic (GR) effects cause apsidal precession, which may suppress Kozai-Lidov oscillations under suitable conditions (e.g. Ford et al. 2000; Fab- rycky & Tremaine 2007). Hence, we must include an additional GR term in the Hamiltonian. We can write the Hamiltonian including the GR term using scaled canonical Delaunay variables, where Ωb is the longitude of ascending node of planet b: (cid:113) (cid:113) G = 1 − e2 1 − e2 b b cos ib H = g = ωb h = Ωb (cid:33)3(cid:34)(cid:0)5 − 3G2(cid:1)(cid:0)3H 2 − G2(cid:1) G2 GM(cid:63) abc2 1 G . + 3nb ab nb mb 1 16 mc M(cid:63) (cid:112)1 − e2 √GM(cid:63)ab, giving (cid:32) H(cid:48) = 15(cid:0)1 − G2(cid:1)(cid:0)G2 − H 2(cid:1) cos 2g mean motion nb =(cid:112)GM(cid:63)/a3. G2 ac + c (cid:35) (2) Here, we have written the expression in terms of the The rapid apsidal precession due to GR may suppress the Kozai resonance, which requires a slowly varying ωb. We calculate the GR precession rate, ωGR = 3nb GM(cid:63) abc2 1 G2 ≈ 2.2 × 10−4 yr−1 (cid:21) which gives a precession period of approximately 30,000 years. In comparison, the Kozai timescale is given by (Kise- leva et al. 1998): (cid:0)1 − e2 c (cid:1)3/2 2P 2 c 3πP 2 b τ = ≈ 4 × 105 years, M(cid:63) mc . an order of magnitude longer. Thus, we expect that the Kozai mechanism is suppressed in this system. To confirm this, we examine the phase space of the Hamiltonian (2). The Hamiltonian admits two integrals of motion: H as well as H(cid:48) itself. Thus, any given phase- space portrait is parameterized by H, which translates to a particular imax, the inclination of the inner planet attained when its orbit is circular. Along level curves of H(cid:48), the variables G, g trace out trajectories in a two- dimensional phase space, where the eccentricity is given 1 − G2, which specifies the instantaneous in- by eb = clination via the conservation of H. √ We plot the phase-space portraits for two different val- ues of imax with and without GR, projected into non- canonical coordinates eb cos ωb and eb sin ωb in Figure 6. In agreement with the simple timescale argument pre- sented above, we find that the fast precession of ωb in the HAT-P-11 system is sufficient to suppress Kozai os- cillations, such that there is no libration of eb for any value of imax. We note that simply because the Kozai effect does not operate within the present-day architecture of the HAT- P-11 system does not rule out the possibility that it could have operated previously. This requires the semi- major axis of planet b to have been larger in the past and to have shrunk to its current configuration due to tidal friction (e.g. Fabrycky & Tremaine 2007). How- ever, the tidal migration timescale is longer than the tidal circularization timescale by a factor of 1/(1 − e2) and correspondingly scaling the Hamiltonian by 2 The semimajor axis ratio in this sytem is ab ac ing a leading-order trunctation of the Hamiltonian. ≈ 0.01, warrant- 9 Figure 6. Phase-space portraits for the two-planet Hamilton in eb, ωb space for different values of imax. Blue trajectories indicate circulation, where e remains roughly constant while ω precesses; red trajectories indicate libration of e. The dashed gray circle shows the current observed eccentricity of HAT-P-11b. Panels (a), (b): At low maximal inclinations imax, only circulatory trajectories exist, where eb remains roughly constant while ωb precesses. Panel (c): GR precession suppresses libratory trajectories even at high values of imax. Panel (d): If GR precession is neglected, the Kozai mechanism can occur, with libratory trajectories taking the planet to high eccentricity and inclination. (Hut 1981). Thus, tides would tend to circularize the orbit faster than they shrink the orbit. Given that the orbit is still eccentric, we consider it unlikely that tidal damping has shrunk the orbit of HAT-P-11b. Given the lack of Kozai cycles, we can average out the Kozai term, which leaves us with a trivial dynamical sys- tem, governed by a Hamiltonian that only depends on the actions, and thus yields only precession as its con- sequence. The longitude of ascending node then evolves according to dΩb dt = = (cid:32) ∂H(cid:48) ∂H 1 2 nb mc M(cid:63) ac (cid:112)1 − e2 ab c (cid:33)3(cid:18) 15 − 9G2 (cid:19) G2 (3) H. Thus, Ωb precesses around the invariant plane defined by the outer planet's orbit, with a period of approxi- mately 3.5 Myr, significantly shorter than the age of the system. If the orbit normal of planet c is misaligned with the spin axis of the star by more than half the cur- rent observed obliquity of planet b, ic (cid:38) 50◦, this would be sufficient to explain HAT-P-11b's approximately po- lar orbit (Figure 7). However, this does not explain the initial misalignment of HAT-P-11c, which may be the result of planet-planet scattering in the outer system. We also check that the stellar spin axis does not it- self precess more quickly than the inner planet, as such an arrangement would result in a coupling between the star and the inner planet, allowing both to precess to- gether and remain aligned. We follow Spalding & Baty- gin (2015) and model the oblateness of the star as a WithGRTermimax35degaApsidalPrecessionApsidalLibrationCurrentEccentricity(e=0.2)-1.0-0.50.00.51.0-1.0-0.50.00.51.0ebcosωbebsinωbWithoutGRTermimax35degb-1.0-0.50.00.51.0-1.0-0.50.00.51.0ebcosωbebsinωbWithGRTermimax80degc-1.0-0.50.00.51.0-1.0-0.50.00.51.0ebcosωbebsinωbWithoutGRTermimax35degd-1.0-0.50.00.51.0-1.0-0.50.00.51.0ebcosωbebsinωb 10 plains the unusual polar orbit of HAT-P-11b. Through precession around the outer planet's orbital plane, HAT- P-11b can attain very high obliquities with respect to the stellar rotation axis, although the angle with respect to the invariant plane remains fixed. A measurement of the mutual inclination between the planetary orbits, for ex- ample, through astrometry (Section 7), could help shed more light on this explanation. Nonetheless, irrespective of the exact scenario, this system would have required a large degree of primordial misalignment, either between the orbits of the two planets as described here, or be- tween the stellar spin axis and HAT-P-11b. 6. THE HAT-P-11 SYSTEM IN CONTEXT Among the planets that have measured obliquities, HAT-P-11b is an outlier. It has the smallest planetary to stellar mass ratio and one of the lowest host star ef- fective temperatures for a misaligned planet (Figure 8). Winn et al. (2010a) first noted a connection be- tween high stellar obliquities and effective temperature, with a higher proportion of planets around hot stars (Teff (cid:38) 6000 K) with misaligned orbits. They suggested that this could be due to the fact that cool stars have larger convective zones, creating strong tidal coupling with their close-in planets that realigns the star to the planet's orbit normal. In contrast, hot stars without these large convective zones have weaker tidal coupling, leading to a longer tidal realignment timescales. Indeed, among the cool stars (Teff < 6000 K) with obliquity measurements, most of the systems exhibiting significant misalignment also have large a/R(cid:63) (Figure 8). This corresponds to a longer realignment timescale, allowing systems to retain any primordial inclination. Seen in this light, HAT-P-11b is no longer an outlier, with a/R(cid:63) = 16.3±0.4. Albrecht et al. (2012) calculated the characteristic realignment timescale for the system to be ∼ 1015 years, vastly longer than the age of the system. Hence, while the secular precession of HAT- P-11b's orbit normal may be a plausible reason for the observed misalignment, it is not a necessary condition since any primordial misalignment of HAT-P-11b would have also been retained. While nodal precession is not necessary to maintain HAT-P-11b's misalignment for Gyr timescales, it may be for shorter-period planets like HATS-14b (Zhou et al. 2015). Another system for which this mechanism may be at work is WASP-8, which is reminiscent of HAT-P- 11 in the following respects: WASP-8 is a cool star with a close-in misaligned planet and a distant giant planet on an eccentric orbit (Knutson et al. 2014). Nodal precession is likely one of several ways to pro- duce misaligned planets. For example, Kepler-420b (Santerne et al. 2014) and HD 80606b (Hébrard et al. 2010) have stellar companions, suggesting star-planet Figure 7. Precession of the longitude of ascending node Ωb of HAT-P-11b's orbit (blue) around the orbital plane of HAT- P-11c (brown) can result in a polar orbit for planet b. (1) Initially, the inner planet's orbit is aligned with the star's rotation axis (black arrow), but the outer planet has an in- clination ∼ 50◦ (normal to orbital plane shown as brown arrow). (2) As time progresses, the longitude of ascending node precesses around the plane of the outer planet's orbit, increasing the inclination of the inner planet's orbit relative to the stellar rotation axis. (3) The inner planet reaches a maximum inclination twice that of the outer planet relative to the stellar rotation axis. point mass surrounded by an orbiting ring with effec- tive semimajor axis (cid:21)1/3 (cid:20) 16ν2k2 2R6 (cid:63) 9I 2GM(cid:63) a = . (4) Here, ν is the stellar rotation frequency, and we have used the dimensionless moment of inertia I = 0.08 and Love number k2 = 0.01. Using Equation 3 for the pre- cession rate due to the torque from the inner planet on the star, we confirm that the stellar precession timescale is on the order of 100 Myr, much slower than the pre- cession of the inner planet. This application of secular theory to the HAT-P-11 system presents a plausible dynamical history that ex- 𝛀Time𝛀Longitude of ascending node precessesaround planet c's orbital plane∼1.8 Myr𝛀(3) Maximally misalignedorbit after 180°precession(2) Obliquity increasesas planet b precesses(1) Initially aligned orbit(low obliquity)Direction of observer 11 Figure 8. HAT-P-11b in the context of other planets with measured obliquities.c Left: Effective temperature and projected stellar obliquities. For clarity, error bars are only shown for planets misaligned by 10◦ or more. A large fraction of planets orbiting stars hotter than ∼ 6000 K are misaligned, perhaps due to the absence of convective envelopes around these stars (see Section 6). HAT-P-11b is one of only a handful of misaligned planets orbiting a star cooler than 6000 K. Right: Obliquities for stars with Teff < 6000 K as a function of a/R(cid:63). Close-in planets tend to be aligned, but this preference vanishes for a/R(cid:63) (cid:38) 15. The mapping between point color and Teff is the same as in the left panel. rather than planet-planet mechanisms. However, if planet-planet nodal precession is a common mechanism to produce misaligned orbits, we predict a correlation between misaligned planets around cool stars and dis- tant, eccentric giants. Such a prediction is testable with future RV, astrometric, or imaging follow-up of known misaligned systems as well as the many more that will soon be discovered by the Transiting Exoplanet Survey Satellite (TESS; Ricker et al. 2014). 7. PROSPECTS FOR FUTURE OBSERVATIONS 7.1. Secondary Eclipse of HAT-P-11b Secondary eclipse observations can provide insight into the albedo and thermal structure of a planet's at- mosphere. Huber et al. (2017) reported the detection of the secondary eclipse of HAT-P-11b based on Ke- pler photometry, at an orbital phase of φ = 0.659 and depth of 5 ppm. Our adopted RV model found an ex- pected secondary eclipse phase of φ = 0.623+0.018 −0.019, con- sistent within 2σ of the Huber et al. (2017) detection. Future observations with the James Webb Space Tele- scope (Gardner et al. 2006) should be able to detect and characterize this secondary eclipse. 7.2. Astrometric Characterization of HAT-P-11c The Gaia mission is currently making extremely high- precision astrometric observations of one billion astro- nomical objects. For HAT-P-11, Gaia is expected to 3 Data compiled from TEPCat 2017 (Southworth 2011, http://www.astro.keele.ac.uk/jkt/tepcat/ rossiter.html). of October as reach an astrometric precision of ∼ 7µas (Gaia Collab- oration et al. 2016b) by the end of its five-year nominal mission. The wide orbit (a = 4.13+0.29−0.16 AU) of HAT-P-11c and the system's relative proximity (37.89 ± 0.33 pc) to Earth means that HAT-P-11c will yield a maximum astrometric signal of 264+42−29 µas. The presence of the planet should in principle be detectable by Gaia during its nominal five-year mission, while a full determination of the orbit's three-dimensional orientation will be pos- sible if the mission duration is extended to 10 years. Although such a measurement will not uniquely deter- mine the mutual inclination of the two planets, it will constrain the difference in angles in a single plane and could help verify the dynamical picture described in Sec- tion 5. orbit up takes 7.3. Direct Imaging of HAT-P-11c it eccentric HAT-P-11c's to 6.61+0.52−0.30 AU from its host star. Taking into ac- count the argument of periastron of the orbit, the maximum sky-projected planet-star separation is 134+15−10 mas. Assuming a radius and albedo similar to that of Jupiter, the reflected-light contrast ratio of the planet will be ∼ 6 × 10−9. This makes HAT-P-11c a poten- tial although challenging target for high-contrast direct imaging studies. For example, the Wide Field Infrared Survey Telescope (WFIRST, Spergel et al. 2015) is ex- pected to have a 10−9 effective contrast and 100 mas inner working angle, and may be able to characterize HAT-P-11c. 40005000600070008000Teff (K)0306090120150180Projected obliquity (deg)HAT-P-11bWASP-107bHAT-P-18bHATS-14bKepler-420bHD 80606bKepler-63bWASP-8b5101520Aligned=0Polar =90Retrograde=180Polar =270a/R*HAT-P-11bHAT-P-17bKepler-63bWASP-8bWASP-107bHAT-P-18bHATS-14b 12 8. CONCLUSION The HAT-P-11 system is one of the best-studied exo- planet systems, with long baseline RV and photometric datasets and RM measurements. Here, we extend the RV baseline to ten years and discover a new 1.5 MJ gi- ant plant on a distant, eccentric orbit. We found that the presence of HAT-P-11c may explain the previously known misalignment of HAT-P-11b through nodal pre- cession. This mechanism may help to explain the di- versity of exoplanet obliquities as a function of orbital distance and host star type. Further characterization of the HAT-P-11 system will soon be possible thanks to upcoming spectroscopic, astrometric, and imaging facil- ities. We thank Joshua Winn for helpful discussions that improved the final manuscript. We thank the many observers who contributed to the measurements re- ported here including M. Bottom, B. Bowler, P. Butler, J. Brewer, J. Crepp, C. Chubak, K. Clubb, I. Crossfield, D. Fischer, E. Ford, E. Gaidos, M. Giguere, J. Johnson, M. Kao, G. Marcy, T. Morton, G. Mandushev, K. Peek, S. Pineda, G. Torres, S. Vogt, P. Worden, M. Zhao. S. W. Y. acknowledges support from the Caltech Mar- cella Bonsall Summer Undergraduate Research Fellow- ship. E. A. P. acknowledges support from a Hubble Fellowship grant HST-HF2-51365.001-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astron- omy, Inc. for NASA under contract NAS 5-26555. The data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partner- ship among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation. The authors wish to recognize and acknowledge the very significant cultural role and rever- ence that the summit of Maunakea has long had within the indigenous Hawaiian community. We are most for- tunate to have the opportunity to conduct observations from this mountain. Software: Numpy/Scipy (Van Der Walt et al. 2011), Matplotlib (Hunter 2007), Pandas (McKinney 2010), Astropy (Astropy Collaboration et al. 2013), emcee (Goodman & Weare 2010; Foreman-Mackey et al. 2013), RadVel (Fulton et al. 2018) Facility: Keck:I (HIRES) REFERENCES Albrecht, S., Winn, J. N., Johnson, J. A., et al. 2012, ApJ, 757, Goodman, J., & Weare, J. 2010, Communications in Applied 18 Mathematics and Computational Science, 5, 65 Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. Haywood, R. D., Collier Cameron, A., Queloz, D., et al. 2014, 2013, A&A, 558, A33 Bakos, G. Á., Torres, G., Pál, A., et al. 2010, ApJ, 710, 1724 Batygin, K. 2012, Nature, 491, 418 Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977 Dai, F., & Winn, J. N. 2017, AJ, 153, 205 Deming, D., Sada, P. V., Jackson, B., et al. 2011, ApJ, 740, 33 Dumusque, X., Boisse, I., & Santos, N. C. 2014, ApJ, 796, 132 Eastman, J., Gaudi, B. S., & Agol, E. 2013, PASP, 125, 83 Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298 Ford, E. B., Kozinsky, B., & Rasio, F. A. 2000, ApJ, 535, 385 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306 Fulton, B., Blunt, S., Petigura, E., et al. 2018, California-Planet-Search/radvel: Version 1.1.0, , , doi:10.5281/zenodo.1143674 MNRAS, 443, 2517 Hébrard, G., Désert, J.-M., Díaz, R. F., et al. 2010, A&A, 516, A95 Hirano, T., Narita, N., Shporer, A., et al. 2011, PASJ, 63, 531 Howard, A. W., & Fulton, B. J. 2016, PASP, 128, 114401 Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, ApJ, 721, 1467 Huber, K. F., Czesla, S., & Schmitt, J. H. M. M. 2017, A&A, 597, A113 Hunter, J. D. 2007, Computing in Science and Engineering, 9, 90 Hut, P. 1981, A&A, 99, 126 Isaacson, H., & Fischer, D. 2010, ApJ, 725, 875 Kane, S. R., Thirumalachari, B., Henry, G. W., et al. 2016, ApJL, 820, L5 Kaula, W. M. 1964, Reviews of Geophysics and Space Physics, 2, Fulton, B. J., Petigura, E. A., Blunt, S., & Sinukoff, E. 2018, 661 ArXiv e-prints, arXiv:1801.01947 Fulton, B. J., Weiss, L. M., Sinukoff, E., et al. 2015, ApJ, 805, Kipping, D. M. 2013, MNRAS, 434, L51 Kiseleva, L. G., Eggleton, P. P., & Mikkola, S. 1998, MNRAS, 175 300, 292 Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2016a, Knutson, H. A., Fulton, B. J., Montet, B. T., et al. 2014, ApJ, A&A, 595, A2 785, 126 Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al. 2016b, Lovis, C., Dumusque, X., Santos, N. C., et al. 2011, ArXiv A&A, 595, A1 Gardner, J. P., Mather, J. C., Clampin, M., et al. 2006, SSRv, 123, 485 e-prints, arXiv:1107.5325 Marcy, G. W., & Butler, R. P. 1992, PASP, 104, 270 Marcy, G. W., Butler, R. P., Vogt, S. S., et al. 2005, ApJ, 619, Gomes da Silva, J., Santos, N. C., Bonfils, X., et al. 2011, A&A, 570 534, A30 Mardling, R. A. 2010, MNRAS, 407, 1048 13 McKinney, W. 2010, in Proceedings of the 9th Python in Science Conference, ed. S. van der Walt & J. Millman, 51 – 56 Meschiari, S., Laughlin, G., Vogt, S. S., et al. 2011, ApJ, 727, 117 Morris, B. M., Hebb, L., Davenport, J. R. A., Rohn, G., & Shine, R. A., & Linsky, J. L. 1974, SoPh, 39, 49 Southworth, J. 2011, MNRAS, 417, 2166 Spalding, C., & Batygin, K. 2015, ApJ, 811, 82 Spergel, D., Gehrels, N., Baltay, C., et al. 2015, ArXiv e-prints, Hawley, S. L. 2017a, ApJ, 846, 99 Morris, B. M., Hawley, S. L., Hebb, L., et al. 2017b, ApJ, 848, 58 Nagasawa, M., Ida, S., & Bessho, T. 2008, ApJ, 678, 498 O'Toole, S. J., Tinney, C. G., Jones, H. R. A., et al. 2009, MNRAS, 392, 641 Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2014, in Proc. SPIE, Vol. 9143, Space Telescopes and Instrumentation 2014: Optical, Infrared, and Millimeter Wave, 914320 Robertson, P., Mahadevan, S., Endl, M., & Roy, A. 2014, Science, 345, 440 arXiv:1503.03757 Valenti, J. A., Butler, R. P., & Marcy, G. W. 1995, PASP, 107, Van Der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, ArXiv e-prints, arXiv:1102.1523 Vaughan, A. H., Preston, G. W., & Wilson, O. C. 1978, PASP, Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, 2198, 362 Winn, J. N., Fabrycky, D., Albrecht, S., & Johnson, J. A. 2010a, 966 90, 267 ApJL, 718, L145 723, L223 L16 Rogers, T. M., Lin, D. N. C., & Lau, H. H. B. 2012, ApJL, 758, Winn, J. N., Johnson, J. A., Howard, A. W., et al. 2010b, ApJL, Sanchis-Ojeda, R., & Winn, J. N. 2011, ApJ, 743, 61 Santerne, A., Hébrard, G., Deleuil, M., et al. 2014, A&A, 571, Zechmeister, M., & Kürster, M. 2009, A&A, 496, 577 Zhou, G., Bayliss, D., Hartman, J. D., et al. 2015, ApJL, 814, L6 A37 Schwarz, G. 1978, Annals of Statistics, 6, 461 APPENDIX A. MODEL COMPARISON AND SELECTION In Section 4, we adopted an RV model incorporating the effect of two planets and an activity-induced signal. This model was chosen after exploring two other possibilities: (1) A single-planet model where the long-term RV trend must be completely accounted for by the activity-RV correlation, and (2) A two-planet model without any activity-RV correlation. Model (3) is the two-planet plus activity model. To compare the models, we compute the BIC, which incorporates the log-likelihood of the model and a penalty for the number of free parameters. A model with lower BIC is preferred, with ∆BIC (cid:38) 10 being strongly favored. We present the results of the RV fits in Table A1. We found that the single-planet model (1) gave the poorest fit, with RMS residuals of almost 17 m s−1 and a high BIC. Models (2) and (3) provided a significant improvement in the RMS residuals, indicating that the long-period signal can be best explained by the presence of an outer planet. The model parameters found by both models for the two planets are similar, typically differing by less than 1-σ. However, model (3) had the lowest BIC (∆BIC = −16) despite the additional model complexity. This is in accordance with our conclusions from Section 3, where we saw that stellar activity does account for some, but not all, of the RV variation. The BIC analysis thus supports our choice of a two-planet RV model with a linear activity-RV correction. We also investigated more complex models to determine if there were additional planets in the system. We performed an iterative search using the two-dimensional Keplerian Lomb-Scargle (2DKLS) periodogram (O'Toole et al. 2009) following the technique of Fulton et al. (2015) and Howard & Fulton (2016). An empirical False Alarm Probability (eFAP) is calculated from a histogram of periodogram amplitudes. Figure A1 shows the periodogram for a three- versus two-planet model. We find no significant peaks above the eFAP threshold of 1%, with only minor peaks close to the stellar rotation period of 29 days and its aliases. Thus, inclusion of a third planet is not justified given the current dataset. B. COMPARISON WITH HD 99492 To validate the joint RV-activity methodology described in section 4, we applied the same analysis to the HD 99492 system, another moderately active mid-K dwarf with a decade of RV measurements. Similar to HAT-P-11, HD 99492 has a relatively short-period (P = 17.1 days) inner planet first discovered by Marcy et al. (2005). Meschiari et al. (2011) later attributed a 5000 day RV signal to a distant giant planet. However, the star's activity cycle, as traced by the SHK index, was found to have similar periodicity by Kane et al. (2016), who subtracted the effect of the inner planet from the RVs and found a strong correlation between these residual RVs and the SHK values. They used this to argue that the outer planet reported by Meschiari et al. (2011) should therefore be attributed to stellar activity. We investigated the HD 99492 system using the same joint activity-RV analysis described previously. We used 89 HIRES spectra taken by the CPS program over 13 years, from 2004 to 2017 (Figure B2). A long-term periodic signal can clearly be seen in the SHK measurements, of almost identical period and phase to the RV variability. We then fit these RV measurements using three different models, similar to those used for fitting HAT-P-11: (1) A single-planet model with activity-RV decorrelation; (2) A two-planet model without any decorrelation; and (3) A 14 Table A1. Comparison of RV-Activity Models Number of planets SHK correction Number of free parameters Planet b parameters Pb (days) T conjb (JD) eb ωb (◦) Kb (m s−1) Planet c parameters Pc (days) T conjc (JD) ec ωc (◦) Kc (m s−1) Global parameters cS (m s−1 SHK γ (m s−1) σjit (m s−1) Model comparison RMS residuals (m s−1) BIC −1) Model 1 Model 2 Model 3 (Adopted) 2 Yes 11 1 Yes 6 2 No 10 ≡ 4.887802443 ≡ 2454957.8132067 0.19+0.14−0.13 33+44−57 10.0+2.2−2.3 0.223+0.036 −0.034 18+15−17 10.51+0.72−0.71 0.218+0.034 −0.031 19+14−16 10.42+0.64−0.66 - - - - - 3335+280−160 2456739+30−39 0.560+0.034 −0.035 140.0+4.8−4.9 32.1±1.4 3407+360−190 2456746+24−32 0.601+0.032 −0.031 143.7+4.8−4.9 30.9±1.3 92+43−44 -10.2±1.5 17.3+1.1−1.0 - 0.223+0.036 −0.034 5.40+0.38−0.34 78.6+16.8−16.1 -1.80+0.84−0.79 4.98+0.36−0.32 16.86 1250.91 5.37 941.99 4.98 925.59 Note-cS is the slope of the linear correlation between RV and SHK two-planet model with activity-RV decorrelation. We compare these three models in Table B2. Table B2. Model Comparison for HD 99492 Model 1 Model 2 Model 3 Number of planets SHK correction RMS residuals (m s−1) BIC (Adopted) 1 Full 3.61 502.98 2 None 3.69 524.53 2 Full 3.39 513.52 For the HD 99492 system, the lowest BIC was achieved for the model with only one planet, with the residual RV signal fully accounted for by stellar activity (∆BIC = −11). The slope of the RV-SHK correlation was found to be c1 = 92± 18 m s−1 SHK −1, so the semi-amplitude of the activity-induced RV signal is ∼ 5 m s−1, similar to that found for HAT-P-11. We also noted that when fitting two planets with activity decorrelation (model 3), an MCMC analysis found the RV semi-amplitude of the outer planet to be 1.8+1.2−2.0 m s−1, not significantly different from zero. Finally, a Keplerian fit to the SHK time series found a cycle whose period and phase are consistent with the fit to the RV time series within 1-σ, suggesting that the two signals are indeed correlated. Here, we see that the same analysis as previously applied to HAT-P-11 now readily rejects the presence of an outer planet in the HD 99492 system and supports the integrity of our methodology. 15 Figure A1. Two-Dimensional Keplerian Lomb-Scargle (2DKLS) periogram where a third Keplerian is fit to the residuals after removing HAT-P-11b and c. No additional significant periodic signals could be detected. Figure B2. RV measurements (top) and SHK indices (bottom) for HD 99492. Similar to HAT-P-11, the SHK index exhibits a long-term variation indicative of an activity cycle. C. POSTERIOR DISTRIBUTION OF RV MODEL PARAMETERS We provide in Figure C3 the posterior distributions for each of the model parameters. To derive these distributions, we explored the likelihood surface with an MCMC analysis. 101102103104Period (Days)0.00.10.20.30.40.5PowerP = 27.12DKLS Periodgram for 3-planet vs 2-planet model3000400050006000700080002010010RV [m s1]300040005000600070008000JD - 24500000.200.250.30SHK index 16 Figure C3. Corner plot showing posterior distributions for each model parameter. The first plot in each column shows the single-variable distribution, with the vertical dashed lines denoting the most probable value and 1σ confidence bounds. The remaining plots show joint distributions between each pair of model parameters. j = 1.64+0.840.788.008.258.508.75logperclogperc = 8.13+0.100.050.720.640.560.48ecoscecosc = 0.62+0.040.030.10.20.30.40.5ecosbecosb = 0.42+0.040.054.04.85.66.4jj = 4.91+0.360.32600650700750800Tconjc+2.456e6Tconjc = 2456749.21+22.2330.660306090120c1_svalue_jc1_svalue_j = 72.10+16.3916.700.0040.0000.0040.008 = 0.00+0.000.0027303336KcKc = 30.96+1.311.257.59.010.512.013.5KbKb = 10.48+0.670.650.40.20.00.20.4esinbesinb = 0.17+0.110.134.53.01.50.01.5j0.300.450.60esinc8.008.258.508.75logperc0.720.640.560.48ecosc0.10.20.30.40.5ecosb4.04.85.66.4j600650700750800Tconjc+2.456e60306090120c1_svalue_j0.0040.0000.0040.00827303336Kc7.59.010.512.013.5Kb0.40.20.00.20.4esinb0.300.450.60esincesinc = 0.46+0.050.05
1302.2147
3
1302
2013-05-21T04:51:24
Catastrophic Evaporation of Rocky Planets
[ "astro-ph.EP" ]
Short-period exoplanets can have dayside surface temperatures surpassing 2000 K, hot enough to vaporize rock and drive a thermal wind. Small enough planets evaporate completely. We construct a radiative-hydrodynamic model of atmospheric escape from strongly irradiated, low-mass rocky planets, accounting for dust-gas energy exchange in the wind. Rocky planets with masses < 0.1 M_Earth (less than twice the mass of Mercury) and surface temperatures > 2000 K are found to disintegrate entirely in < 10 Gyr. When our model is applied to Kepler planet candidate KIC 12557548b --- which is believed to be a rocky body evaporating at a rate of dM/dt > 0.1 M_Earth/Gyr --- our model yields a present-day planet mass of < 0.02 M_Earth or less than about twice the mass of the Moon. Mass loss rates depend so strongly on planet mass that bodies can reside on close-in orbits for Gyrs with initial masses comparable to or less than that of Mercury, before entering a final short-lived phase of catastrophic mass loss (which KIC 12557548b has entered). Because this catastrophic stage lasts only up to a few percent of the planet's life, we estimate that for every object like KIC 12557548b, there should be 10--100 close-in quiescent progenitors with sub-day periods whose hard-surface transits may be detectable by Kepler --- if the progenitors are as large as their maximal, Mercury-like sizes (alternatively, the progenitors could be smaller and more numerous). According to our calculations, KIC 12557548b may have lost ~70% of its formation mass; today we may be observing its naked iron core.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 14 August 2018 (MN LATEX style file v2.2) Catastrophic Evaporation of Rocky Planets Daniel Perez-Becker1⋆ and Eugene Chiang2 1Department of Physics, University of California at Berkeley, 366 LeConte Hall, Berkeley CA 94720-7300, USA 2Departments of Astronomy and of Earth and Planetary Science, University of California at Berkeley, Hearst Field Annex B-20, Berkeley CA 94720-3411, USA Submitted: 14 August 2018 ABSTRACT Short-period exoplanets can have dayside surface temperatures surpassing 2000 K, hot enough to vaporize rock and drive a thermal wind. Small enough planets evaporate completely. We construct a radiative-hydrodynamic model of atmospheric escape from strongly irradiated, low-mass rocky planets, accounting for dust-gas energy exchange in the wind. Rocky planets with masses . 0.1M⊕ (less than twice the mass of Mercury) and surface temperatures & 2000 K are found to disintegrate entirely in . 10 Gyr. When our model is applied to Kepler planet candidate KIC 12557548b -- which is believed to be a rocky body evaporating at a rate of M & 0.1 M⊕/Gyr -- our model yields a present-day planet mass of . 0.02 M⊕ or less than about twice the mass of the Moon. Mass loss rates depend so strongly on planet mass that bodies can reside on close-in orbits for Gyrs with initial masses comparable to or less than that of Mercury, before entering a final short-lived phase of catastrophic mass loss (which KIC 12557548b has entered). Because this catastrophic stage lasts only up to a few percent of the planet's life, we estimate that for every object like KIC 12557548b, there should be 10 -- 100 close-in quiescent progenitors with sub-day periods whose hard-surface transits may be detectable by Kepler -- if the progenitors are as large as their maximal, Mercury-like sizes (alternatively, the progenitors could be smaller and more numerous). According to our calculations, KIC 12557548b may have lost ∼70% of its formation mass; today we may be observing its naked iron core. Key words: hydrodynamics -- planets and satellites: atmospheres -- planets and satellites: composition -- planets and satellites: physical evolution -- planet-star interaction 1 INTRODUCTION Atmospheric escape shapes the faces of planets. Mechanisms for mass loss vary across the Solar System. Planets lacking mag- netic fields can be stripped of their atmospheres by the Solar wind: the atmospheres of both Mercury and Mars are continu- ously eroded and refreshed by Solar wind ions (Potter & Morgan 1990, 1997; Lammer & Bauer 1997; Bida et al. 2000; Killen et al. 2004; Jakosky & Phillips 2001). Venus demonstrates the extreme sensitivity of atmospheres to stellar insolation. Although its dis- tance to the Sun is only 30% less than that of the Earth, Venus has lost its water to hydrodynamic escape powered by Solar ra- diation (Kasting & Pollack 1983; Zahnle & Kasting 1986). Im- pacts with comets and asteroids can also purge planets of their atmospheres, as has thought to have happened to some extent on Mars (Melosh & Vickery 1989; Jakosky & Phillips 2001) and ⋆ e-mail: [email protected]. c(cid:13) 0000 RAS some giant planet satellites (Zahnle et al. 1992). An introductory overview of atmospheric escape in the Solar System is given by Catling & Zahnle (2009). Smaller bodies lose their atmospheres more readily because of their lower surface gravities. Bodies closer to their host star also lose more mass because they are heated more strongly. Extrasolar planets on close-in orbits are expected to have highly processed at- mospheres. In extreme cases, stellar irradiation might even evapo- rate planets in their entirety -- comets certainly fall in this category. Hot Jupiters are the best studied case for atmospheric erosion in extrasolar planets see Yelle 2004, 2006; Tian et al. 2005; Garc´ıa Munoz 2007; Holmstrom et al. 2008; Murray-Clay et al. 2009; Ekenback et al. 2010; Tremblin & Chiang 2013). For these gas giants, mass loss is driven by stellar ultraviolet (UV) radiation which photo-ionizes hydrogen and imparts enough energy per proton to overcome the planet's large escape velocity (∼60 km/s). Mass loss for M ∼ 1010 -- 1011 g/s hot Jupiters typically occurs at a rate of theoretical models, (for 2 D. Perez-Becker & E. Chiang so that ∼1% of the planet mass is lost over its lifetime (Yelle 2006; Garc´ıa Munoz 2007; Murray-Clay et al. 2009). Although this erosion hardly affects a hot Jupiter's internal structure, the outflow is observable. Winds escaping from hot Jupiters have been spectroscopically detected by the Hubble Space Telescope (HST). During hot Jupiter transits, several absorp- tion lines (H i, O i, C ii, Mg ii, and Si iii) deepen by ∼2 -- 10% (Vidal-Madjar et al. 2003, 2004, 2008; Ben-Jaffel 2007, 2008; Fossati et al. 2010; Lecavelier Des Etangs et al. 2010; Linsky et al. 2010; Lecavelier des Etangs et al. 2012). Because of their lower escape velocities, lower mass super- Earths should have their atmospheres more significantly sculpted by evaporation. Lopez et al. (2012) found that hydrogen-dominated atmospheres of close-in super-Earths could be stripped entirely by UV photoevaporation (e.g., Kepler-11b). At even lower masses and stellocentric distances, planets might evaporate altogether. When dayside temperatures surpass ∼2000 K, rocks, particularly silicates and iron (and to a lesser ex- tent, Ti, Al, and Ca), begin to vaporize. If the planet mass is low enough, thermal velocities of the high-metallicity vapor could ex- ceed escape velocities. Under these circumstances, mass loss is not limited by stellar UV photons, but is powered by the incident bolo- metric flux. 1.1 KIC 12557548b: A Catastrophically Evaporating Planet In fact, the Kepler spacecraft may have enabled the discovery of an evaporating, close-in rocky planet near the end of its life. As re- ported by Rappaport et al. (2012), the K-star KIC 12557548 dims every 15.7 hours by a fractional amount f that varies erratically from a maximum of f ≈ 1.3% to a minimum of f . 0.2%. These occultations are thought to arise from dust embedded in the time- variable outflow emitted by an evaporating rocky planet -- here- after KIC 1255b. Because Kepler observes in the broadband op- tical, the obscuring material cannot be gas which absorbs only in narrow lines; it must take the form of dust which absorbs and scat- ters efficiently in the continuum. As we will estimate shortly, the amount of dust lost by the planet is prodigious. From the ∼2000 K surface of the planet is launched a ther- mal ("Parker-type") wind out of which dust condenses as the gas expands and cools. The composition of the high-metallicity wind reflects the composition of the evaporating rocky planet (see, e.g., Schaefer & Fegley 2009): the wind may consist of Mg, SiO, O, and O2 -- and Fe if it has evaporated down to its iron core. Stellar grav- ity, together with Coriolis and radiation-pressure forces, then shape the dusty outflow into a comet-like tail. The peculiar shape of the transit light curve of KIC 1255b supports the interpretation that the occultations are caused by a dusty tail. First, the light curve evinces a flux excess just before ingress, which could be caused by the for- ward scattering of starlight toward the observer. Second, the flux changes more gradually during egress than during ingress, as ex- pected for a trailing tail. Both features of the light curve and their explanations were elucidated by Rappaport et al. (2012), and fur- ther quantified by Brogi et al. (2012) who presented a phenomeno- logical light-scattering model of a dusty tail.1 1 Alternative ideas might include Roche lobe overflow from a compan- ion to the star (c.f. Lai et al. 2010). A problem with this scenario is that the mass transfer accretion stream, funneled through the L1 Lagrange point, Spectra of KIC 12557548 taken using the Keck Telescope re- veal no radial velocity variations down to a limiting accuracy of ∼100 m/s (A. Howard and G. Marcy 2012, personal communica- tion). A deep Keck image also does not reveal any background blended stars. These null results are consistent with the interpreta- tion that KIC 1255b is a small, catastrophically evaporating planet. The characteristic mass loss rate for KIC 1255b -- a quantity that we will reference throughout this paper -- can be estimated from the observations as follows. First compute the total mass in grains required to absorb/scatter a fraction f of the starlight, as- suming that the dust is optically thin (or marginally so). Out of the total area πR2 ∗ presented by the stellar face, take the dust cloud to cover an area A and to have a characteristic line-of-sight optical depth Σdκd, where Σd is the grain mass per unit area (on the sky) and κd is the grain opacity. Then A πR2 ∗ Σdκd . f = (1) Assume further that grains can be modeled as a monodispersion of spheres with radii s and internal material density ρint so that κd = 3 4ρints . The total mass in dust covering the star is then Md = ΣdA = 4π 3 f ρintR2 ∗ s . (2) (3) To obtain a mass loss rate, divide this mass by the orbital period of 15.7 hours on the grounds that transit depths change by up to an order of magnitude from orbit to orbit. For f = 1% -- when KIC 1255b is in a state of relatively high mass loss -- the resultant mass loss rate in grains is M1255b, dust ∼ 0.4 R∗ 0.65R⊙!2 f 0.01! s 0.5 µm! ρint 1 g/cm3! M⊕ Gyr−1 . (4) This is the mass loss rate in dust only, by construction. The gas- to-dust ratio by mass, η, must be at least on the order of unity if outflowing gas is to carry, by Epstein drag, embedded grains out of the gravity well of the planet (Rappaport et al. 2012). Henceforth we will adopt as our fiducial total mass loss rate (for the relatively high value of f = 1%): M1255b = (1 + η) M1255b, dust ∼ 1 M⊕ Gyr−1 . (5) would lead the companion in its orbit, not trail it as is required by obser- vation. By comparison, in our dusty wind model, there is no focussing of material at L1 because the wind is not in hydrostatic equilibrium and there- fore does not conform to Roche lobe equipotentials. Moreover, stellar ra- diation pressure, usually neglected in standard Roche-overflow models, is crucial for diverting the flow so that it trails the planet in its orbit (see also Mura et al. 2011). Another alternative involves bow shocks of planets pass- ing through stellar coronae (Vidotto et al. 2010, 2011a,b; Llama et al. 2011; Tremblin & Chiang 2013) which may explain the ingress-egress asymmetry observed for the exoplanet WASP-12b (Fossati et al. 2010). Bow shocks be- tween the planetary wind and the stellar wind are also expected to occur for KIC 1255b, but would be hard to detect through Kepler's broadband filter since the shocked gas alone would absorb only in narrow absorption lines. At the low gas densities characterizing exoplanet winds, only dust can gen- erate the large continuum opacities required by the deep transits observed by Kepler. Rappaport et al. (2012) reviews these and additional arguments supporting the dust tail model. A similar interpretation was proposed for the light curve of beta Pictoris (Lecavelier Des Etangs et al. 1997; Lamers et al. 1997). c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The total mass loss rate ∝ ρints(1 + η). Given the uncertainties in these three factors (some rough guesses: ρint & 0.5 g cm−3 for possi- bly porous grains; s & 0.1 µm; η & 1), the total mass loss rate could take any value M1255b > 0.1 M⊕ / Gyr (but see also our Figures 8 and 10 which place upper bounds on mass loss rates). Rappaport et al. (2012) claimed that the present-day mass of KIC 1255b is ∼0.1 M⊕. They estimated, to order-of-magnitude, the maximum planet mass that could reproduce the observed mass loss rate given by equation (5). Here we revisit the issue of planet mass -- and the evaporative lifetime it implies -- with a first-principles solution of the hydrodynamic equations for a planetary wind, in- cluding an explicit treatment of dust-gas thermodynamics. We will obtain not only improved estimates of the present-day mass, but also full dynamical histories of KIC 1255b. 1.2 Plan of This Paper We develop a radiative-hydrodynamic wind model to study mass loss of close-in rocky planets with dayside temperatures high enough to vaporize rock. Although our central application will be to understand KIC 1255b -- both today and in the past -- the model is general and can be easily be modified for parameters of other rocky planets. The paper is structured as follows. In §2 we present the model that computes M as a function of planet mass M. We give results in §3, including estimates for the maximum present-day mass and the maximum formation mass of KIC 1255b. A wide-ranging discus- sion of the implications of our results -- including, e.g., an explana- tion of why mass loss in this context does not obey the often-used energy-limited mass loss formula (cf. Lopez et al. 2012), as well as the prospects of detecting close-in Mercuries and stripped iron cores -- is given in §4. Our findings are summarized in §5. 2 THE MODEL We construct a 1D model for the thermally-driven atmospheric mass loss from a close-in rocky planet, assuming that it occurs in the form of a transonic Parker wind. The atmosphere of the planet consists of a high-metallicity gas created by the sublimation of sil- icates (and possibly iron, as we discuss in §4.7) from the planetary surface. We focus on the flow that streams radially toward the star from the substellar point on the planet. The substellar ray is of spe- cial interest because the mass flux it carries will be maximal insofar as the substellar point has the highest temperature -- and thus the highest rate of rock sublimation -- and insofar as stellar gravity will act most strongly to accelerate gas away from the planet. As the high-metallicity gas flows away from the surface of the planet, it will expand and cool, triggering the condensation of dust grains. Dust grains can affect the outflow in several ways: (a) latent heat released during the condensation of grains will heat the gas; (b) continuously heated by starlight, grains will transfer their energy to the gas via gas-grain collisions; and (c) absorption/scattering of starlight by grains will reduce the stellar flux reaching the planet, reducing the surface temperature and sublimation rate. In this work, we do not account for the microphysics of grain condensation, as cloud formation is difficult to calculate from first principles (see, e.g., the series of five papers beginning with Helling et al. 2001; see also Helling et al. 2008a and Helling et al. 2008b). Instead, we treat the dust-to-gas mass fraction xdust as a free function and explore c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 3 Figure 1. Equilibrium vapor pressures of pyroxene and olivine vs. temper- ature, obtained from equation (7). how our results depend on this function. In our 1-fluid model, dust grains are perfectly entrained in the gas flow, an assumption we test a posteriori (§4.2). We begin in §2.1 by evaluating thermodynamic variables at the surface of the planet -- the inner boundary of our model. In §2.2 we introduce the 1D equations of mass, momentum, and energy conservation. In §2.3 we explain the relaxation method employed to solve the conservation laws. 2.1 Base Conditions of the Atmosphere Dayside planet temperatures peak at approximately Tsurface ≈ T∗r R∗ a e−τsurface/4 ≈ 2150 e−τsurface/4 K, (6) where for our numerical evaluation we have used parameters ap- propriate to KIC 12557548 (orbital semimajor axis a ≈ 0.013 AU; effective stellar temperature T∗ ≈ 4450 K; and stellar ra- dius R∗ ≈ 0.65 R⊙; Rappaport et al. 2012). Here τsurface is the opti- cal depth between the planet's surface and the star due to absorp- tion/scattering by dust grains. We expect rocky extrasolar planets to consist of the same silicates predominantly found in the mantle of the Earth: pyrox- ene ([Fe,Mg]SiO3) and olivine ([Fe,Mg]2SiO4). At the estimated Tsurface, these silicates will vaporize and form an atmosphere whose equilibrium vapor pressure can be described by Pvapor = exp − mLsub kTsurface + b! , (7) where m is the mass of either a pyroxene or olivine molecule, k is Boltzmann's constant, Lsub is the latent heat of sublimation, and b is a constant. We follow Kimura et al. (2002) in our choice of param- eters for equation (7). For olivine, m = 169 mH (Mg1.1Fe0.9SiO4) where mH is the mass of atomic hydrogen, Lsub = 3.21 × 1010 4 D. Perez-Becker & E. Chiang erg g−1, and eb = 6.72 × 1014 dyne cm−2; these values were found experimentally by evaporating forsterite (Mg2SiO4), an end- member of olivine (Nagahara et al. 1994). For pyroxene, m = 60 mH, Lsub = 9.61 × 1010 erg g−1, and eb = 3.13 × 1011 dyne cm−2; these values were determined by Hashimoto (1990) for SiO2. Kimura et al. (2002) argue that the parameters for SiO2 are ap- propriate for pyroxene because evaporation experiments of en- statite (MgSiO3) show that SiO2 escapes preferentially, leaving the mineral with a polycrystalline forsterite coating (Tachibana et al. 2002). We plot equation (7) for both olivine and pyroxene in Fig- ure 1. According to Schaefer & Fegley (2009), the atmosphere of a hot rocky planet at temperatures similar to Tsurface is primarily com- posed of Mg, SiO, O, and O2. In our model we choose a mean molecular weight for the atmospheric gas of µ = 30 mH, (8) similar to the average molecular mass of Mg and SiO. Additionally, we set the adiabatic index of the gas to be γ ≡ cP/cV = 1.3, ap- propriate for diatomic gases at high temperatures (e.g., γT =2000K = 1.318, Lange & Forker 1967). The gas density at the surface of the planet is H2 ρvapor = µPvapor kTsurface . (9) 2.2 Equations Solved In our 1D hydrodynamic model, we solve the equations for mass, momentum, and energy conservation of the gas, assuming a steady flow. From mass conservation, ∂ ∂r (r2ρv) = 0, (10) where r is the radial distance from the center of the planet, and ρ and v are the density and velocity of the gas. In the frame rotating at the orbital angular frequency of the planet, momentum conser- vation -- with the Coriolis force omitted -- implies ρv ∂v ∂r = − ∂P ∂r − GMρ r2 + 3GM⋆ρr a3 , (11) where G is the gravitational constant, P = ρkT /µ is the gas pres- sure, T is the gas temperature, and M and M∗ = 0.7M⊙ are the masses of the planet and the host star (KIC 12557548), respectively. The last term on the right-hand side of eq. (11), which we refer to as the tidal gravity term, is the sum of the centrifugal force and the gravitational attraction from the star. In deriving eq. (11), we have neglected terms of order (r/a)2. We do not include the contribution of the Coriolis force because its magnitude will only be compara- ble to the gravitational attraction of the planet when the gas moves at bulk speeds comparable to the escape velocity of the planet at its Hill sphere (a.k.a. the Roche lobe). This generally occurs near the outer boundary of our calculation, which is set by the location of the sonic point. Although the sonic point radius and Hill sphere radius are distinct quantities, we will find in practice for our mod- els that they are close to one another, which is not surprising since the wind most easily accelerates to supersonic velocities where the effective gravity is weakest -- i.e., near the Hill sphere. The steady-state energy conservation equation, ∇ · (ρuv) = −P∇ · v + Γ, together with equation (10) and the specific internal energy u = kT /[(γ − 1)µ], yields ρv ∂ ∂r " kT (γ − 1)µ# = kT v µ ∂ρ ∂r + Γcol + Γlat. (12) The left hand side of equation (12) tracks changes in internal en- ergy, while the terms on the right hand side track cooling due to PdV work, heating from gas-grain collisions, and latent heat re- leased from grain condensation. We have omitted conduction be- cause we found it to be negligible compared to all other terms. For gas heating by dust-gas collisions, we assume that each collision between a gas molecule and a dust grain transfers an en- ergy k(Tdust − T ) to the gas molecule. Then the rate of gas heating per unit volume is Γcol = 3xdust 4sρint k µ (Tdust − T ) ρ2cs, (13) where xdust(r) ≡ ρdust/ρ is the spatially varying dust-to-gas mass ratio, s = 1 µm is the assumed grain size, ρint = 3 g cm−3 is the grain bulk density, and cs = s γkT µ (14) is the sound speed of the gas. As discussed in §4.2, our assumption that micron-sized grains are entrained in the flow is valid for the lowest mass planets we consider (a group that includes, as best we will gauge, KIC 1255b) but not the highest mass planets. We adopt Tdust = T∗r R∗ a e−τ/4 (15) for simplicity, so that Tdust equals Tsurface at the surface of the planet. The dust optical depth to starlight from r to the star is τ = 3 4sρint Z rs r ρxdustdr + τs. (16) The constant τs accounts for the optical depth from the outer bound- ary of our calculation -- the sonic point radius rs -- to the star. We arbitrarily set τs = 0.01. (17) Our results are insensitive to our assumed value of τs as long as it is ≪ 1. The optical depth at the planet's surface is τsurface ≡ τ(r = R), where R is the planetary radius, calculated for a given M by assuming a constant bulk density of 5.4 g/cm3, the mean density of Mercury. Latent heat from grain formation is released at a rate given by Γlat = Lsubρv dxdust dr , (18) where we have used Lsub appropriate for pyroxene, as it is more thermally stable than olivine (see Figure 1; in any case, Lsub for olivine is not that different). Because we do not model dust condensation from first princi- ples, the dust-to-gas mass ratio xdust(r) is a free function, which we parametrize as follows: log[xdust(r)] = log(xdust, max) − log(xdust, amp) ·  1 − arctan" 8(r − R) 3!#−1 3(rs − R)#"arctan 8 increases monotonically with r to a max- The function xdust imum value of xdust, max, starting from a minimum value of . (19) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 xdust, max/xdust, amp. We have two reasons for choosing xdust to in- crease rather than decrease with increasing height. The first is phys- ical: as the wind launches into space, it expands and cools, allowing the metal-rich vapor to more easily saturate and condense (see also §4.8). The second is motivated by observations: on length scales on the order of the stellar radius, far from the planet, the wind must have a relatively high dust content so as to produce the deep tran- sits observed by Kepler. Of course, neither of these reasons consti- tutes a first-principles proof that xdust actually does increase with r. Indeed the increasingly low density of the wind may frustrate con- densation. A physics-based calculation of how grains may (or may not) form in the wind is sorely needed (§4.8); our present study merely parameterizes this difficult problem. As dust condenses from gas and xdust rises, the gas density ρ must fall by mass conservation. For simplicity we neglect the dependence of ρ on xdust and so restrict our analysis to xdust, max ≪ 1; in practice, the maximum value of xdust, max that we consider is 10−0.5 ≈ 0.32. 2.3 Relaxation Code The structure of the wind is found by the simultaneous solution of equations (10), (11), (12), & (16) with the appropriate bound- ary conditions. This system constitutes a two-point boundary value problem, as the boundary conditions for ρ and T are defined at the base of the flow, while those for v and τ are defined at the sonic point. Although generally more complicated, relaxation codes are preferred over shooting methods when solving for the transonic Parker wind. This is because for each transonic solution there are an infinite number of "breeze solutions" where the bulk speed of the wind never reaches supersonic velocities. It is easier to begin with an approximate solution that already crosses the sonic point and to then refine this solution, than to exhaustively shoot in mul- tidimensional space for the sonic point. Murray-Clay et al. (2009) used a relaxation code to find the transonic wind from a hot Jupiter. Here we follow their methodology; we develop a relaxation code based on Section 17.3 of Press et al. (1992). 2.3.1 Finite Difference Equations We set up a grid of N = 100 logarithmically spaced radii that run from the base of the flow located at the surface of the planet (r = R) to the sonic point (r = rs). We replace our system of differential equations by a set of finite-difference relations. For ρ and T : E1, j ≡ ∆ jρ − ∆ jr = ∆ jρ + ρ 2 r + 1 v dv dr! ∆ jr = 0, dρ dr dT dr (20) (21) E2, j ≡ ∆ jT − ∆ jr = ∆ jT − (γ − 1)" T ρ dρ dr + µ kρv (Γcol + Γlat)# ∆ jr = 0, where finite differences are calculated upwind (∆ j x ≡ x j − x j−1), as ρ and T (and by extension E1 and E2) have boundary con- ditions defined at the surface of the planet ( j = 0). Following Press et al. (1992), we average across adjacent grid points when c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 5 evaluating variables in (20) and (21), e.g., ρ ≡ (ρ j−1 + ρ j)/2. The finite-difference relations for v and τ are: E3, j ≡ ∆ jv − = ∆ jv − dv dr ∆ jr v v2 − γkT /µ" 2γkT µr − γ − 1 ρv (Γcol + Γlat) −GM/r2 + 3GM∗r/a3#∆ jr = 0, E4, j ≡ ∆ jτ − = ∆ jτ + dτ ∆ jr dr 3 4sρint ρxdust∆ jr = 0. (22) (23) Note that differences are now computed downwind (∆ jx ≡ x j+1 − x j), as v and τ have boundary conditions defined at the sonic point ( j = N). When evaluating variables in (22) and (23), we average across adjacent gridpoints downwind, e.g., ρ ≡ (ρ j + ρ j+1)/2. 2.3.2 Boundary Conditions We need four boundary conditions to solve the system of finite dif- ference equations (20) -- (23). At the base of the atmosphere, we set the boundary conditions for gas density and temperature using re- lations from §2.1: E1,0 = ρ0 − ρvapor = 0 E2,0 = T0 − Tsurface = 0. (24) At our outer boundary -- the sonic point -- we require the bulk velocity of gas to equal the local sound speed, and τ to equal our specified value (§2.2): E3,N = vN − cs = 0 E4,N = τN − τs = 0. (25) At every step of the iteration we determine rs by demanding dv/dr to be finite at the sonic point. As the denominator of equation (22) vanishes at v = cs, we require that " 2γkT µr − γ − 1 ρv (Γcol + Γlat) − GM r2 + 3GM∗r a3 #rs = 0. (26) 2.3.3 Method of Solution The relaxation method determines the solution by starting with an appropriate guess and iteratively improving it. We use the multi- dimensional Newton's method as our iteration scheme, which re- quires us to evaluate partial derivatives of all Ei, j with respect to all 4N dependent variables (ρ j, T j, v j, τ j). We evaluate partial deriva- tives numerically, by introducing changes of order 10−8 in depen- dent variables and computing the appropriate finite differences. Newton's method produces a 4N × 4N matrix, which we invert us- ing the numpy library in python. At each iteration we numerically solve equation (26) for rs and re-map all N gridpoints between R and rs. We iterate until variables change by less than one part in 1010. We found our iteration scheme to be rather fragile. The code only converges if initial guesses are already close to the solution. 6 D. Perez-Becker & E. Chiang For this reason, we started with a simplified version of the prob- lem with a known analytic solution as the initial guess. We then gradually added the missing physics to the code, obtaining a solu- tion with each new physics input until the full problem was solved. Specifically, we began by solving the isothermal Parker wind prob- lem (e.g., Lamers & Cassinelli 1999). We enforced the isothermal condition by declaring γ to be unity, so that our energy equation read dT /dr = 0 (see eq. 21). Additionally, we set M∗ and xdust to zero to remove the effects of tidal gravity and dust. We then slowly increased each of the parameters M∗, γ, and xdust (in that order) to their nominal values. Any subsequent parameter change (e.g., M) was also performed in small increments. In summary, our code contains three main input parameters: xdust, amp and xdust, max which prescribe the dust-to-gas profile xdust(r), and the planet mass M. Our goal is to explore the dependence of the mass loss rate M = Ωρvr2 (27) on these three parameters. Here Ω is the solid angle over which the wind is launched, measured from the center of the planet; we set Ω = 1 since the high surface temperatures required to produce a wind are likely to be reached only near the substellar point. We independently vary the parameters xdust, max and xdust, amp to find the maximum M for a given M. 3 RESULTS We begin in §3.1 with an isothermal gas model. The isothermal model serves both as a limiting case and as a starting point for de- veloping the full solution which includes a realistic treatment of the energy equation. We provide results for our full model in §3.2. In the full solution we focus on three possible planet masses M = 0.01, 0.03, and 0.07 M⊕, finding that M ≈ 0.01 M⊕ yields a maxi- mum M that is compatible with the observationally inferred M1255b. We conclude that in the context of our energy equation, the present- day mass of KIC 1255b is at most ∼0.02 M⊕ (since smaller mass M that are also compatible with planets can generate still higher M1255b). In §3.3 we integrate back in time to compute the maximum formation mass of KIC 1255b. 3.1 Isothermal Solution We begin by solving a steady, dust-free, isothermal wind described by equations (10) and (11) with M∗ set equal to zero. Although this is a highly idealized problem, its solution provides us with a starting point (i.e., an initial model) from which we are able to solve more complicated problems (see §2.3.3). Our code accepts the standard isothermal wind solution (e.g., Chapter 3 of Lamers & Cassinelli 1999) as an initial guess, with minimal relaxation. Mass loss rates derived for the isothermal model (T =2145 K, the temperature given by eq. 6 with τsurface = τs = 0.01) are plot- ted in Figure 2. They depend strongly on planet mass, with large gains in M for comparatively small reductions in M, for the basic reason that atmospheric densities are exponentially sensitive to sur- face gravity. Tidal gravity boosts M by reducing the total effective gravity. As M decreases, gravity becomes increasingly irrelevant, the thermal speed of the gas eventually exceeds the surface escape M on M weakens. In the "free- velocity, and the dependence of streaming" limit (dotted lines in Figure 2), mass loss is no longer "free streaming" limit no tidal gravity with tidal gravity T = 2145 K 102 101 100 10−1 10−2 10−3 10−4 ] r y G / ⊕ M [ M olivine KIC 12557548b pyroxene 10−5 10−3 10−2 M/M⊕ 10−1 100 Figure 2. Evaporative mass loss rates M vs. planet mass M for isothermal dust-free winds. Solid lines are for models that include stellar tidal grav- ity, while dashed curves are for models that do not. As the planet mass is reduced, all solutions converge toward the free-streaming limit where M is not influenced by gravity but instead scales with the surface area of the planet (eq. 28). As inferred from observations, possible present-day mass loss rates M1255b for KIC 1255b are marked in gray. Technically we have only a lower limit on M1255b of 0.1M⊕ / Gyr; for purposes of discussion throughout this paper, we adopt 0.1 -- 1M⊕ / Gyr as our fiducial range (see discussion surrounding equation 5). Clearly KIC 1255b cannot be a pure pyroxene planet. Subsequent figures will refer to planets with pure olivine surfaces (except in §4.7 where we consider iron). influenced by gravity, and occurs at a rate Mfree ≈ ρvaporcsR2. (28) Mfree decreases when M is reduced, as less surface area Note that (R2 ∝ M2/3) is available for evaporation. For M ∼ 0.03M⊕, a pure olivine surface can produce a (dust-free, isothermal) wind for which M becomes compatible with the observationally inferred M1255b ∼ 1 M⊕/Gyr (this is value for KIC 1255b, on the order of the observed mass loss rate for both dust and gas combined; see our discussion surrounding equation 5 and Rappaport et al. 2012). By contrast, for a pure pyroxene surface, M ≪ M1255b always. Hence- forth we will calculate the density ρvapor of gas at the surface using parameters appropriate for olivine. 3.2 Full Solution We relax the isothermal condition in our code by slowly increasing γ from 1 to 1.3. With this parameter change, we are solving the full energy equation (eq. 12). Grains are gradually added by modifying xdust, amp and xdust, max, necessitating the calculation of optical depth (eq. 16). In Figure 3 we show the solution for which M was maximized (over the space of possible values of xdust) for M = 0.03 M⊕, or about half the mass of Mercury. Recall from §3.1 that when the wind was assumed to be isothermal, a planet of this mass was able c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 7 nearly balancing. Only when velocities are nearly sonic does advec- tion become a significant term in the momentum equation (bottom right panel of Figure 3). Flow speeds are still high enough at the base of the atmosphere to lift dust grains against gravity for all but the highest planet masses considered (§4.2; gas speeds as low as ∼1 m/s in a ∼1 µbar atmosphere can enable micron-sized grains to es- cape sub-Mercury-sized planets, as can be verified by equation 30; furthermore, the condition that grains not slip relative to gas is con- veniently independent of gas velocity in the Epstein free-molecular drag regime, as can be seen in equation 29. Our flows are entirely in the Epstein drag regime, as explained below equation 29.) In Figure 4, we show how M and τsurface vary with the function xdust. As just discussed, max M is reached for a specific choice of parameters which balance gas heating by dust and surface obscu- ration by dust. (We emphasize, here and elsewhere, that we have not identified a physical reason why actual systems like KIC 1255b should have mass loss rates M that equal their theoretically allowed maximum values.) Apart from the region near the peak in M, con- tours of constant τsurface are roughly parallel to contours of constant M, suggesting that the value of τsurface is more important for deter- mining M than the specific functional form of xdust. This finding increases the confidence we have in the robustness of our solution. It is clear from Figures 3 and 4 that a planet of mass M = 0.03 M⊕ can only emit winds for which max M < min M1255b ∼ 0.1M⊕ / Gyr (at least within the context of our energy equation). In Figure 5 we show results for M = 0.01 M⊕, for which max M > min M1255b. For this M, the mass loss rate M is maximized when the planet surface is essentially unobscured by dust. At our arbitrarily chosen value for xdust ∼ 3×10−7, the wind is not significantly heated by dust and expands practically adiabatically. The maximum M is reached for an essentially dust-free solution because the wind is blowing at too high a speed for dust-gas collisions to be important. That is, the timescale over which gas travels from the planet's surface to the sonic point is shorter than the time it takes a gas particle to collide with a dust grain: gas is thermally decoupled from dust. Were we to increase xdust to make dust-gas heating important, the flow would become optically thick and the wind would shut down. If the model shown in Figure 5 does represent KIC 1255b, then the dust grains that occult the star must condense outside the sonic point, beyond the Hill sphere. Figure 6 shows the solution for M = 0.07 M⊕. In this case M is maximized when enough dust is present to heat the gas signifi- cantly above the adiabat. At this comparatively large planet mass, initial wind speeds are low enough that a gas particle collides many times with dust over the gas travel time, so that dust-gas collisions heat the gas effectively near the surface of the planet. Further down- wind, near the sonic point, latent heating overtakes PdV cooling and actually increases the temperature of the gas with increasing altitude. The wind is much more sensitive to heating near the base of the flow than near the sonic point: if we omit latent heating from our model -- so that gas cools to ∼500 K at the sonic point -- then M is reduced by only a factor of two. In comparison, M drops by several orders of magnitude if dust-gas collisions are omitted. The dependence of M on xdust in both the "dust-free low- mass" and the "dusty high-mass" limits is illustrated in Figure 7. In the low-mass limit, M = max M when no dust is present. In this regime, gas moves too quickly for heating by dust to be significant; dust influences the flow only by attenuating starlight, and as long as τsurface ≪ 1, dust hardly affects M. By contrast, in the high-mass limit, M = max M for the dustiest flows we consider (i.e., the high- Figure 3. Radial dependence of wind properties in the full solution for a planet of mass M = 0.03M⊕. Parameters of the xdust function are chosen to maximize M. In the upper six panels, the full solution is shown by solid curves, while the dust-free isothermal solution is marked by dashed curves. Values for v have been normalized to the sound speed at T = 2145 K, and values for ρ have been normalized to the density ρ0,2145K = µPvapor/(kT ) evaluated also at T = 2145 K. The ×-symbol marks the location of the sonic point (the outer boundary of our solution), which occurs close to the Hill radius RHill (marked by a vertical line). The two lower panels show the contributions of the individual terms in the energy and momentum equa- tions (left and right panels, respectively). to reproduce the inferred mass loss rate of KIC 1255b. Now, with our treatment of the full energy equation and the inclusion of dust, max M plummets by a factor of 40. Relative to the isothermal so- lution, the reduction in mass loss rate in the full model is mainly caused by the gas temperature dropping from 2095 K at the planet surface to below 500 K at the sonic point, and the consequent reduc- tion in gas pressure. The temperature drops because gas expands in the wind and does PdV work. Gas heating by dust-gas collisions or grain formation could, in principle, offset some of the temper- ature reduction, but having too many grains also obscures the sur- face from incident starlight, decreasing Tsurface and ρvapor (the latter quantity depends exponentially on the former). The particular dust- to-gas profile xdust that is used in Figure 3 is such that the ability of dust to heat gas is balanced against the attenuation of stellar flux by dust, so that M is maximized (for this M). Note that the flow at the base begins with velocities of ∼10−3 the sound speed. At these subsonic velocities, the atmosphere is practically in hydrostatic equilibrium, with gas pressure and gravity c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 8 D. Perez-Becker & E. Chiang M = 0.03 M⊕ −1 −2 −3 −4 −5 ) x a m , t s u d x ( g o l −6 0 1 2 3 4 5 log(xdust, amp) 6 7 8 −1.2 −1.9 −2.6 −3.3 −4.0 −4.7 −5.4 −6.1 −6.8 −7.5 ) 1 − r y G ⊕ M / M ( g o l M = 0.03 M⊕ −1 −2 −3 −4 −5 ) x a m , t s u d x ( g o l −6 0 1 2 3 4 5 log(xdust, amp) 6 7 8 −0.1 −0.4 −0.7 −1.0 −1.3 −1.6 −1.9 ) e c a f r u s τ ( g o l M on dust abundance for a planet of mass M = 0.03M⊕. The dust abundance parameters xdust, max and xdust, amp are Figure 4. Left panel: Dependence of defined in equation (19); see also Figure 3. Both parameters are varied independently to find the maximum M, which occurs at log(xdust, amp, xdust, max) ∼ (1.7, −2.2). The full model corresponding to this maximum M is detailed in Figure 3. Each black dot corresponds to a full model solution. Colour contours for M are interpolated using a cubic polynomial. Right panel: Optical depth τsurface between the star and the planetary surface as a function of dust abundance for a planet of mass M = 0.03M⊕. Apart from the region near the peak in M, contours of constant τsurface are roughly parallel to contours of constant M, suggesting that the value of τsurface is more important for determining M than the specific functional form of xdust. M is very sensitive to dust est values of xdust, max). In this regime, abundance, with values spanning six orders of magnitude over the explored parameter space. In Figure 8 we show mass loss rates as a function of planet mass for the full model. We emphasize that these are maximum mass loss rates, found by varying xdust. At low planet masses, max M for the full model converges with M for the isothermal model because both approach the free-streaming limit, where grav- ity becomes irrelevant and M is set entirely by conditions at the surface of the planet (eq. 28). To reach M1255b > 0.1M⊕ / Gyr, the present-day mass of KIC 1255b must be . 0.02 M⊕, or less than about twice a lunar mass. We have verified a posteriori for the models shown in Figures 3, 5, and 6 that the sonic point is attained at an altitude where the collisional mean free path of gas molecules is smaller than rs, so that the hydrodynamic approximation embodied in equations (10) -- (12) is valid. In other words, the exobase lies outside the sonic point in these M = max M models. The margin of safety is largest for the lowest mass models. 3.3 Mass-Loss Histories We calculate mass-loss histories M(t) by time-integrating our solu- tion, shown in Figure 8, for max M(M). Before we integrate, how- ever, we introduce 0 < fduty < 1 to account for the duty cycle of the wind: we define fduty · max M as the time-averaged mass-loss rate. We estimate that fduty ∼ 0.5 based on the statistics of transit depths compiled by Brogi et al. (2012, see their figure 2). We time- integrate fduty · max M to obtain the M(t) curves shown in Figure 9. (This factor of 2 correction for the duty cycle should not obscure the fact that the actual mass loss rate, time-averaged or otherwise, could still be much lower than the theoretically allowed max M, a possibility we return to throughout this paper.) We highlight the case of a planet with a lifetime of tlife = 5 Gyr. Under our full (non-isothermal) model (right panel of Figure 9), such a planet has a mass at formation of 0.06 M⊕ and grad- ually erodes over several Gyr until it reaches M ∼ 0.03 M⊕ -- whereupon the remaining mass is lost catastrophically over a short time (just how short is estimated for the specific case of KIC 1255b below). Contrast this example with planets having initial masses & 0.07 M⊕ -- these have such low M that they barely lose any mass over Gyr timescales. For comparison, we also present mass-loss histories using a dust-free isothermal model with stellar tidal gravity (left panel of Figure 9). Compared to our non-isothermal solutions -- for which gas temperatures fall immediately as the wind lifts off the planet surface -- the isothermal solution corresponds to a flow which stays relatively pressurized and which therefore enjoys the largest M for a given M. The isothermal wind thus represents an endmem- ber case. However, the mass-loss histories under the isothermal ap- proximation do not differ qualitatively from those using the full energy equation. For isothermal winds, the dividing mass between planet survival and destruction within 10 Gyr is about 0.11 M⊕, only ∼40% larger than the value cited above for our non-isothermal solutions. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 9 Figure 5. Same as Figure 3, but for a planet of mass M = 0.01M⊕. For this M is maximized when essentially no dust is present (i.e., the planet mass, atmosphere is essentially transparent) as gas moves too quickly for dust-gas collisions to heat the gas. As such, all heating terms are negligible and the gas expands adiabatically. Figure 6. Same as Figure 3, but for a planet of mass M = 0.07M⊕. For this M is maximized for the dustiest flows considered in this paper. The mass, outflow is launched at such low velocities that collisional dust-gas heating is important. Near the sonic point, latent heating overtakes PdV cooling and the temperature of the gas rises. 3.4 Possible Mass Loss Histories for KIC 1255b Figures 8 and 9 can be used to sketch possible mass-loss histories for KIC 1255b. For our first scenario, we employ the full model that solves the full energy equation. Today, to satisfy the obser- M1255b > 0.1 M⊕/Gyr (see the derivation vational constraint that of our equation 5), the planet must have a present-day mass of at most M ≈ 0.02 M⊕, or roughly twice the mass of the Moon (Fig- ure 8). Such a low mass implies that currently KIC 1255b is in a "catastrophic evaporation" phase (vertical straight lines in Figure 9). Depending on the age of the planet (i.e., the age of the star: 1 -- 10 Gyr), KIC 1255b originally had a mass of at most 0.04 -- 0.07 M⊕: 2 -- 4 times its maximum current mass, or about the mass of Mercury. Starting from today, the time the planet has before it dis- integrates completely is on the order of tevap ∼ M/( fduty · max M), which is as long as 400 (40) Myr for M = 0.02 M⊕, fduty = 0.5, and max M = 0.1 (1) M⊕ / Gyr. If our energy equation were somehow in error and the wind better described as isothermal, then the numbers cited above would change somewhat. The present-day mass of KIC 1255b would be at most M ∼ 0.07M⊕; the maximum formation mass would be be- tween 0.08 -- 0.11 M⊕; and the evaporation timescale starting from c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 today could be as long as tevap ∼ 400 Myr if the present-day mass M ∼ 0.07M⊕. As we have stressed throughout, a critical assumption we have made is that the time-averaged mass loss rate is fduty · max M with fduty as high as 0.5. We have not identified a physical reason why the actual mass loss rate should be comparable to the theoretically allowed max M for a given M. If our assumption were in error and actual mass loss rates M ≪ max M, then the present-day mass of KIC 1255b, the corresponding evaporation time, and the formation mass of KIC 1255 would all decrease from the upper bounds cited above. The degeneracy of possible masses and evaporative histories outlined in this section could be broken with observational searches for the progenitors of KIC 1255b-like objects (§4.6). 4 DISCUSSION We discuss how mass loss is not "energy-limited" in §4.1; how gas entrains dust in §4.2; how the evaporative wind can be time- variable in §4.3; the effects of winds on orbital evolution in §4.4; the interaction of the planetary wind with the stellar wind in §4.5; the occurrence rate of quiescent progenitors of catastrophically 10 D. Perez-Becker & E. Chiang M = 0.006 M⊕ −2 −3 −4 −5 −6 ) x a m , t s u d x ( g o l −7 0 1 2 3 4 5 log(xdust, amp) 6 7 8 1.5 1.2 0.9 0.6 0.3 0.0 −0.3 −0.6 ) 1 − r y G ⊕ M / M ( g o l M = 0.1 M⊕ −1 −2 −3 −4 −5 ) x a m , t s u d x ( g o l −6 0 1 2 3 4 5 log(xdust, amp) 6 7 8 −4.0 −4.8 −5.6 −6.4 −7.2 −8.0 −8.8 −9.6 −10.4 −11.2 ) 1 − r y G ⊕ M / M ( g o l M is maximized when no dust is Figure 7. Left panel: Same as the left panel of Figure 4, but for a planet of mass M = 0.006M⊕. In this low-mass limit, present (i.e., when the atmosphere is transparent) because gas is moving too quickly for heating by dust to be significant. Right panel: Same as left panel, but for a planet of mass M = 0.1M⊕. In this high-mass limit, M is maximized for the dustiest flows we consider (i.e., the highest values of xdust,max) because flow speeds near the surface are slow enough for dust-gas energy exchange to be significant. evaporating planets in §4.6; the possibility that evaporating planets can be naked iron cores in §4.7; and the ability of dust to condense out of the wind in §4.8. and gas (1-fluid approximation). Grains are well entrained if their momentum stopping times tstop ∼ mgrainvrel ρvrelcs s2 ∼ sρint ρcs (29) 4.1 Mass Loss is Not Energy-Limited Atmospheric mass loss rates are commonly assumed to be "UV- energy-limited": it is assumed that a fixed, order-unity fraction ǫ of the incident UV flux FUV does PdV work and lifts material out of the planet's gravitational well. Under the energy-limited assump- tion, the mass loss rate is M ∼ ǫπR2FUV/(GM/R). The validity of this formula has been tested by Watson et al. (1981) for terrestrial atmospheres and by Murray-Clay et al. (2009) for hot Jupiters. The energy-limited formula does not apply at all to the evap- orating rocky planets considered in this paper. The stellar UV flux is not essential for low-mass planets because their escape veloci- ties are so small that energy deposition by photoionization is not necessary for driving a wind. Even optical photons -- essentially the stellar bolometric spectrum -- can vaporize close-in planets. Merely replacing FUV with Fbolometric in the energy-limited formula would still be misleading, however, because most of the incident energy is radiated away and does no mechanical work. More to the point, ǫ is not constant; as discussed in §3.2, the energetics of the flow changes qualitatively with planet mass. If we were to insist on using the energy-limited formula, we would find ǫ ∼ 10−8 for M ∼ 0.1M⊕, and ǫ ∼ 10−4 for M ∼ 0.01M⊕ (see Figure 8). are shorter than the grain advection time tadv ∼ R/v. In equation (29), mgrain is the mass of an individual grain and vrel is the relative gas-grain velocity. The aerodynamic drag force in the denominator is given by the Epstein law, which is appropriate for vrel . cs and grain sizes s smaller than the gas collisional mean free path (for our flows, λmfp ∼ µ/(ρσ) & 2 cm, where σ ∼ 3 × 10−15 cm2). We find using our full model (for M = max M) that tstop/tadv < 1 everywhere for M < 0.03 M⊕, confirming our assumption that micron-sized (and smaller) grains are well-entrained in the winds emanating from such low-mass planets. When M = 0.03 M⊕, tstop < tadv close to the planet's surface, but as gas nears the sonic point, tstop becomes comparable to tadv. For M > 0.03 M⊕, the 1- fluid approximation breaks down near the sonic point. For such large planet masses, future models should account for gas-grain relative motion -- in addition to other effects that become increas- ingly important near the sonic point / Hill sphere boundary (e.g., Coriolis forces and stellar radiation pressure). Note that the 1-fluid approximation should be valid for the present-day dynamics of KIC 1255b, since according to the full model its current mass is at most M ≈ 0.02 M⊕ < 0.03 M⊕. Although dust can slip relative to gas, dust can still be trans- ported outward and escape the planet if the aerodynamic drag force exceeds the force of gravity: 4.2 Dust-Gas Dynamics ρvrelcss2 & ρints3g, (30) In our model we have assumed that dust grains condensing within the wind are carried along without any relative motion between dust where g = G(M/r2 − 3M∗r/a3) is the total gravitational accelera- tion. Grains that manage to be lifted beyond the Hill sphere (which c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 ] 1 − r y G ⊕ M [ M x a m 103 102 101 100 10−1 10−2 10−3 10−4 10−5 isothermal (T = 2145K), with tidal gravity full model KIC 12557548b 10−2 10−1 M/M⊕ 100 Figure 8. Maximum mass loss rates M vs. planet mass M for the full model. At each ×-marked mass, max M was found by varying xdust as described, e.g., in Figure 4. The solid curve is a cubic spline interpolation. Mass loss rates for the full model are generally lower than for the isothermal model (dashed curve) because the dust-gas heating terms we have included in our full model turn out to be inefficient. At low planet masses, the isothermal and full models converge because both approach the free-streaming limit, where gravity becomes irrelevant and M is set entirely by conditions at the surface (i.e., surface area, equilibrium vapor pressure, and sound speed; eq. 28). According to the full model, the present-day mass of KIC 1255b is . 0.02 M⊕, or less than twice the mass of the Moon; for such masses, max M > min M1255b ∼ 0.1 M⊕ / Gyr. isothermal (T=2145K) full model 0.14 0.12 0.10 ⊕ M / M 0.08 0.06 0.04 0.02 0.00 0 2 4 6 t [Gyr] 8 0 2 4 6 t [Gyr] 8 10 Figure 9. Mass-loss histories M(t) obtained by time-integrating fduty · max M(M) with fduty = 0.5 for the isothermal and full models. We high- light the case of a planet with a 5 Gyr lifetime. For the full model (right panel), this corresponds to an initial mass of ∼0.06 M⊕, slightly larger than the mass of Mercury. Such a planet could have formed in situ and slowly eroded over several Gyr until reaching ∼0.03 M⊕, whereupon the planet evaporates completely in a few hundred Myr. By contrast, planets with for- mation masses & 0.07M⊕ survive (in the full model) for tens of Gyrs with- out significant mass loss. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 11 is situated close to the sonic point in all our models) are no longer bound to the planet. Outside the Hill sphere, dust decouples from gas and is swept into a comet-like tail by stellar radiation pressure and Coriolis forces. For M ≈ 0.03 M⊕ and our adopted grain size s = 1 µm, the forces in equation (30) are comparable near the sonic point when vrel ∼ cs (i.e., when grains are barely lifted by drag). For M > 0.03 M⊕, micron-sized and larger grains will not be dragged past the sonic point according to (30) -- thus our assumption that they do fails for such high-mass planets, and will need to be recti- fied in future models. Smaller, sub-micron-sized grains can, how- ever, be lifted outward. Moreover, sub-micron sized grains may heat the gas more effectively because they have greater geomet- ric surface area for dust-gas collisions (at fixed xdust) and because their temperatures exceed those of blackbodies (their efficiencies for emission at infrared wavelengths are much less than their ab- sorption efficiencies at optical wavelengths). The superior momen- tum and thermal coupling enjoyed by grains having sizes s ≪ 1 µm, plus their relative transparency at optical wavelengths -- which helps them avoid shadowing the planet surface from stellar radi- ation -- motivate their inclusion in the next generation of models. Our model could be improved still further by self-consistently accounting for how the gas density ρ must decrease as dust grains condense out of gas. Moreover, the restriction that xdust < 1 could be lifted. 4.3 Time Variability Occultations of KIC 12557548 vary in depth from a maxi- mum of 1.3% to a minimum of . 0.2% on orbital timescales. Rappaport et al. (2012) discussed qualitatively the possibility that such transit depth variations arise from a limit cycle that alter- nates between high- M and low- M phases. A high- M phase that produces a deep eclipse would also shadow the planet surface from starlight. The resultant cooling would lower the surface vapor pres- sure and lead to a low- M phase -- after which the atmosphere would clear, the surface would re-heat, and the cycle would begin anew. Such limit cycle behavior could be punctuated by random explosive events that release dust, similar to those observed on Io (Geissler 2003; see also the references in §4.2 of Rappaport et al. 2012). Order-of-magnitude variations in M arise from only small fractional changes in surface temperature because the vapor pres- sure of gas over rock depends exponentially on temperature. For example, from equations (6) and (7), we see that increasing the sur- face optical depth τsurface from 0.1 to 0.4 reduces the surface temper- ature Tsurface by 150 K and the base density ρvapor by a factor of 10. Such changes would cause M to drop by more than a factor of 10, because M scales super-linearly with ρvapor: a linear dependence re- sults simply because M scales with gas density, while an additional dependence arises because the wind speed increases with the gas pressure gradient, which in turn scales as ρvaporTsurface. To reproduce the orbit-to-orbit variations in transit depth ob- served for KIC 1255b, the dynamical time tdyn of the wind cannot be much longer than the planet's orbital period of Porb = 15.7 hr. The dynamical time is that required for dust to be advected from the planet surface to the end of the comet-like tail that occults the star: it is the minimum timescale over which the planet's transit signature "refreshes". If tdyn ≫ Porb, then we would expect transit 12 D. Perez-Becker & E. Chiang depths to be correlated from one orbit to the next -- in violation of the observations. Referring to the full model for a KIC 1255b-like mass of M ≈ 0.01M⊕ (Figure 5), we estimate that tdyn =Z rs R dr v +Z 0.1R⋆ rs dr v ∼ 13 hr + 0.1R∗ v2 ∼ 14 hr, (31) where we have split tdyn into two parts: the first integral is the time for dust to reach the sonic point (the outer boundary of our calcu- lation), while the second integral is the time for dust to travel out to 0.1R∗ (a circular, optically disk of this radius would generate a 1% transit depth). The first integral is performed numerically us- ing our full model (upper left panel of Figure 5), while the second integral is estimated to order-of-magnitude using a characteristic grain velocity (well outside the planet's Hill sphere) of v2 ∼ 10 km s−1 (Rappaport et al. 2012; see their equation 6 and related com- mentary). Since tdyn . Porb, we conclude that the wind/cometary tail can refresh itself quickly enough to change its appearance from orbit to orbit. We can go one step further. The timescale over which M changes should be the timescale over which the stellar insolation at the planetary surface changes -- in other words, the timescale over which the "weather" at the substellar point, where the wind is launched, changes from, e.g., "overcast" to "clear". Ignoring the possibility of volcanic eruptions, we estimate this variability timescale as the time for the wind to reach the Hill sphere bound- ary, at which point the Coriolis force has turned the wind by an order-unity angle away from the substellar ray joining the planet to the star (i.e., beyond the Hill sphere, the dust-laden wind no longer blocks stellar radiation from hitting the substellar region where the wind is launched). Because the Hill radius is situated near the sonic point, this variability timescale is given approximately by the first integral in equation (31): ta dyn ≈ 13 hr. The fact that ta dyn is neither much shorter than nor much longer than Porb supports our proposal that the observed time variability of KIC 1255b is driven by a kind of limit cycle involving stellar inso- lation and mass loss. Clearly if ta dyn ≫ Porb, orbit-to-orbit variations in M would be impossible. Conversely, if ta dyn ≪ Porb -- or more precisely if ta dyn were much shorter than the transit duration of 1.5 hr -- then the wind would vary so rapidly that each transit observa- tion would time-integrate over many cycles, yielding a smeared-out average transit depth that would not vary from orbit to orbit as is observed. 4.4 Long-Term Orbital Evolution When computing planet lifetimes in §3.3, we have assumed that the planet does not undergo any orbital evolution while losing mass. This approximation is valid because gas leaves the planet's surface at velocities vlaunch . cs ∼ 1 km s−1. Launch velocities are so low compared to the orbital velocity of the planet vorb ∼ 200 km s−1 that the total momentum imparted by the wind is a tiny fraction of the planet's orbital momentum. We quantify this as follows. We estimate the total change in semimajor axis a and eccentricity e from Gauss' equations (see, e.g., Murray & Dermott 1999). Consider the case in which the gas is launched at a velocity vlaunch at an angle α from the substellar ray and in the plane of the planet's orbit. The angle α could be non- zero because of surface inhomogeneities or asynchronous rotation of the planet. Neglecting terms of order e2 and αe, we find that a and e evolve according to ∼ M M M M ∼ a (1 + α), (e + α) , vlaunch vorb vlaunch vorb da dt de dt where the first term in parentheses on the right-hand-side of each equation is due to the radial component of the perturbation and the second term is due to the azimuthal component (where radius and azimuth are measured in a cylindrical coordinate system centered at the star and in the plane of the planet's orbit). Integrating equations (32) yields the changes in a and e accumulated over the planet's age: (32) ∆a a ∼ ln Minitial ∆e ∼ ln Mintial Mfinal ! vlaunch Mfinal! vlaunch vorb vorb (e + α) . 0.01 (e + α) , . 0.01. (33) When evaluating equation (33), we have used the maximum possi- ble launch velocity vlaunch ∼ cs (valid in the "free-streaming" limit) and Minitial/Mfinal ∼ 5 (see §3.3). The fact that evaporating planets undergo negligible orbital evolution suggests they could have resided on their current orbits for Gyrs -- indeed they might even have formed in situ. Swift et al. (2013) would disfavor in-situ formation of KIC 1225b because at the planet's orbital distance, dust grains readily sublimate, and os- tensibly there would have been no solid material in the primor- dial disk out of which rocky planets could have formed. How- ever, a less strict in-situ formation scenario is still viable. Solid bodies larger than dust grains obviously have longer evaporation times. Such large planetesimals could have drifted inward through the primordial gas disk and then assembled into the progenitors of objects like KIC 1255b at their current close-in distances (e.g., Youdin & Shu 2002; Hansen & Murray 2012; Chiang & Laughlin 2013). Solid particles can avoid vaporization if they merge faster than they evaporate. 4.5 Flow Confinement by Stellar Wind When computing the transonic solution for the planetary wind, we implicitly assumed that the outflow was expanding into a vacuum. In reality, the outflow from the planet will collide with the stellar wind and form a "bubble" around the planet.2 For our solution to be valid, the radius of the bubble -- i.e., the surface of pressure bal- ance between the winds -- has to be downstream of the sonic point where the flow is supersonic. This way, any pressure disturbance created at the interface cannot propagate upstream and influence our solution. At the wind-wind interface, the normal components of the pressures of the planetary and stellar winds balance. The total pres- sure includes the ram, thermal, and magnetic contributions. We compare the pressure of the stellar wind at a ∼ 0.01 AU ∼ 4R∗ 2 The considerations in this section parallel those for hot Jupiter winds and their bow shocks; see Tremblin & Chiang (2013), Vidotto et al. (2010, 2011a,b), and Llama et al. (2011). c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 with that of the planetary wind computed at rs to gauge whether the interface occurs downstream of rs. For the pressure P∗ of a main-sequence, solar-type star we are guided by the Solar wind. Flow speeds of the Solar wind are measured by tracking the tra- jectories of coronal features (Sheeley et al. 1997; Qu´emerais et al. 2007). At a = 0.01 AU the Solar wind3 is still accelerating with typ- ical flow speeds of v∗ ∼ 100 km s−1. For this v∗, and a canonical So- lar mass-loss rate of 2×10−14 M⊙ yr−1, the proton number density is n∗ = 2×105 cm−3. We take the proton temperature to be T∗ ∼ 106 K (Sheeley et al. 1997) and the heliospheric magnetic field strength to be B∗ ≈ 0.1 G at 0.01 AU (Kim et al. 2012). The total stellar pres- sure at 0.01 AU is then P∗ ∼ n∗mHv2 ∗/(8π) ∼ 4 × 10−4 dyne cm−2, with the magnetic pressure dominating other terms by an order of magnitude. ∗ + n∗kT∗ + B2 The ram and thermal pressures of the planet's wind are equal M = max M) Ps ∼ 2 × 10−3 at the sonic point and add up to (for (1 × 10−4) dyne cm−2 when M = 0.03 (0.07) M⊕. Thus for M . 0.03 M⊕ (a range that includes possible present-day masses of KIC 1255b), Ps > P∗ so that the planetary wind blows a bubble that extends beyond the sonic point and the transonic solutions we have computed are self-consistent. By contrast for M ∼ 0.07 M⊕, Ps ∼ P∗/4 < P∗ so that the stellar wind pressure will balance the planetary wind pressure inside of rs. The stellar wind will prevent the planetary wind from reaching supersonic velocities; the plan- etary outflow will conform to a "breeze" solution with a reduced M. If the colliding winds in our problem behave like those of hot Jupiters and their host stars, then Ps ∼ P∗/4 will reduce M by ∼30% (see figure 12 of Murray-Clay et al. 2009). We have found P∗ to be dominated by magnetic pressure. Since magnetic fields only interact with the ionized component of the planetary wind, and since the planetary wind may not be fully ionized, we might be over-estimating the effect of P∗. Dust grains in the planetary wind may absorb free charges so that the wind may be too weakly ionized to couple with the stellar magnetic field. If we ignore the magnetic contribution to the stellar wind, then Ps > P∗ for all planet masses that we have considered, and all our transonic solutions would be self-consistent. What about the planet's magnetic field? A sufficiently ion- ized outflow could be confined by the planet's own magnetic field. To assess the plausibility of this scenario, we use Mercury's field as a guide for KIC 1255b. The flyby of Mariner 10 and recent measurements by MESSENGER show that Mercury possesses a weak magnetic field with a surface strength of ∼ 3 × 10−3 Gauss (Ness et al. 1974; Anderson et al. 2008). The associated magnetic pressure ∼ 4 × 10−7(R/r)6 dyne cm−2 is small compared to both the thermal pressures that we have computed at the planet surface and the hydrodynamic pressures at the sonic point. For planetary mag- netic fields to confine the wind, they are required to have surface strengths in excess of ∼ 30 Gauss. 3 These stellar wind parameters are appropriate for the "slow" Solar wind blowing primarily near the equatorial plane of the Sun. In addition, the Solar wind also contains a "fast" component which emerges from coronal holes. During Solar minimum, the fast Solar wind is confined mainly to large heliographic latitudes, but may reach the equator plane during periods of increased solar activity (Kohl et al. 1998; McComas et al. 2003). Because most planets are expected to orbit near their stellar equatorial planes, we take the slow component of the Solar wind to be a better guide. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 13 4.6 Occurrence Rates of Close-In Progenitors According to our analysis in §3.4, currently KIC 1255b has a mass of at most 0.02 -- 0.07 M⊕, and is in a final, short-lived, catastroph- ically evaporating phase possibly lasting another tevap ∼ 40 -- 400 Myr during which dust is ejected at a large enough rate to produce eclipse depths of order 1%. But for most of its presumably Gyrs- long life, KIC 1255b was a more quiescent planet -- larger but still less than roughly Mercury in size (§3.4), with a stronger gravity and emitting a much more tenuous wind. We therefore expect that for every KIC 1255b-like object discovered, there are many more progenitors in the quiescent phase -- i.e., planets having sizes up to about that of Mercury orbiting main sequence K-type stars at ∼0.01 AU. Out of the ∼45,000 main sequence K-type stars in the Kepler Input Catalogue (Batalha et al. 2010), there is apparently only 1 object like KIC 1255b. Then fobserved ∼ fprogenitor ftransit fevap ∼ 1 45000 , (34) where fprogenitor is the intrinsic occurrence rate of progenitors around K stars with orbital periods shorter than a day, ftransit ∼ R∗/a ∼ 1/4 is the geometric probability of transit, and fevap = tevap/tlife ∼ {40, 400} Myr / 5 Gyr ∼ {0.8, 8}% is the fraction of the planet's lifetime spent in the catastrophic mass-loss stage. Inverting equa- tion (34), we estimate that fprogenitor ∼ {1, 0.1}% of K stars harbor a close-in planet having less than the mass of Mercury. For each KIC 1255b-like object there should be f −1 evap = tlife/tevap ∼ {130, 13} planets with radii no larger than about Mer- cury's (. 0.4 R⊕), transiting K-type stars with sub-day periods. These planets have transit depths of (R/R∗)2 . 10−5 -- small but possibly detectable by Kepler if light curves are folded over enough periods (S. Rappaport 2012, personal communication) and if M ∼ max M so that the progenitor masses attain their maximum, M were actually ≪ max M, the progen- Mercury-like values. If itor sizes would be less than that of Mercury. The smaller sizes would decrease their lifetimes tevap and thus increase their expected number f −1 evap, but would also render them undetectable even with Kepler. 4.7 Iron Planets In our model, we found the (maximum) present-day mass of KIC 1255b to be about 1/3 of its (maximum) mass at formation. If KIC 1255b began its life similar in composition to the rocky planets in the Solar System, evaporation may have stripped the planet of its silicate mantle, so that only its iron core remains: KIC 1255b could be an evaporating iron planet today. We estimate mass loss rates of a pure iron planet using our isothermal model, including tidal gravity. We set the mean molec- ular weight of the gas to µFe ≈ 56mH and use a bulk density for the planet of 8.0 g cm−3, appropriate for a pure iron planet of mass 0.01M⊕ (Fortney et al. 2007). We fit laboratory measurements of the iron vapor pressure4 at T = 2200 K (Desai 1986) to equation (7), obtaining eb = 7.8 × 1011 dyne cm−2 for a latent heat of subli- mation Lsub = 6.3 × 1010 erg g−1 (Desai 1986) and m = µFe. In Figure 10 we compare mass loss rates derived for the 4 We have found measurements of the iron vapor pressure at T ∼ 2000 K to differ by a factor of two in the literature (see, e.g., Nuth et al. 2003). 14 D. Perez-Becker & E. Chiang 103 102 101 100 10−1 ] 1 − r y G ⊕ M [ M 10−2 10−3 10−4 10−5 10−3 iron, isothermal olivine, isothermal T = 2145 K KIC 12557548b 0.07 0.06 0.05 ⊕ M / p M 0.04 0.03 0.02 0.01 0.00 0 iron, isothermal (T = 2145 K) 2 6 4 t [Gyr] 8 10 10−2 M/M⊕ 10−1 100 Figure 10. Mass loss rates M, computed using the isothermal model at T = 2145 K, for iron planets and olivine planets of varying M. In the low-M, free-streaming limit, the higher vapor pressure of iron allows for a higher M than for olivine. At larger M, mass loss rates are lower for iron planets than for olivine planets because of the higher molecular weight of iron. Our estimate for the maximum present-day mass of KIC 1255b varies by a factor of two between the iron and olivine scenarios. isothermal model with T = 2145 K for both an iron planet and a pure olivine planet. On the one hand, at this temperature, the vapor pressure for iron Pvapor, Fe = 1.8 × 103 dyne cm−2 is about 50 times higher than for olivine, which raises the mass loss rate by a similar factor in the free-streaming limit appropriate for low masses (see equation 28). On the other hand, at higher masses, M for iron drops below that for olivine because the iron atmosphere, with its higher molecular weight, is harder to blow off. These two effects counter- act each other so that an iron planet and a silicate planet both reach M1255b ∼ 1 M⊕ / Gyr at a similar M. Thus at least within the con- text of isothermal winds, our estimate for the present-day mass of KIC 1255b is insensitive to whether the evaporating surface of the planet is composed of iron or silicates. Figure 11 displays mass loss histories for an iron planet using our isothermal model at T = 2145 K. Initial planet masses are up to a factor of two lower than those of our full model using silicates. According to Figure 11, iron planets with M & 0.05M⊕ survive for tens of Gyrs. Compare this result to its counterpart in Figure 9 (left panel), which shows that olivine planets with M . 0.1M⊕ evaporate within ∼10 Gyr. This comparison suggests that for planets with the right proportion of silicates in the mantle to iron in the core, mantles may be completely vaporized, leaving behind essentially non-evaporating iron cores. Such massive, qui- escent iron cores might be detectable by Kepler via direct transits (see §4.6). Figure 11. Mass-loss histories M(t) for an iron planet, obtained by time- integrating fduty · max M with fduty = 0.5 for an isothermal wind. An iron planet with a 5 Gyr lifetime will have an initial mass of ∼0.044 M⊕. As was the case for olivine planets (see Figure 9), the catastrophic evaporation stage lasts only for ∼100 Myr. Iron planets (or planetary iron cores) with masses greater than 0.05 M⊕ survive for over 10 Gyr. 4.8 Dust Condensation Recall that we have not modelled the microphysics of dust forma- tion, but treated the dust-to-gas ratio as a free function. Are the con- ditions of our flow actually favorable for the condensation of dust? Any gas whose partial pressure is greater than its vapor pressure (at that T ) might condense and form droplets (modulo the many complications discussed below). Condensation might proceed until all vapor in excess of saturation is in cloud particles. If the conden- sates have sedimentation velocities larger than updraft speeds they will rain out of the atmosphere; otherwise they remain aloft (see, e.g., Lewis 1969; Marley et al. 1999; Ackerman & Marley 2001; Helling et al. 2001). In Figure 12, we compare gas pressures of our solutions (for M = max M) at M = 0.01, 0.03, and 0.07 M⊕ with the equilibrium vapor pressures of olivine and pyroxene. At the base of the flow, the gas pressure equals the saturation vapor pressure of olivine by construction. As gas expands and cools, its pressure generally re- mains above the saturation pressure of pyroxene, so conditions are favorable for the condensation of pyroxene grains. Unfortunately we have an embarrassment of riches: at T . 1700 K, our gas pres- sures are many orders of magnitude higher than silicate vapor pres- sures, and the concern is that gas condensation will lower pres- sures precipitously to the point where winds shut down. This fate could be avoided if grains take too long to condense out of the outflowing and rapidly rarefying gas -- i.e., the finite timescale of grain condensation might permit gas to remain supersaturated. Furthermore, even before bulk condensation can begin, seed parti- cles must first nucleate from the gas phase. These issues call for a time-dependent analysis of grain formation, perhaps along the lines made for brown dwarf and exoplanet atmospheres (Helling et al. 2001, 2008a,b; Helling & Rietmeijer 2009). Accounting for the ex- tra heating from small grains with sizes ≪ 1 µm as they first con- dense out of the wind (§4.2) would keep the gas closer to isothermal and help to prevent catastrophic condensation. Of course, by the wind leaves the sonic point and is diverted into the trailing comet-like tail, it must be full of grains to explain c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 ] 2 − m c e n y d [ r o p a v P 102 101 100 10−1 10−2 10−3 10−4 10−5 10−6 10−7 10−8 10−9 10−10 10−11 10−12 1 M ⊕ 3 M ⊕ 0 . 0 0 . 0 7 M ⊕ 0 . 0 olivine pyroxene 500 1000 1500 T [K] 2000 2500 Figure 12. Trajectories of the wind in pressure-temperature space, com- puted using the full model for M = 0.01, 0.03, and 0.07M⊕. Gas pressures generally remain above the equilibrium vapor pressures of pyroxene, per- mitting the condensation of pyroxene grains. the observed occultations of KIC 1255b. In this paper, we have not modeled the dusty tail at all. Future work should address the dy- namics of the tail and compute its extinction profile, with the aim of reproducing the observed transit lightcurves. 5 CONCLUSIONS Our work supports the hypothesis that the observed occultations of the K-star KIC 12557548 originate from a dusty wind emitted by an evaporating planetary companion ("KIC 1255b"). We reach the following conclusions. (i) Maximum present-day mass Our estimates for the maximum present-day mass of KIC 1255b range from max M ≈ 0.02M⊕ (roughly twice the mass of the Moon) to max M ≈ 0.07M⊕ (slightly larger than the mass of Mer- cury), depending on how strongly the wind is heated by dust and can remain isothermal (Figure 8). For these maximum masses, cal- culated mass loss rates peak at max M ∼ 0.1 M⊕/Gyr, which is just large enough to produce the observed transit depths of order 1%. Smaller planets with weaker gravities yield larger mass loss rates and are also compatible with the observations. (ii) Maximum formation mass For an assumed planet age of ∼5 Gyr, the maximum mass of KIC 1255b at formation ranges from 0.06 M⊕ to 0.1 M⊕, again depend- ing on the energetics of the wind (Figure 9). (iii) Mass threshold for catastrophic evaporation A pure rock planet of mass & 0.1M⊕ and surface temperature . 2200 K will survive with negligible mass loss for tens of Gyrs. (iv) Time variability The observed occultations of KIC 12557548 vary by up to an order of magnitude in depth without any apparent correlation between orbits. The implied order-of-magnitude variations in M can be ex- M to condi- plained in principle by the exponential sensitivity of c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Catastrophic Evaporation of Rocky Planets 15 tions at the planet surface. The dynamical "refresh" timescale of the wind tdyn ∼ 14 h is similar to the orbital period Porb = 15.7 h. This supports our dusty wind model because were tdyn ≫ Porb or tdyn ≪ Porb, then eclipse depths would correlate from orbit to orbit, in violation of the observations. (v) Progenitors and occurrence rates KIC 1255b's current catastrophic mass loss phase may represent only the final few percent of the planet's life. As such, for every KIC 1255b-like object, there could be anywhere from 10 to 100 larger planets in earlier stages of mass loss. These close-in, rela- tively quiescent progenitors may be detectable by Kepler through conventional "hard-sphere" transits if they are as large as Mercury. We cannot, however, rule out the possibility that the progenitors are lunar-sized or even smaller. If KIC 1255b remains the only catas- trophically evaporating planet in the Kepler database, then the oc- currence rate of close-in progenitors orbiting K-stars with sub-day periods is > 0.1%, with larger occurrence rates for progenitors in- creasingly smaller than Mercury. (vi) Iron planet KIC 1255b may have lost ∼70% of its formation mass to its thermal wind. It seems possible that today only the iron core of KIC 1255b remains and is evaporating. Transmission spectra of the occulting dust cloud might reveal whether the planet's surface is composed primarily of iron or silicates. (vii) Future modelling and observations Keeping the wind hot as it lifts off the planet surface significantly enhances mass loss rates. In our model we have included heating of the gas by micron-sized grains embedded in the flow. But these grains can also shadow the surface from starlight and reduce M if they are too abundant. Additional heating may be provided by super-blackbody grains with sizes ≪ 1 µm which do not signif- icantly attenuate starlight. Future models should incorporate such tiny condensates -- and treat the dusty comet-like tail that our pa- per has completely ignored. More information about grain size dis- tributions and compositions might be revealed by Hubble observa- tions of KIC 12557548 scheduled for early 2013. 6 ACKNOWLEDGMENTS We thank Saul Rappaport for alerting us to the feasibility of de- tecting Mercury-sized progenitors of KIC 1255b using Kepler, and Andrew Howard and Geoff Marcy for sharing their Keck observa- tions of KIC 12557548. We are also grateful to Andrew Ackerman, Tom Barclay, Bill Borucki, Raymond Jeanloz, Hiroshi Kimura, Ed- win Kite, Michael Manga, Mark Marley, and Subu Mohanty for useful and encouraging conversations; to Paul Kalas for pointing us to work on beta Pictoris that presaged the story of KIC 1255b; and to an anonymous referee for a helpful report that motivated us to review the various modeling uncertainties and introduced us to the cloud formation literature. D.P.-B. acknowledges the support of a UC MEXUS-CONACyT Fellowship. E.C. acknowledges Hubble Space Telescope grant HST-AR-12823.01-A. REFERENCES Ackerman, A. S. & Marley, M. S. 2001, ApJ, 556, 872 Anderson, B. J., Acuna, M. H., Korth, H., et al. 2008, Science, 321, 82 16 D. Perez-Becker & E. Chiang Batalha, N. M., Borucki, W. J., Koch, D. G., et al. 2010, ApJ Let- ters, 713, L109 Ben-Jaffel, L. 2007, ApJ Letters, 671, L61 Ben-Jaffel, L. 2008, ApJ, 688, 1352 Bida, T. A., Killen, R. M., & Morgan, T. H. 2000, Nature, 404, 159 Brogi, M., Keller, C. U., de Juan Ovelar, M., et al. 2012, A&A, 545, L5 Catling, D. C. & Zahnle, K. J. 2009, Scientific American, 300, 050000 Chiang, E. & Laughlin, G. 2013, MNRAS Desai, P. D. 1986, Journal of Physical and Chemical Reference Data, 15, 967 Ekenback, A., Holmstrom, M., Wurz, P., et al. 2010, ApJ, 709, 670 Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661 Fossati, L., Haswell, C. A., Froning, C. S., et al. 2010, ApJ Letters, 714, L222 Garc´ıa Munoz, A. 2007, Planetary & Space Science, 55, 1426 Geissler, P. E. 2003, Annual Review of Earth and Planetary Sci- ences, 31, 175 Hansen, B. M. S. & Murray, N. 2012, ApJ, 751, 158 Hashimoto, A. 1990, Nature, 347, 53 Helling, C., Ackerman, A., Allard, F., et al. 2008a, MNRAS, 391, 1854 Helling, C., Oevermann, M., Luttke, M. J. H., Klein, R., & Sedl- mayr, E. 2001, A&A, 376, 194 Helling, C. & Rietmeijer, F. J. M. 2009, International Journal of Astrobiology, 8, 3 Helling, C., Woitke, P., & Thi, W.-F. 2008b, A&A, 485, 547 Holmstrom, M., Ekenback, A., Selsis, F., et al. 2008, Nature, 451, 970 Jakosky, B. M. & Phillips, R. J. 2001, Nature, 412, 237 Kasting, J. F. & Pollack, J. B. 1983, Icarus, 53, 479 Killen, R. M., Sarantos, M., Potter, A. E., & Reiff, P. 2004, Icarus, 171, 1 Kim, R.-S., Gopalswamy, N., Moon, Y.-J., Cho, K.-S., & Yashiro, S. 2012, ApJ, 746, 118 Kimura, H., Mann, I., Biesecker, D. A., & Jessberger, E. K. 2002, Icarus, 159, 529 Kohl, J. L., Noci, G., Antonucci, E., et al. 1998, ApJ Letters, 501, L127 Lai, D., Helling, C., & van den Heuvel, E. P. J. 2010, ApJ, 721, 923 Lamers, H. J. G. L. M. & Cassinelli, J. P. 1999, Introduction to Stellar Winds Lamers, H. J. G. L. M., Lecavelier Des Etangs, A., & Vidal- Madjar, A. 1997, A&A, 328, 321 Lammer, H. & Bauer, S. J. 1997, Planetary & Space Science, 45, 73 Lange, N. A. & Forker, G. M. 1967, Handbook of Chemistry Lecavelier des Etangs, A., Bourrier, V., Wheatley, P. J., et al. 2012, A&A, 543, L4 Lecavelier Des Etangs, A., Ehrenreich, D., Vidal-Madjar, A., et al. 2010, A&A, 514, A72 Lecavelier Des Etangs, A., Vidal-Madjar, A., Burki, G., et al. 1997, A&A, 328, 311 Lewis, J. S. 1969, Icarus, 10, 365 Linsky, J. L., Yang, H., France, K., et al. 2010, ApJ, 717, 1291 Llama, J., Wood, K., Jardine, M., et al. 2011, MNRAS, 416, L41 Lopez, E. D., Fortney, J. J., & Miller, N. 2012, ApJ, 761, 59 Marley, M. S., Gelino, C., Stephens, D., Lunine, J. I., & Freed- man, R. 1999, ApJ, 513, 879 McComas, D. J., Elliott, H. A., Schwadron, N. A., et al. 2003, Geophys. Res. Lett., 30, 1517 Melosh, H. J. & Vickery, A. M. 1989, Nature, 338, 487 Mura, A., Wurz, P., Schneider, J., et al. 2011, Icarus, 211, 1 Murray, C. D. & Dermott, S. F. 1999, Solar system dynamics Murray-Clay, R. A., Chiang, E. I., & Murray, N. 2009, ApJ, 693, 23 Nagahara, H., Mysen, B. O., & Kushiro, I. 1994, Geochimica et Cosmochimica Acta, 58, 1951 Ness, N. F., Behannon, K. W., Lepping, R. P., Whang, Y. C., & Schatten, K. H. 1974, Science, 185, 151 Nuth, III, J. A., Ferguson, F. T., Johnson, N., & Martinez, D. 2003, in Lunar and Planetary Inst. Technical Report, Vol. 34, Lunar and Planetary Institute Science Conference Abstracts, ed. S. Mack- well & E. Stansbery, 1598 Potter, A. E. & Morgan, T. H. 1990, Science, 248, 835 Potter, A. E. & Morgan, T. H. 1997, Planetary & Space Science, 45, 95 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, Numerical recipes in C. The art of scientific computing Qu´emerais, E., Lallement, R., Koutroumpa, D., & Lamy, P. 2007, ApJ, 667, 1229 Rappaport, S., Levine, A., Chiang, E., et al. 2012, ApJ, 752, 1 Schaefer, L. & Fegley, B. 2009, ApJ Letters, 703, L113 Sheeley, Jr., N. R., Wang, Y.-M., Hawley, S. H., et al. 1997, ApJ, 484, 472 Swift, J. J., Johnson, J. A., Morton, T. D., et al. 2013, ApJ, 764, 105 Tachibana, S., Tsuchiyama, A., & Nagahara, H. 2002, Geochim- ica et Cosmochimica Acta, 66, 713 Tian, F., Toon, O. B., Pavlov, A. A., & De Sterck, H. 2005, ApJ, 621, 1049 Tremblin, P. & Chiang, E. 2013, MNRAS, 428, 2565 Vidal-Madjar, A., D´esert, J.-M., Lecavelier des Etangs, A., et al. 2004, ApJ Letters, 604, L69 Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., et al. 2003, Nature, 422, 143 Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., et al. 2008, ApJ Letters, 676, L57 Vidotto, A. A., Jardine, M., & Helling, C. 2010, ApJ Letters, 722, L168 Vidotto, A. A., Jardine, M., & Helling, C. 2011a, MNRAS, 411, L46 Vidotto, A. A., Jardine, M., & Helling, C. 2011b, MNRAS, 414, 1573 Watson, A. J., Donahue, T. M., & Walker, J. C. G. 1981, Icarus, 48, 150 Yelle, R. V. 2004, Icarus, 170, 167 Yelle, R. V. 2006, Icarus, 183, 508 Youdin, A. N. & Shu, F. H. 2002, ApJ, 580, 494 Zahnle, K., Pollack, J. B., Grinspoon, D., & Dones, L. 1992, Icarus, 95, 1 Zahnle, K. J. & Kasting, J. F. 1986, Icarus, 68, 462 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
1012.1516
1
1012
2010-12-07T14:24:08
Preface: Planetary Systems Beyond the Main Sequence 2010
[ "astro-ph.EP", "astro-ph.SR" ]
Preface of Planetary Systems Beyond the Main Sequence including conference scope and summary, short overview of programme, acknowledgements of patronage, sponsors, the scientific organising committee, and the local organising committee.
astro-ph.EP
astro-ph
PREFACE Planetary and brown dwarf companions to evolved stars have only recently been dis- covered. The aim of this conference was to discuss observational results and techniques for their detection, explore theoretical predictions for the formation and the fate of sub- stellar objects orbiting evolving stars, and assess their potential impact on the evolution of the host stars. It was central to the conference to explore the importance of new and upcoming space missions like Kepler, GAIA and PLATO for this emerging field. The meeting, held at the Harmoniesäle of the city of Bamberg, Germany from Au- gust 11 -- 14, 2010, was organised by the Dr. Remeis-Sternwarte and the University of Erlangen-Nürnberg. The meeting attracted 68 delegates from 16 nations and 4 conti- nents. The first planets outside the solar system were discovered around a pulsar. Therefore the programme was opened by a review on the present state of research in the field of pulsar planets. The next session focussed on the interaction of planets with their host stars during their evolution from a theoretical point of view, and included aspects of binary stellar evolution as well as second and third generation planet formation. In the third session the results from surveys of planets around giant/massive stars were presented. The inventory of exoplanets around G-K giants as well as a very hot planet transiting a rapidly rotating A star were highlighted. Stellar oscillations in the planet hosting giant stars were discussed as a major obstacle for the discovery of planets. The conference programme retraced the sequence of stellar evolutionary stages where the horizontal branch (HB) follows after the red-giant stage. The timing method to measure light travel time variations in pulsating stars and eclipsing binaries has been very successful in finding substellar companions to HB stars. Several talks reported on such discoveries, including the detection of a planetary companion around a metal-poor star with extragalactic origin. The formation and existence of planetary systems around white dwarf stars were the subjects of the last two sessions. The first part was devoted to debris and gaseous disks around white dwarfs, while the second part focussed on planets and brown dwarfs orbiting white dwarfs, which included evidence for two planets orbiting the post-CE binary NN Ser. The scientific programme ended with a lively round table discussion and concluded with the decision to hold a second meeting on this subject in 2012. The meeting finished with a reception at the Dr. Remeis-Sternwarte. The philosophy of the meeting was to provide an equal platform for all delegates, and especially to foster the full participation of young scientists. We are confident that we have achieved this goal. Invited reviews were scheduled at the beginning of the sessions to set the stage for the contributed talks. Ample time for discussion was provided following each talk, and extra discussion sessions concluded the sections individually as well as the entire meeting. We thank the Lord Mayor of Bamberg for his patronage and for the reception of the participants in the beautiful surroundings of the "Geyerswörth" castle by the Mayor Werner Hipelius. We are extremely grateful to the Deutsche Forschungsgemeinschaft, the University of Erlangen-Nürnberg and the Erlangen Centre for Astroparticle Physics for the gener- ous financial support, which made possible the participation of young researchers and facilitated the attendance of several of the key note speakers. Last but not least we thank the Scientific Organising Committee for their valuable help with the scientific planning and scheduling of the meeting. Scientific Organising Committee: Noam Soker & Uli Heber (co-chairs), Matt Burleigh (UK), Stephan Geier (Germany), Paul Groot (The Netherlands), Artie Hatzes (Germany), Michael Jura (USA), Mukremin Kilic (USA), Orsola De Marco (Australia), Pierre Maxted (UK), Gijs Nelemans (The Netherlands), Sonja Schuh (Germany), Roberto Silvotti (Italy). Local Organising Committee: Horst Drechsel (chair), Edith Day, Stephan Geier, Uli Heber, Andreas Irrgang, Ingo Kreykenbohm, Thomas Kupfer, Veronika Schaffenroth, Florian Schiller, Fritz-Walter Schwarm. Patronage: Andreas Starke, Lord Mayor of Bamberg. Sponsors: Deutsche Forschungsgemeinschaft (DFG), Friedrich-Alexander-Universität Erlangen- Nürnberg, Erlangen-Centre for Astroparticle Physics.
1808.03010
2
1808
2018-10-09T22:54:54
The California-Kepler Survey. VI: Kepler Multis and Singles Have Similar Planet and Stellar Properties Indicating a Common Origin
[ "astro-ph.EP" ]
The California-Kepler Survey (CKS) catalog contains precise stellar and planetary properties for the \Kepler\ planet candidates, including systems with multiple detected transiting planets ("multis") and systems with just one detected transiting planet ("singles," although additional planets could exist). We compared the stellar and planetary properties of the multis and singles in a homogenous subset of the full CKS-Gaia catalog. We found that sub-Neptune sized singles and multis do not differ in their stellar properties or planet radii. In particular: (1.) The distributions of stellar properties $M_\star$, [Fe/H], and $v\mathrm{sin}i$ for the Kepler sub Neptune-sized singles and multis are statistically indistinguishable. (2.) The radius distributions of the sub-Neptune sized singles and multis with $P > 3$ days are indistinguishable, and both have a valley at $\sim1.8~R_\oplus$. However, there are significantly more detected short-period ($P < 3$ days), sub-Neptune sized singles than multis. The similarity of the host star properties, planet radii, and radius valley for singles and multis suggests a common origin. The similar radius valley, which is likely sculpted by photo-evaporation from the host star within the first 100 Myr, suggests that planets in both singles and multis spend much of the first 100 Myr near their present, close-in locations. One explanation that is consistent with the similar fundamental properties of singles and multis is that many of the singles are members of multi-planet systems that underwent planet-planet scattering.
astro-ph.EP
astro-ph
Draft version October 11, 2018 Typeset using LATEX preprint2 style in AASTeX62 The California-Kepler Survey. VI: Kepler Multis and Singles Have Similar Planet and Stellar Properties Indicating a Common Origin∗ Lauren M. Weiss,1, 2, 3, 4 Howard T. Isaacson,5 Geoffrey W. Marcy,5 Andrew W. Howard,6 Erik A. Petigura,6, 7 Benjamin J. Fulton,8, 6, 9 Joshua N. Winn,10 Lea Hirsch,5 Evan Sinukoff,1, 11 and Jason F. Rowe12 The California Kepler Survey 8 1 0 2 t c O 9 . ] P E h p - o r t s a [ 2 v 0 1 0 3 0 . 8 0 8 1 : v i X r a 1Institute for Astronomy, University of Hawaii at Manoa, Honolulu, HI, USA 2Parrent Fellow 3Université de Montréal, Montréal, QC, Canada 4Trottier Fellow 5University of California at Berkeley, Berkeley, CA, USA 6California Institute of Technology, Pasadena, CA, USA 7NASA Sagan Fellow 8NASA Exoplanet Science Institute, Pasadena, CA, USA 9Texaco Fellow 10Princeton University, Princeton, NJ, USA 11NSERC Graduate Research Fellow 12Bishops University, Sherbrooke, QC, Canada (Accepted October 01, 2018) Submitted to The Astronomical Journal ABSTRACT The California-Kepler Survey (CKS) catalog contains precise stellar and planetary properties for the Kepler planet candidates, including systems with multiple detected transiting planets ("multis") and systems with just one detected transiting planet ("sin- gles," although additional planets could exist). We compared the stellar and planetary properties of the multis and singles in a homogenous subset of the full CKS-Gaia cat- alog. We found that sub-Neptune sized singles and multis do not differ in their stellar properties or planet radii. In particular: (1.) The distributions of stellar properties M(cid:63), [Fe/H], and v sin i for the Kepler sub Neptune-sized singles and multis are statistically indistinguishable. (2.) The radius distributions of the sub-Neptune sized singles and multis with P > 3 days are indistinguishable, and both have a valley at ∼ 1.8 R⊕. However, there are significantly more detected short-period (P < 3 days), sub-Neptune sized singles than multis. The similarity of the host star properties, planet radii, and radius valley for singles and multis suggests a common origin. The similar radius valley, ∗ Based on observations obtained at the W. M. Keck Observatory, which is operated jointly by the University of California and the California Institute of Technology. Keck time was granted for this project by the University of California, and California Institute of Technology, the University of Hawaii, and NASA. 2 which is likely sculpted by photo-evaporation from the host star within the first 100 Myr, suggests that planets in both singles and multis spend much of the first 100 Myr near their present, close-in locations. One explanation that is consistent with the simi- lar fundamental properties of singles and multis is that many of the singles are members of multi-planet systems that underwent planet-planet scattering. Keywords: catalogs, stars: fundamental parameters, planets and satellites: fundamental parameters, planets and satellites: formation 1. INTRODUCTION Comparisons between planetary systems with multiple planets and those with just one known planet have long been used to probe planet for- mation. A decade after the discovery of the first multi-planet system around a main sequence star (Butler et al. 1999), Wright et al. (2009) conducted a statistical study of 28 multi-planet systems, all of which were discovered and char- acterized with radial velocities. They compared the multi-planet systems to systems with only one known planet and found that multi-planet systems were spaced uniformly in log-period (unlike the single-planet systems) and typically had lower eccentricities and m sin i values than the single-planet systems. More recently, the Kepler Mission (Borucki et al. 2010) has detected hundreds of multi- planet systems (Latham et al. 2011; Lissauer et al. 2011; Fabrycky et al. 2014; Lissauer et al. 2014; Rowe et al. 2014). In the Kepler multi- planet systems, multiple planet candidates tran- sit the star, resulting in measured orbital pe- riods, planet-to-star radius ratios, and transit durations for each planet. The vast majority of the Kepler planet candidates in multis are bona- fide planets, based on statistical arguments (Lis- sauer et al. 2012, 2014). The Kepler multi- planet systems differ from the previously stud- ied RV multi-planet systems in that Kepler was sensitive to smaller (lower-mass) planets. The majority of the Kepler single-planet and multi- planet systems have sub-Neptune sized planets rather than giant planets (Latham et al. 2011). Also, Kepler only detected transiting planets. In systems with multiple transiting planets, the planets are very likely nearly coplanar by virtue of the fact that they all transit (Lissauer et al. 2011). However, not all multi-planet sys- tems must be nearly coplanar. A sufficiently non-coplanar system might result in only one transiting planet detected by Kepler, although multiple planets might exist. The systems with just one detected transiting planet ("singles") might belong to the tail of a single underlying distribution that describes systems with multi- ple detected transiting planets ("multis"). On the other hand, a high fraction of the singles might belong to a population with different for- mation conditions or a different dynamical his- tory. We would like to understand whether the Ke- pler singles and multis differ in their orbital and physical parameters. Some orbital parameters of interest include multiplicity, orbital periods, eccentricities, and inclinations. Physical param- eters of interest include host star mass, metal- licity, and rotation velocity, as well as planet ra- dius and mass. If the singles differ from the mul- tis in their underlying distributions of orbital and/or physical parameters, such a distinction likely points to a divergence in the planet for- mation and/or evolution of the Kepler singles versus multis. Past research has considered the hypothesis that a large fraction of the Kepler singles be- long to a distinct population from the multi- planet systems. Some examples of a distinct population are a dynamically hot population (high mutual inclinations and eccentricities for the singles) or a population with wider spacing in the orbital period ratios for the singles than is typical for the multis. Lissauer et al. (2011) found that the typical mutual inclinations in the Kepler multis were < 10◦ and noted that these small mutual inclinations seemed inconsis- tent with the large number of observed singles. Hansen & Murray (2013) explored the multi- plicity vectors and period distributions of the Kepler singles and multis through a model of in situ planet formation. They found that the number of Kepler singles is too high to result from an in situ formation scenario (although the authors required each system to have at least three initially coplanar planets). In an- other study that required a minimum number of planets per system, Ballard & Johnson (2016) found that there is an excess of singles among the Kepler M-dwarfs. Xie et al. (2016) used stellar spectra from LAMOST and the transit durations from Kepler lightcurves to estimate of the mean eccentricities and inclinations for singles and multis. They found that the mean eccentricity of the singles was ∼ 0.3, whereas the multis were on nearly circular orbits (e = 0.04 ± 0.04). In a sample of stars with aster- oseismically determined properties, Van Eylen et al. (2018) also found higher eccentricities for the singles than the multis. However, other studies have found no need for a large fraction of the singles to have distinct underlying architectures. Ford et al. (2011) found that the prevalence of TTVs in singles was consistent with the multis, suggesting that many singles belong to compact, multi-planet systems. Tremaine & Dong (2012) explored a variety of possible orbital geometries and found that no separate population was needed to ex- plain the apparent excess of Kepler singles, if high mutual inclinations were allowed in a small 3 fraction of the multis. Fang & Margot (2012) modeled the transit duration ratios as well as the transiting planet multiplicity. They found that an underlying distribution in which most multi-planet systems have mutual inclination distributions of < 3◦, and 75% of systems have 1-2 planets with P < 200 days (like the solar system) describes the observed planet multiplic- ities and transit duration ratios. Gaidos et al. (2016) found that with improved stellar param- eters and an exponentially-distributed number of planets per star, the large number of M dwarf singles compared to multis announced in Bal- lard & Johnson (2016) could be reconciled. Zhu et al. (2018) used spectra from LAMOST to measure the properties of Kepler planet candi- date host stars, giving special attention to the differences between multis and singles that did and did not exhibit transit timing variations. They found that the stellar properties of the singles and multis did not differ substantially. Munoz Romero & Kempton (2018) compared the metallicities determined by the California- Kepler Survey (described below) for singles and multis and found no significant differences in the stellar metallicities. We push the comparison of the fundamental properties of the Kepler singles versus multis into new regions of parameter space by lever- aging the precise stellar and planetary param- eters of The California-Kepler Survey (CKS) combined with Gaia DR2. CKS obtained high- resolution (R=60,000) spectra for 1305 Kepler systems with transiting planets (Petigura et al. 2017). The improved stellar and planetary pa- rameters (Johnson et al. 2017, CKS II) enable a more accurate and precise characterization of the Kepler systems than was previously avail- able, yielding 2025 transiting planet candidates with precise radii and host star properties. Ful- ton & Petigura 2018 (CKS VII) revised the stel- lar properties and planet radii based on paral- laxes from the Gaia DR2 catalog (Gaia Collab- 4 oration et al. 2018). CKS and Gaia have dra- matically improved the characterization of the Kepler stellar radii, metallicities, masses, and rotations, as well as the planet radii and equi- librium temperatures, compared to what was available before the CKS project (e.g., Brown et al. 2011). In this paper (CKS VI), we use the refined stellar and planetary properties presented in CKS VII to compare a large, homogeneous, high-purity sample of Kepler singles and multis. Where applicable, we also examine how the stel- lar and planetary properties of the multis differ for system with 2, 3, and 4 or more transiting planets. In section 2, we discuss the cuts to the CKS catalog needed to generate homogenous samples for comparison. In section 3, we compare the distributions of the stellar properties for the sin- gles vs. the multis. In section 4, we compare the distributions of the planet radii and orbital pe- riods for the singles vs. the multis. We conclude in section 5. 2. THE SAMPLE By construction, CKS was not a homogenous survey (Petigura et al. 2017). The largest com- ponent of CKS is Kepler planet hosts with Kp < 14.2. However, CKS was expanded to in- clude fainter Kepler stars that addressed special interests, including ultra-short period planets, planets in the habitable zone, and multi-planet systems. Since this paper addresses multi-planet sys- tems, we would like to include the full breadth of multi-planet systems wherever possible. How- ever, the population of CKS singles is heteroge- nous: most orbit stars with Kp < 14.2, and any singles orbiting fainter stars are ultra-short pe- riod planets or habitable-zone planets. Thus, the sample of singles is only homogenous for Kp < 14.2. The different magnitude limits of homogenous sub-samples of the singles and multis are prob- lematic because fundamental stellar properties, in particular stellar mass and radius, are cor- related with stellar magnitude. Therefore, to fairly compare the singles and multis, we must down-select the multis to those with Kp < 14.2. However, in comparing multis to each other, we can use the full sample of CKS-Gaia. In addition to ensuring homogenous samples for comparison, we make several cuts to ensure high-precision stellar and planetary parameters. 2.1. Selecting High-Purity Planet Samples The CKS planet candidates and the succes- sive cuts we made are summarized in Table 1. The initial CKS dataset consists of 1944 sig- nals that were at one time flagged as transit- ing planet candidates, orbiting 1222 stars that have Gaia properties reported in DR2. From these, we discarded the signals that are now known to be false positives (as determined on either the NASA Exoplanet Archive or in CKS I), removing 156 non-planetary signals around 104 stars. We then discarded stars that are di- luted by at least 5% by a second star in the Kepler aperture (as determined in the stellar companion catalog of Furlan et al. 2017), re- moving 88 planet candidates around 58 stars. We discarded planets for which Mullally et al. (2015) measured b > 0.9, for which the high impact parameters adversely affected our abil- ity to determine accurate planet radii, removing 137 planet candidates from around 70 stars. We removed planet candidates for which the mea- sured signal-to-noise ratio (SNR) is less than 10 since these planets have poorly determined radii and impact parameters, removing 48 planet candidates. We also removed planet candidates with Rp > 22.4 R⊕, which are likely eclips- ing binaries rather than planets. Of the four planet candidates with Rp > 22.4 R⊕, all four were singles, and three orbited giant stars with log g < 3.9. Systems that were originally multis but had been purified to the extent that only one planet remained were excluded. Table 1. Successive Cuts a N(cid:63) Ntp,multi Ntp 1944 1222 1788 1118 1700 1060 990 1563 1563 990 952 1495 948 1491 700 997 843 578 1176 1092 1042 940 940 908 892b 492 434c Cut No FPs dilution < 5% b < 0.9 Rp/R(cid:63) < 0.5 SNR > 10.0 Rp < 22.4 Kp < 14.2 SNR 1.5 R⊕, 30 days > 10.0 aNumber of transiting planets bThese are the "CKS Multis" sample cThese are the Bm sample After these cuts, our sample included 892 high-purity planet candidates in multi-planet systems around 349 stars1. The number of plan- ets and stars in the various subsamples are sum- marized in Table 4. In this sample of multis, we compare the systems with 2 transiting plan- ets (446 planets around 223 stars), 3 transiting planets (228 planets around 76 stars), and 4+ transiting planets (218 planets around 50 stars). Figure 1 shows the number of stars with various multiplicities (blue histogram). We present the catalog of planets in the high-purity sample of multis in Table 2. 2.2. Selecting Singles and Multis for Comparison As discussed above, a homogenous compari- son of singles and multis can only be performed for Kp < 14.2. In the CKS sample, we ob- serve a significant correlation between Kepler magnitude and stellar mass (Pearson r = −3, p < 10−5, see Figure 2). This is in part be- 1 This number differs slightly from (Weiss et al. 2018) because a few stars from that study did not have paral- laxes in Gaia DR2. 5 cause host star apparent magnitude correlates with luminosity (Malmquist 1922), which cor- relates with both stellar mass and radius. In addition, the 150,000 stars selected for moni- toring in the Kepler mission were chosen based on both their magnitudes and their colors (as a proxy for spectral type), and so the selection criteria might have contributed to the correla- tion. Hence, a common apparent magnitude cut for the singles and multis ensures that any ob- served difference in the host star properties is astrophysical rather than the result of selection biases. From the multis in the purified sample, we down-selected to those orbiting stars with Kp < 14.2, resulting in 426 planets orbiting 166 stars. Among both the singles and multis, we did not want to include systems for which it would have been unlikely to detect additional planets, and so we made a final cut based on the de- tectability of a hypothetical planet around each star in our sample. To determine the ease of detecting planets around a given star, we cal- culated the model signal-to-noise ratio (SNR) for a 1.5 R⊕ planet orbiting at 30 days around each of the single and multi host stars via the following equations: (Rp/R(cid:63))2(cid:112)3.5yr/P (cid:112)6hr/T CDPP6h SNR = (1) T = 13hr (P/1yr)1/3(ρ(cid:63)/ρ(cid:12))−1/3 (2) where Rp is the planet radius, R(cid:63) is the stel- lar radius, ρ(cid:63)/ρ(cid:12) is the stellar density in units of solar density, and CDPP6h is the combined differential photometric precision in the Kepler light curve over 6 hours. We remove all sin- gles and multis stars for which the model SNR < 10, ensuring that a planet of 1.5 R⊕ at 30 days would have been detectable around all the stars in our sample. Our final sample of Kp < 14.2 singles, hereafter Bs, contains 376 singles. Our final sample of Kp < 14.2 multis, hereafter Bm, 6 Figure 1. Histograms of the number of stars with various transiting planet multiplicities in the full CKS-Gaia multis sample (blue) and in the magnitude-limited Bs + Bm samples (orange). contains 166 stars hosting 426 planets. The Bm and Bs samples together comprise the orange histogram in Figure 1. The planet and stellar properties of the Bm and Bs samples are listed in Table 3. We tabulate the number of stars and plan- ets in the initial CKS-Gaia sample, the cleaned sample of multis, the subsets with 2, 3, or 4+ transiting planets (labeled Ntp = 2, Ntp = 3, and Ntp ≥ 4), and the Kp < 14.2 singles and multis (labeled Bs and Bm) in Table 4. In ad- dition, we include the subsets of Bs and Bm in which the planets are smaller than 4 R⊕ (Bs,4 and Bm,4). The host star effective temperatures and radii of the CKS stars, the high-purity mul- tis, Bs, and Bm are summarized in Figure 3. As a sanity check, we compare the magnitudes of the host stars of the Kp < 14.2 singles and multis. The distribution of host star magni- tudes is indistinguishable for Bs and Bm, sup- porting that we have selected samples of sin- gles and multis with similar host star bright- nesses. We also find that the CDPP over 6 hour timescales is indistinguishable for Bs and Bm. From the similarity of the CDPP distributions, we conclude that we have not inadvertently se- Figure 2. Stellar mass vs. Kepler magnitude for the purified CKS-Gaia singles (gray circles) and multis (black squares), and the homogenized, magnitude-limited singles (red) and multis (blue). There is a correlation between stellar mass and magnitude, motivating a homogenous, magnitude- limited sample of singles and multis to ensure a fair comparison of their host star properties. Figure 3. Stellar radius vs. effective tempera- ture for the purified CKS-Gaia singles (gray circles) and multis (black squares), and the homogenized, magnitude-limited singles (red) and multis (blue). lected photometrically noisy stars among either the singles or the multis (see Figure 4). 3. STELLAR PROPERTIES We present the distributions of host star and planetary properties among several samples: 1234567Number of Transiting Planets050100150200250300350Number of Stars0376223106763736151162111CKS-Gaia Multis StarsKp < 14.2 Stars810121416Kepler Magnitude0.60.81.01.21.41.61.82.0Stellar Mass [Solar]CKSCKS multisms450047505000525055005750600062506500Teff [K]110Stellar Radius [Solar]CKSCKS multisSingles Kp < 14.2Multis Kp < 14.2 Table 2. High Purity CKS-Gaia Multis 7 KOI Teff (K) log g [Fe/H] v sin i (km s−1) Kepmag CDPP6h (ppm) 23.33 23.33 23.33 54.61 54.61 39.42 39.42 39.42 39.42 39.42 11.2 11.2 11.2 13.77 13.77 12.5 12.5 12.5 12.5 12.5 4.11 K00041.01 4.11 K00041.02 4.11 K00041.03 4.06 K00046.01 4.06 K00046.02 4.51 K00070.01 4.51 K00070.02 4.51 K00070.03 4.51 K00070.04 K00070.05 4.51 aNumber of transiting planets surviving our cuts. Note -- This table is downloadable in full online. A portion has been reproduced here for form and content. 5873.54 5873.54 5873.54 5686.15 5686.15 5483.75 5483.75 5483.75 5483.75 5483.75 0.09 0.09 0.09 0.38 0.38 0.08 0.08 0.08 0.08 0.08 2.7 2.7 2.7 2.5 2.5 0 0 0 0 0 3 3 3 2 2 5 5 5 5 5 Period Rp Ntp a (days) 12.82 6.89 35.33 3.49 6.03 10.85 3.7 77.61 6.1 19.58 (R⊕) 2.36 1.35 1.54 6.19 1.29 2.93 2.04 2.53 0.8 0.96 M(cid:63) (M(cid:12)) 1.1 1.1 1.1 1.24 1.24 0.95 0.95 0.95 0.95 0.95 R(cid:63) (R(cid:12)) 1.54 1.54 1.54 1.72 1.72 0.89 0.89 0.89 0.89 0.89 Table 3. Bright Multis and Singles R(cid:63) (R(cid:12)) 1.05 0 KOI log g [Fe/H] v sin i (km s−1) Kepmag CDPP6h (ppm) 17.72 21.36 30.05 42.06 46.12 51.55 50.59 23.33 23.33 23.33 11.34 10.46 12.21 13.3 13.37 13.44 13.44 11.2 11.2 11.2 2 -0.01 0.18 0.17 0.34 K00001.01 K00002.01 K00007.01 K00017.01 K00018.01 K00020.01 K00022.01 K00041.01 K00041.02 K00041.03 aModel SNR for a 1.5 R⊕ planet at 30 days. bNumber of transiting planets surviving our cuts. Note -- This table is downloadable in full online. A portion has been reproduced here for form and content. 4.39 4.02 4.13 4.26 4.08 4.1 4.26 4.11 4.11 4.11 1.54 1.28 1.74 1.56 1.29 1.54 1.54 1.54 1.3 5.2 2.8 2.6 4.4 1.7 1.9 2.7 2.7 2.7 0.02 0.18 0.09 0.09 0.09 a Period Rp Model SNR Ntp b (days) 2.47 2.2 3.21 3.23 3.55 4.44 7.89 12.82 6.89 35.33 (R⊕) 14.24 16.44 4.16 13.35 15.25 20.12 13.28 2.36 1.35 1.54 61.72 25.56 20.67 16.02 12.41 10.54 14.31 26.16 26.16 26.16 1 1 1 1 1 1 1 3 3 3 Teff (K) 5820.3 6448.66 5844.82 5664.08 6326.65 5945.39 5880.28 5873.54 5873.54 5873.54 M(cid:63) (M(cid:12)) 0.99 1.53 1.16 1.1 1.32 1.1 1.13 1.1 1.1 1.1 8 Figure 4. Left, top: The host star magnitude in the Kepler bandpass for systems with one transiting planet (red) and multiple transiting planets (blue) in the magnitude-limited CKS samples Bs and Bm. Right, top: the same, but the cumulative distribution function (CDF). An Anderson-Darling test indicates there is no significant distinction between the magnitudes of stars that host one versus multiple transiting planets for Kp < 14.2 (p= 0.75). Bottom: same as top, but for the combined differential photometric precision (CDPP) in the Kepler bandpass over 6 hour timescales. There is no significant distinction between the photometric noise of our final samples of singles vs. multis. These sanity checks show that our selection criteria have not inadvertently favored bright or quiet stars for either the singles or the multis. 891011121314Kepler Magnitude01020304050Number of StarsKp < 14.2 SinglesKp < 14.2 Multis891011121314Kepler Magnitude0.00.20.40.60.81.0CDFp = 0.75Kp < 14.2 SinglesKp < 14.2 Multis0204060801001206 hour CDPP [ppm]0102030Number of StarsKp < 14.2 SinglesKp < 14.2 Multis0204060801001206 hour CDPP [ppm]0.00.20.40.60.81.0CDFp = 0.37Kp < 14.2 SinglesKp < 14.2 Multis Table 4. The Samples Name Description High-Purity CKS Multis 2 transiting planets 3 transiting planets 4+ transiting planets CKS Multis Ntp = 2 Ntp = 3 Ntp ≥ 4 Bs Bs,4 Bm Bm,4 aNumber of transiting planets bMagnitude limit in the Kepler bandpass. ...of which Rp < 4 R⊕ Bright (Kp < 14.2b) Singles ...of which Rp < 4 R⊕ Bright (Kp < 14.2b) Multis a N(cid:63) Ntp 892 349 446 223 76 228 218 50 376 376 342 342 426 166 415 169 Kp < 14.2 singles, Kp < 14.2 multis, all the CKS-Gaia multis, and the subsets of CKS-Gaia multis with 2, 3, and 4+ transiting planets. The stars have effective temperatures from 4500 to 6300 K, masses from 0.5 to 1.6 M(cid:12), radii from to 0.6 to 2.1 R(cid:12), and projected ro- tation velocities of < 20 km s−1. The uncer- tainties in stellar effective temperature, mass, radius, metallicity, v sin i, and age are typically 60 K, 0.03 M(cid:12), 0.03 R(cid:12), 0.04 dex, 1 km s−1, and 1 Gyr, respectively. We compared the magnitude-limited singles and multis, Bs and Bm. Figure 5 shows how the stellar mass, metallicity, and projected ro- tation velocity distributions of the singles and multis differ. The panels on the left are his- tograms of the number of stars; the panels on the right are normalized cumulative distribution functions. Comparing the distributions of sin- gles and multis with Anderson-Darling tests, we found: • no significant difference between the stel- lar mass (M(cid:63)) distributions of the singles and the multis (p = 0.47), 9 • no significant difference between the stel- lar metallicity ([Fe/H]) distributions of the singles and the multis (p = 0.29), • no significant difference between the pro- jected stellar rotation (v sin i) distribu- tions of the singles and the multis (p = 0.83). The stellar effective temperatures, radii, and isochrone-determined ages also do not differ significantly. The host star properties of the Kp < 14.2 singles and multis are summarized in Table 5. In a recent study of the CKS singles and mul- tis, Munoz Romero & Kempton (2018) com- pared the CKS-determined metallicities of the singles and multis. Their sample selection re- moved evolved stars, but did not make the mag- nitude cuts and detectability cuts we used here. Nonetheless, they also found no significant dif- ferences between the host star metallicities of the CKS singles and multis. We found no difference in the stellar v sin i dis- tributions of the singles vs. multis. However, the projected rotational velocities are only well- calibrated for 2 km s−1 < v sin i < 20 km s−1 (Petigura 2015). Thus, the sample of singles and multis with measurable v sin i is smaller than the Bs and Bm samples: only 245 sin- gle stars and 112 multi stars have 2 km s−1 < v sin i < 20 km s−1. We also compared stars with different transit- ing planet multiplicities using the full CKS-Gaia sample. We found no significant differences in the distributions of stellar properties for the sys- tems with 2, 3, and 4+ transiting planets (Fig- ure 6). The lack of significant correlations between host star properties and planet multiplicities suggests that, if there is some divergence in the planet evolution of the singles vs. multis, such evolution is not dependent on host star type. 10 Figure 5. Top left: a histogram of the host star masses for systems with one transiting planet (red) and multiple transiting planets (blue) in the magnitude-limited CKS samples Bs and Bm. Top right: the same, but the cumulative distribution function (CDF). The typical uncertainty in stellar mass is 3%. The subsequent rows compare the singles vs. multis distributions for stellar metallicity [Fe/H] (typical uncertainty 0.04 dex), and projected rotation velocity v sin i (typical uncertainty 1 km s−1). With Anderson-Darling tests, we find no significant distinctions between the distributions of singles and multis for any of these stellar parameters (p > 0.01). 0.60.81.01.21.41.6M [Solar]0102030Number of StarsKp < 14.2 SinglesKp < 14.2 Multis0.60.81.01.21.41.6M [Solar]0.00.20.40.60.81.0CDFp = 0.47Kp < 14.2 SinglesKp < 14.2 Multis0.60.40.20.00.20.4[Fe/H]010203040Number of StarsKp < 14.2 SinglesKp < 14.2 Multis0.60.40.20.00.20.4[Fe/H]0.00.20.40.60.81.0CDFp = 0.29Kp < 14.2 SinglesKp < 14.2 Multis2.002.834.005.668.0011.316.0vsini [km/s]0510152025Number of StarsKp < 14.2 SinglesKp < 14.2 Multis2.002.834.005.668.0011.316.0vsini [km/s]0.00.20.40.60.81.0CDFp = 0.83Kp < 14.2 SinglesKp < 14.2 Multis 11 Figure 6. The same as Figure 5, but for the purified CKS-Gaia sample. The subsets with Ntp = 2 (green), Ntp = 3 (cyan), and Ntp ≥ 4 (violet) are shown. With Anderson-Darling tests between Ntp = 2 and Ntp > 2, we find no significant distinctions between the distributions of stellar properties for the various planet multiplicities. 0.60.81.01.21.41.6M [Solar]0102030Number of Stars2 tp.3 tp.4+ tp.All CKS multis0.60.81.01.21.41.6M [Solar]0.00.20.40.60.81.0CDFp = 0.212 tp.3 tp.4+ tp.All CKS multis0.60.40.20.00.20.4[Fe/H]010203040Number of Stars2 tp.3 tp.4+ tp.All CKS multis0.60.40.20.00.20.4[Fe/H]0.00.20.40.60.81.0CDFp = 0.382 tp.3 tp.4+ tp.All CKS multis2.002.834.005.668.0011.316.0vsini [km/s]0510152025Number of Stars2 tp.3 tp.4+ tp.All CKS multis2.002.834.005.668.0011.316.0vsini [km/s]0.00.20.40.60.81.0CDFp = 0.782 tp.3 tp.4+ tp.All CKS multis 12 4. PLANET PROPERTIES We examined the distributions of singles and multis in the planet radius-orbital period plane (Figure 7). The fraction of detected planets that are singles, f, in each grid cell is given. The uncertainty in f is the 68% confidence in- terval calculated from binomial statistics. Note that because our definitions of singles and mul- tis are based on the transiting planet multiplic- ity, the values of f and their uncertainties do not necessarily describe the true planet multiplicity. Rather, f is an observed quantity that should be reproduced by future attempts to model un- derlying planetary architectures. In this sec- tion, we identify regions in period-radius space where contiguous cells have similar values of f and discuss possible interpretations. We follow with a discussion of the the 1-D distributions of planet radius and orbital period for the singles and multis. 4.1. Sub-Neptunes with P > 3 days Sub-Neptune sized planets (Rp < 4R⊕) are the majority of the planets in the CKS-Gaia sample, and they are also intrinsically common (Petigura et al. 2013; Fressin et al. 2013). Most sub-Neptunes have P > 3 days (we will explore the population with P < 3 days below). Figure 7 shows the sub-Neptunes with P > 3 days and 0.5 R⊕ < Rp < 4 R⊕ in a cyan box. Within the cyan box, a detected planet is slightly less likely to be a single than a multi (f = 43± 2%). How- ever, there is very little variation in the frac- tion of singles as a function of period and radius within the cyan box. Thus, for the majority of the Kepler planets, the orbital period and size of the planet are not good predictors of whether the planet will have additional transiting com- panions. The sub-Neptunes dominate the sample, hence their host star properties are very similar to the host star properties of the sample over- all. The mean mass of the sub-Neptunes with P > 3 days is 1.01 M(cid:12)for the singles (1.01 for the multis), and the mean value of [Fe/H] is -0.03 for the singles (-0.02 for the multis). 4.2. Hot Jupiters The hot Jupiters (P < 10 days, 8 R⊕ < Rp < 22 R⊕, yellow box in Figure 7) in the CKS- Gaia sample are all singles. The high fraction of singles among the hot Jupiters confirms the well-studied phenomenon that hot Jupiters are lonely (e.g., Steffen et al. 2012). Recall that in our sample of hot Jupiters, we would have been able to detect any transiting planets with Rp > 1.5 R⊕ and P < 30 days, based on our selection critera. If the next hot Jupiter discovered had transiting companions, the hot Jupiters would have f = 92±6%. Combining this fraction with the overall occurrence of hot Jupiters (∼ 1%), hot Jupiters with nearby coplanar companions occur around no more than ∼ 0.1% of stars. Thus, WASP-47, the only known hot Jupiter with small transiting companions (Becker et al. 2015), belongs to a rare population. Hot Jupiter host stars have higher masses and metallicities than field stars (Johnson et al. 2007; Fischer & Valenti 2005). We compare the host star properties of the hot Jupiters to the host star properties of the small exoplan- ets (Rp < 4 R⊕). The hot Jupiters orbit more massive and more metal-rich stars than the sub- Neptunes. The mean mass and metallicity of a hot Jupiter host star are M(cid:63) = 1.17 M(cid:12) and [Fe/H] = 0.16, whereas the mean mass and metallicity for a sub-Neptune's host star are M(cid:63) = 1.01 M(cid:12) and [Fe/H] = 0.00. The typical uncertainties in the stellar masses and metal- licities are 3% and 0.04 dex, respectively, and Anderson-Darling tests yield with > 99% confi- dence (p < 0.01) that the host star masses and metallicities of the hot Jupiters are drawn from different distributions than the host star masses and metallicities of the sub-Neptunes. 4.3. Cold Jupiters 13 Figure 7. The CKS singles (red) and multis (blue) as a function of orbital period and planet size. The fraction of detected planets that are in singles, f, is displayed in each grid cell with at least 4 planets. The uncertainties are calculated using binomial statistics. The fraction of detected sub-Neptunes with P > 3 days (i.e., the majority of the the Kepler planets) that are singles is 43 ± 2% (cyan box). However, among the detected sub-Neptunes with P < 3 days, 68± 5% are singles (magenta box). The 11 hot Jupiters (yellow box) are all singles. The population of cold giants (orange box) is also predominantly singles (71 ± 9%). We explore each of these regions of parameter space in the text. 0.321.03.1610.031.62100.0316.231000.0Orbital Period [days]0.250.51.02.04.08.016.032.0Planet Size [RE]f = 0.75 ± 0.12f = 0.92 ± 0.04f = 0.73 ± 0.07f = 0.49 ± 0.08f = 0.86 ± 0.07f = 0.41 ± 0.06f = 0.45 ± 0.04f = 0.33 ± 0.06f = 0.80 ± 0.10f = 1.00 ± 0.00f = 0.25 ± 0.12f = 0.42 ± 0.05f = 0.43 ± 0.05f = 0.29 ± 0.18f = 0.39 ± 0.10f = 0.49 ± 0.06f = 0.40 ± 0.20f = 0.17 ± 0.19f = 0.48 ± 0.11f = 0.67 ± 0.13f = 0.71 ± 0.12f = 0.83 ± 0.08Kp < 14.2 singlesKp < 14.2 multis 14 Stellar Properties Kp CDPP (ppm) R(cid:63) (R(cid:12)) Teff [K] M(cid:63) (M(cid:12)) [Fe/H] v sin i (km/s)b Age (Gyr) Planet Properties Rp Per (days) Planet Properties (Rp < 4 R⊕) Rp Per (days) Planet Properties (Rp < 4 R⊕ & P > 3 days) Rp Per (days) aAnderson-Darling test p-value bFor v sin i > 2 km s−1 Table 5. Statistics of Kp < 14.2 Singles vs. Multis Parameter Singles Multis Unc. p-value a Mean RMS Mean RMS 0.96 13.08 13.08 0.98 51.71 18.23 52.12 21.36 0.31 1.19 391 5737 1.01 0.15 0.17 -0.01 2.74 4.86 5.69 3.27 1.18 5733 1.02 0.01 4.68 5.51 0.33 398 0.15 0.17 3.03 3.38 0.03 53 0.03 0.04 1.00 1.54 2.74 2.51 35.97 89.60 28.14 67.39 0.00021 0.09 2.03 1.42 0.79 1.76 22.27 42.49 23.01 39.65 0.00021 0.09 0.73 1.80 0.78 1.86 26.11 45.29 24.47 40.60 0.00023 0.09 1.84 0.73 0.66 0.36 0.78 0.70 0.43 0.22 0.86 0.68 0.02 0.001 0.072 0.002 0.44 0.32 Cold Jupiters are giant planet candidates with P > 100 days, 5 R⊕ < Rp < 16 R⊕ (orange box in Figure 7). The fraction of singles among the cold Jupiters is high compared to the sub- Neptunes (f = 71 ± 9%), but not as high as the hot Jupiters. Formally, the excess of gi- ant cold singles (compared to the near-parity of sub-Neptune sized singles and multis) is sig- nificant with 3σ confidence, but concluding that most of the cold giant planets are indeed singles is premature. Not all of the long-period giant planet candidates are confirmed, and false pos- itives are common for planets of these sizes and orbital periods (e.g., as many as 35-50% of the unconfirmed giant planets could be false posi- tives, Santerne et al. 2016). We have already excluded known and likely false positives from the CKS sample. However, only 9 out of 17 cold giant singles are already confirmed. (The giant planets in multis are statistically validated, e.g. Lissauer et al. 2012). Vetting the remainder of the single, cold giant planet candidates would clarify whether the excess of single cold giants (compared to sub-Neptunes) is real. The host star properties of the single, cold giant planet candidates do not differ substan- tially from the host star properties of the sub- Neptunes (Anderson-Darling p > 0.1 in com- parisons of stellar mass, metallicity, and v sin i). The average mass of the cold Jupiter host stars is 1.07 M(cid:12), which is slightly higher than the typical stellar mass for the sub-Neptunes (1.0 M(cid:12)), but lower than the typical host-star mass for hot Jupiters (1.17 M(cid:12)). The average metal- licity of the cold giant planet host stars is [Fe/H] = 0.01, which does not differ significantly from the average metallicity of the sub-Neptune host stars ([Fe/H] = 0.0). 4.4. Hot Sub-Neptunes In particular, there is an excess of singles rel- ative to multis among the hot sub-Neptunes (Rp < 4 R⊕, P < 3 days, magenta box in Figure 7). There are 80 planets in this size and period 15 range, of which 54 are singles (f = 68 ± 5%), resulting in a significantly higher fraction of sin- gles than for the sub-Neptunes with P > 3 days (f = 43 ± 2%). The vast majority (74/80) of the sub-Neptunes with P < 3 days have Rp < 1.8 R⊕. There is mounting evidence that planets smaller than 1.8 R⊕ and close to their stars are rocky: the masses of many planets with Rp < 1.5 R⊕ and P < 100 days are consistent with rocky com- positions (Weiss & Marcy 2014; Rogers 2015). Furthermore, there are two distinct size pop- ulations of small planets, with a valley in the planet radius distribution at 1.8 R⊕ (Fulton et al. 2017). Planets smaller than 1.8 R⊕ and near their stars are likely the rocky cores of photo-evaporated planets (Fulton et al. 2017; Owen & Wu 2017). Hence, the planets with P < 3 days and Rp < 1.8 R⊕ might better be described as "hot super-Earths" than sub- Neptunes. The host star properties of the hot sub- Neptunes are very similar to the host star prop- erties of the sub-Neptunes with P > 3 days. The metallicities are slightly higher on aver- age ([Fe/H] = 0.01) for the hot sub-Neptunes than the cool sub-Neputunes ([Fe/H] = −0.03). However, an Anderson-Darling test yields p > 0.1, indicating that the distribution of host star metallicities for the hot super-Earths is consis- tent with the distribution of host star metal- licities for the cool sub-Neptunes. Petigura et al. (2018) also found slightly higher host star metallicities for the hot super-Earths than for the cool sub-Neptunes. However, the scatter in host star metallicities for both the singles and multis (RMS=0.17 dex) is much larger than the slight difference in their average host star metallicities (0.04 dex). Steffen & Coughlin (2016) also identified an excess of detected short-period, Earth-sized sin- gles compared to multis. That study used a ker- nel estimation to identify clusters of singles and 16 multis in (Rp, P ) space, whereas we have di- vided (Rp, P ) space into a grid. Our study con- firms the population of detected short-period planets that are predominantly singles. 4.5. Planet Radius The top panel of Figure 8 shows the distribu- tion of planet radii for Bs and Bm. There is a valley in planet radius at 1.8 R⊕ in both the singles and multis. This valley was announced in Fulton et al. (2017) (CKSIII), but that paper did not specifically address whether the valley exists in multi-planet systems. Here, we find that multi-planet systems indeed have a val- ley in the distribution of planet radii at about 1.8 R⊕. The middle panel of Figure 8 shows the dis- tribution of planet sizes for planets that belong to systems with 2, 3, and 4+ transiting plan- ets. The radius valley at 1.8 R⊕ is evident in the complete CKS-Gaia sample (black), which includes multis having Kp > 14.2. This large sample (892 transiting planets) clarifies the exis- tence of the radius valley. The sub-samples with Ntp = 2 (green), Ntp = 3 (cyan), Ntp ≥ 4 (pur- ple) are all consistent with a valley, although the valley is less clear in the high-multiplicity systems simply because there are fewer planets. As discussed above, there is an excess of single giant planet candidates. An Anderson-Darling test comparing the all of the planet radii in Bs and Bm yields a p-value 0.02 (Figure 8). Fo- cusing on the sub-Neptune sized planets with P > 3 days only (cyan box in Figure 7), the p-value is 0.44, indicating no significant differ- ence between the radius distributions of the sin- gles and multis (bottom panel of Figure 8). The CDFs of the sub-Neptune sized singles and mul- tis have the greatest differences near the valley at 1.8 R⊕, but the differences in their distri- butions are not statistically significant. This is because there are not enough sub-Neptune sized planets near the valley to determine whether the shape or position of the valley differs between the singles and multis. Nonetheless, we detect the presence of the valley in the magnitude- limited singles and multis (Bs,4 and Bm,4) as well as in the full sample of CKS-Gaia. The radius valley exists for the planets in sin- gles and multis. If some physical process (such as planet-planet scattering) disrupts multi- planet systems to make the singles, that pro- cess does not erase or significantly alter the radius valley. The radius valley is thought to be sculpted within the first 100 Myr of the planetary system's lifetime. Photo-evaporation, which scales with the inverse square of orbital distance, is likely responsible at least in part for the presence of the radius valley (Owen & Wu 2017), although other mechanisms to strip a planet's volatile envelope have been proposed (Ginzburg et al. 2018). The similar radius distributions thus suggest that the singles and multis have similar migra- tion histories (or lack thereof), and that any large-scale migration likely happened in the first 100 Myr, bringing the planets close enough to the stars for photo-evaporation to produce the radius valley. If the planets of one population (say, the singles) had migrated great distances after photo-evaporation turned off (i.e., after 100 Myr), we would have seen less of a gap in that population, as photo-evaporation is weak at large distances and at late times. 4.6. Planet orbital period The top panels of figure 9 show the distri- bution of planet orbital periods for the sub- Neptunes in Bs,4 and Bm,4. There are more singles at short orbital periods (P < 3 days) than multis. Beyond 10 days, the orbital pe- riods of the singles and multis are similar. An Anderson-Darling comparison of the period dis- tributions of singles vs. multis yields a p-value of 0.001, indicating that these distributions are not drawn from the same underlying distribu- tion with > 99% confidence. 17 Figure 8. Top panel, left: the distribution of planet radius for the CKS Kp < 14.2 systems with one (red) and multiple (blue) transiting planets. Top panel, right: the same, but the cumulative distribution function (CDF). The majority of planets in singles and multis are smaller than 4 R⊕, and both distributions show a valley at 1.8 R⊕. The tail of the singles distribution includes more giant planets than the multis. Middle panel: same as the top panels, but for all of the CKS-Gaia multis (black), and for the sub-samples with 2, 3, and 4+ transiting planets, all of which are consistent with a valley at 1.8 R⊕. Bottom panel: zoom of the valley in singles and multis for 1 R⊕ < Rp < 4 R⊕ and P > 3 days, including Poisson errors. The prevalence of the valley at 1.8 R⊕ in the CKS-Gaia singles and multis indicates that the dynamical history that makes singles "singles" is unrelated to the formation of the radius valley. 0.501.002.004.008.0016.0RP [Earth]010203040Number of PlanetsKp < 14.2 SinglesKp < 14.2 Multis0.501.002.004.008.0016.0RP [Earth]0.00.20.40.60.81.0CDFp = 0.02Kp < 14.2 SinglesKp < 14.2 Multis0.501.002.004.008.0016.0RP [Earth]020406080Number of Planets2 tp.3 tp.4+ tp.All CKS multis0.501.002.004.008.0016.0RP [Earth]0.00.20.40.60.81.0CDFp = 0.772 tp.3 tp.4+ tp.All CKS multis1.001.191.411.682.002.382.833.364.00RP [Earth]0.00.20.40.60.81.0Fraction of PlanetsKp < 14.2 SinglesKp < 14.2 Multis1.001.191.411.682.002.382.833.364.00RP [Earth]0.00.20.40.60.81.0CDFp = 0.42Kp < 14.2 SinglesKp < 14.2 Multis 18 Figure 9. Top left: the distribution of orbital periods for the CKS Kp < 14.2 sub-Neptunes that are singles (Bs,4, red) and multis (Bm,4, blue). Top right: the same, but the cumulative distribution function (CDF). There is an excess of short-period singles (P < 3 days) among the sub-Neptune sized planets. Bottom left: the distribution of planet orbital period for all of the CKS-Gaia (black), and for the sub-samples with 2, 3, and 4+ transiting planets. Bottom right: the CDF. There are no significant differences in the distributions of orbital periods based on transiting planet multiplicity. If the excess of singles at P < 3 days were due to geometric effects, we would likely also see an excess of systems with 2 transiting planets as compared to 3, and 3 transiting planets as compared to 4+, due to geometric effects. The similarity of the orbital period distributions of multis with different transiting planet multiplicities suggests that some difference in orbital architecture might account for the excess of short-period singles. 1.003.1610.031.6100.316.1000Period [days]010203040Number of PlanetsKp < 14.2 SinglesRp < 4 ReKp < 14.2 MultisRp < 4 Re1.003.1610.031.6100.316.1000Period [days]0.00.20.40.60.81.0CDFp = 0.001Kp < 14.2 SinglesRp < 4 ReKp < 14.2 MultisRp < 4 Re1.003.1610.031.6100.316.1000Period [days]020406080Number of Planets2 tp.3 tp.4+ tp.All CKS multis1.003.1610.031.6100.316.1000Period [days]0.00.20.40.60.81.0CDFp = 1.02 tp.3 tp.4+ tp.All CKS multis Planets at short orbital periods are more likely to transit than planets at long orbital periods. Could the excess of short period singles be the result of a geometrical viewing effect that has nothing to do with astrophysics? If geometry alone is responsible for the excess of short pe- riod singles, we might expect to see a difference between the Ntp = 2, 3, and 4+ sub-samples in this period range. However, the bottom half of Figure 9 shows that the orbital distributions for multiple transiting planets are indistinguish- able for Ntp = 2, 3, and 4+. In other words, there are just as many planets with P < 3 days that belong to 4-transiting planet systems as there are planets that belong to 2 or 3-transiting planet systems. That the various multi-planet systems have very similar orbital period distri- butions, whereas the distribution of orbital pe- riods of singles is unique, suggests that geomet- rical bias alone is unlikely to account for the excess of singles, although a detailed suite of forward-modeling is necessary to demonstrate this claim. In principle, the excess of singles at short or- bital periods could possibly be related to more frequent false positives at short periods. How- ever, false positives are unlikely to mimic such a large number of planets considering the high purity of our sample. Alternatively, the hot super-Earths with P < 3 days might have a different dynamical his- tory from the longer-period sub-Neptunes, as suggested in Steffen & Coughlin (2016). The ultra-short periods planets (P < 1 day) with additional transiting planets have wider pe- riod ratios between the innermost pair of plan- ets than the farther-out planet pairs (Sanchis- Ojeda et al. 2014). In contrast, the majority of transiting planets in multi-planet systems have very regular orbital period spacing (Weiss et al. 2018). Thus, it is likely that a dynamical pro- cess is responsible for the inward migration of the innermost planet in multi-planet systems. 19 Planet-planet scattering (e.g., Chatterjee et al. 2008), secular chaos (Petrovich et al. 2018), and tidal inspiral (Lee & Chiang 2017) are theories that systematically bring the innermost planet of a multi-planet system to short orbital peri- ods. Why do the inward-moved small planets in our sample tend to be singles? The large period ra- tio between the innermost planet and the next planet out might be enough to explain why the short-period planets tend to be singles. Suppose the short-period singles really belong to nearly- coplanar multi-planet systems whose midplanes are slightly misaligned relative to our line of sight. In this case, the combination of the or- bital distance and inclination of the innermost planet might allow it to transit, while the other planets are too distant from the star to tran- sit. However, this explanation is inconsistent with the lack of observed short-period planets in multis. If a dynamical process were frequently moving the innermost planet closer to the star without disrupting the coplanarity of the sys- tem, we would expect to frequently detect the short-period (P < 3 days) sub-Neptunes among the multis. Therefore, in addition to a mecha- nism that moves the innermost planet closer to the star, a mechanism that increases the mu- tual inclinations of the planets is likely needed to explain why there are so many short-period singles and so few short-period multis among the sub-Neptunes. Recently, Dai et al. (2018) found that the mutual inclinations of planets in multis with a very short-period planet are larger than the mutual inclinations in multis with no transiting short-period planets. Ultimately, the orbital period distributions of the Bs singles, Bm multis, and the Ntp = 2, 3, and 4+ samples are all empirical distributions that should be reproduced by any model that aims to describe the underlying distributions of the planet multiplicities, orbital periods, and in- clinations. Fang & Margot (2012); Tremaine & 20 Dong (2012); Ballard & Johnson (2016); Gai- dos et al. (2016) and Zhu et al. (2018) all at- tempted to determine the underlying multiplic- ity and inclination distributions of the Kepler multis, but none of these attempts sought to re- produce the orbital period distributions for the different transiting planet multiplicities. Such an exercise would be valuable but is beyond the scope of this paper. A new tool presented in Mulders et al. (2018) is a promising step toward self-consistently modeling the various observed distributions of the singles and multis. 5. CONCLUSIONS In this paper, we explored how the physical properties of the CKS systems containing mul- tiple detected transiting planets (multis) com- pare to systems with just one detected tran- siting planet (singles). Although other studies have examined the relationships between stellar and/or planetary properties and planet multi- plicity, our study presents three advantages: (1) The CKS-Gaia dataset enables the largest, most accurate, and most precise comparison of the fundamental host star properties of the Kepler singles and multis so far. (2) As a result of strin- gent magnitude, detection threshold, and false positive cuts, our comparison of singles vs. mul- tis suffers from fewer observational biases than other studies. (3) In addition to comparing the properties of singles vs. multis, we compare the host star and planet properties as a function of the number of transiting planets. Our conclusions are as follows: 1. The distributions of stellar mass, metal- licity, and projected rotation velocity do not differ significantly for the singles and multis. The lack of a relationship be- tween stellar physical properties and ap- parent planet multiplicity suggests that any physical process that preferentially creates "singles" occurs late in planet for- mation and in a manner that is not related to the properties of the host star. Also, stellar properties are not particularly use- ful in predicting the number of transiting planets around a star. 2. Transiting planets of various multiplici- ties exhibit a valley in the radius distri- bution at ∼ 1.8 R⊕. The statistically in- distinguishable size distributions of small planets (Rp < 4R⊕) in singles and multis suggests that the acquisition of and sub- sequent evaporation of a volatile envelope around the planetary core is the same for the singles and multis. Because photo- evaporation happens within the first 100 Myr and is only effective at short orbital periods, the singles and multis likely ar- rive near their present orbital distances within the first 100 Myr. 3. For the sub-Neptune sized planets, there is a significant (p = 0.001) excess of short- period singles (P < 3 days) compared to multis. However, among the multis, the Ntp = 2, 3, and 4+ orbital period dis- tributions are the same, suggesting that geometrical bias alone is unlikely to ex- plain the excess of short-period singles. False positives are unlikely to mimic such a large number of planets considering the high purity of our sample. The excess of short-period planets could also be the hallmark of a late mechanism, such as planet-planet scattering or tidal migra- tion, that produces more singles than mul- tis at very short orbital periods. 4. Hot Jupiters are almost always single transiting planets that orbit high-mass and high-metallicity host stars, but these systems are intrinsically rare. Our main finding is that host star properties and planet radii for the majority of the sub- Neptune sized planets have no strong relation- ship with whether there are multiple transit- ing planets in the system. The similarity of the singles and multis suggests that they have a common origin. The majority of the singles with P > 3 days and Rp < 4R⊕ likely be- long to multi-planet systems with higher mu- tual inclinations than the CKS multis. Perhaps the singles are multi-planet systems that have undergone planet-planet scattering, resulting in systems of multiple planets in which only one planet transits. By contrast, the multis likely had dynamically quieter histories, as evidenced by their current low-entropy states. How might planet composition relate to mul- tiplicity? At face value, the observed radii and orbital periods of the singles and multis are in- consistent with the prediction in Dawson et al. (2016). That study predicted that planets in multi-planet systems should have preferentially volatile-rich compositions, whereas the singles should preferentially have rocky compositions. The idea behind the prediction was that the multi-planet systems form a little bit earlier than the singles, while the gas disk is a little bit denser, which contributes to both (1) eccentric- ity damping, resulting in more circular (hence, stable) orbits for the planets, and (2) more gas- rich planets. In contrast, we observe that both the singles and multis include significant pop- ulations of planets larger and smaller than the transition from rocky to volatile-rich planets at ∼ 1.8 R⊕. In other words, there is no evidence that the singles are preferentially rocky, or that the multis are preferentially gas-rich. Further- more, if there were an especially large popula- tion of rocky singles with 10 < P < 30 days and 1 R⊕ < Rp < 1.8 R⊕, many such planets would have been detected in the Kepler Mission and included in our sample. Perhaps the singles and multis generally form in compact multi-planet systems. These plan- ets can either form early while gas is abundant (forming volatile-rich planets) or later when there is less gas (forming rocky planets). In ei- 21 ther case, whatever small eccentricities the plan- ets have acquired will grow after the gas disk dissipates, sometimes leading to dynamical in- stability. For instance, Obertas et al. (2017) found that the Lyapunov time can vary by a couple orders of magnitude for compact multi- planet systems based on slight differences in the initial orbital conditions2. Thus, whether a multi-planet system becomes unstable on a timescale of gigayears might be primarily as- signed at birth. On the other hand, external influences such as passing stars might also play a role in dynamically disrupting initially copla- nar multis (Spalding & Batygin 2016). The sin- gles have higher eccentricities on average than the planets in multis (Xie et al. 2016; Van Eylen et al. 2018), suggesting that dynamical heating and perhaps instability play a role in the forma- tion of the singles. Single sub-Neptunes at very short orbital pe- riods likely have a different dynamical history than the multis. The singles with P < 3 days likely belong to multi-planet systems in which some combination of planet-planet scat- tering, secular chaos, and/or tidal inspiral has moved the innermost planet close to its star. Additional measurements, especially of the im- pact parameters, orbital obliquities, eccentrici- ties, and masses of the planets in both singles and multis with P < 3 days, will clarify which astrophysical process best explains the appar- ent excess of sub-Neptune sized singles at short orbital periods. We thank James Owen, Hilke Schlichting, Daniel Fabrycky, and the anonymous referee for helpful conversations and feedback. Most of the 2 Although it might not be obvious how the initial conditions of some compact multi-planet systems (but not others) lead to eventual instability, Tamayo et al. (2018) found that machine learning techniques are able to predict the outcomes of N-body simulations with high fidelity. 22 data presented here were determined directly from observations at the W. M. Keck Observa- tory, which is operated as a scientific partner- ship among the California Institute of Technol- ogy, the University of California, and NASA. We are grateful to the time assignment com- mittees of the University of Hawaii, the Uni- versity of California, the California Institute of Technology, and NASA for their generous allo- cations of observing time that enabled this large project. Kepler was competitively selected as the tenth NASA Discovery mission. Funding for this mission is provided by the NASA Sci- ence Mission Directorate. LMW acknowledges support from the Parrent Fellowship, the Trot- tier Family, and the Levy Family. AWH ac- knowledges NASA grant NNX12AJ23G. EAP acknowledges support from Hubble Fellowship grant HST-HF2-51365.001-A awarded by the Space Telescope Science Institute, which is op- erated by the Association of Universities for Re- search in Astronomy, Inc. for NASA under con- tract NAS 5-26555. LH acknowledges National Science Foundation grant AST-1009810. ES is supported by a post-graduate scholarship from the Natural Sciences and Engineering Research Council of Canada. PAC acknowledges Na- tional Science Foundation grant AST-1109612. Finally, the authors wish to recognize and ac- knowledge the very significant cultural role and reverence that the summit of Maunakea has al- ways had within the indigenous Hawaiian com- munity. We are most fortunate to have the opportunity to conduct observations from this mountain. Facilities: Keck:I (HIRES), Kepler Software: python2.7, scipy, pandas REFERENCES Ballard, S., & Johnson, J. A. 2016, ApJ, 816, 66, doi: 10.3847/0004-637X/816/2/66 Becker, J. C., Vanderburg, A., Adams, F. C., Rappaport, S. A., & Schwengeler, H. M. 2015, ApJL, 812, L18, doi: 10.1088/2041-8205/812/2/L18 Borucki, W. J., Koch, D. G., Brown, T. M., et al. 2010, ApJL, 713, L126, doi: 10.1088/2041-8205/713/2/L126 Brown, T. M., Latham, D. W., Everett, M. E., & Esquerdo, G. a. 2011, The Astronomical Journal, 142, 112, doi: 10.1088/0004-6256/142/4/112 Butler, R. P., Marcy, G. W., Fischer, D. A., et al. 1999, ApJ, 526, 916, doi: 10.1086/308035 Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686, 580, doi: 10.1086/590227 Dai, F., Masuda, K., & Winn, J. N. 2018, ArXiv e-prints, arXiv:1808.08475. https://arxiv.org/abs/1808.08475 Dawson, R. I., Lee, E. J., & Chiang, E. 2016, ApJ, 822, 54, doi: 10.3847/0004-637X/822/1/54 Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., et al. 2014, ApJ, 790, 146, doi: 10.1088/0004-637X/790/2/146 Fang, J., & Margot, J.-L. 2012, ApJ, 761, 92, doi: 10.1088/0004-637X/761/2/92 Fischer, D. A., & Valenti, J. 2005, ApJ, 622, 1102, doi: 10.1086/428383 Ford, E. B., Rowe, J. F., Fabrycky, D. C., et al. 2011, ApJS, 197, 2, doi: 10.1088/0067-0049/197/1/2 Fressin, F., Torres, G., Charbonneau, D., et al. 2013, The Astrophysical Journal, 766, 81 Fulton, B. J., Petigura, E. A., Howard, A. W., et al. 2017, AJ, 154, 109, doi: 10.3847/1538-3881/aa80eb Furlan, E., Ciardi, D. R., Everett, M. E., et al. 2017, AJ, 153, 71, doi: 10.3847/1538-3881/153/2/71 Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, ArXiv e-prints. https://arxiv.org/abs/1804.09365 Gaidos, E., Mann, A. W., Kraus, A. L., & Ireland, M. 2016, MNRAS, 457, 2877, doi: 10.1093/mnras/stw097 Ginzburg, S., Schlichting, H. E., & Sari, R. 2018, MNRAS, 476, 759, doi: 10.1093/mnras/sty290 Hansen, B. M. S., & Murray, N. 2013, The Astrophysical Journal, 775, 53, doi: 10.1088/0004-637X/775/1/53 Johnson, J. A., Butler, R. P., Marcy, G. W., et al. 2007, The Astrophysical Journal, 670, 833 Johnson, J. A., Petigura, E. A., Fulton, B. J., et al. 2017, AJ, 154, 108, doi: 10.3847/1538-3881/aa80e7 Latham, D. W., Rowe, J. F., Quinn, S. N., et al. 2011, ApJL, 732, L24, doi: 10.1088/2041-8205/732/2/L24 Lee, E. J., & Chiang, E. 2017, ApJ, 842, 40, doi: 10.3847/1538-4357/aa6fb3 Lissauer, J. J., Ragozzine, D., Fabrycky, D. C., et al. 2011, ApJS, 197, 8, doi: 10.1088/0067-0049/197/1/8 Lissauer, J. J., Marcy, G. W., Rowe, J. F., et al. 2012, ApJ, 750, 112, doi: 10.1088/0004-637X/750/2/112 Lissauer, J. J., Marcy, G. W., Bryson, S. T., et al. 2014, ApJ, 784, 44, doi: 10.1088/0004-637X/784/1/44 Mulders, G. D., Pascucci, I., Apai, D., & Ciesla, F. J. 2018, AJ, 156, 24, doi: 10.3847/1538-3881/aac5ea Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015, The Astrophysical Journal Supplement Series, 217, 31, doi: 10.1088/0067-0049/217/2/31 Munoz Romero, C. E., & Kempton, E. M.-R. 2018, AJ, 155, 134, doi: 10.3847/1538-3881/aaab5e Obertas, A., Van Laerhoven, C., & Tamayo, D. 2017, Icarus, 293, 52, doi: 10.1016/j.icarus.2017.04.010 Owen, J. E., & Wu, Y. 2017, ArXiv e-prints. https://arxiv.org/abs/1705.10810 Petigura, E. A. 2015, PhD thesis, University of California, Berkeley Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Proceedings of the National Academy of Sciences, 110, 19273, doi: 10.1073/pnas.1319909110 Petigura, E. A., Howard, A. W., Marcy, G. W., Petigura, E. A., Marcy, G. W., Winn, J. N., et al. et al. 2017, AJ, 154, 107, doi: 10.3847/1538-3881/aa80de 2018, AJ, 155, 89, doi: 10.3847/1538-3881/aaa54c 23 Petrovich, C., Deibert, E., & Wu, Y. 2018, ArXiv e-prints, arXiv:1804.05065. https://arxiv.org/abs/1804.05065 Rogers, L. a. 2015, ApJ, 801, 41, doi: 10.1088/0004-637X/801/1/41 Rowe, J. F., Bryson, S. T., Marcy, G. W., et al. 2014, ApJ, 784, 45, doi: 10.1088/0004-637X/784/1/45 Sanchis-Ojeda, R., Rappaport, S., Winn, J. N., et al. 2014, ApJ, 787, 47, doi: 10.1088/0004-637X/787/1/47 Santerne, A., Moutou, C., Tsantaki, M., et al. 2016, A&A, 587, A64, doi: 10.1051/0004-6361/201527329 Spalding, C., & Batygin, K. 2016, ApJ, 830, 5, doi: 10.3847/0004-637X/830/1/5 Steffen, J. H., & Coughlin, J. L. 2016, Proceedings of the National Academy of Science, 113, 12023, doi: 10.1073/pnas.1606658113 Steffen, J. H., Ragozzine, D., Fabrycky, D. C., et al. 2012, Proceedings of the National Academy of Science, 109, 7982, doi: 10.1073/pnas.1120970109 Tamayo, D., Hadden, S., Hussain, N., et al. 2018, in AAS/Division of Dynamical Astronomy Meeting, 201.02 Tremaine, S., & Dong, S. 2012, AJ, 143, 94, doi: 10.1088/0004-6256/143/4/94 Van Eylen, V., Albrecht, S., Huang, X., et al. 2018, ArXiv e-prints, arXiv:1807.00549. https://arxiv.org/abs/1807.00549 Weiss, L. M., & Marcy, G. W. 2014, ApJ, 783, L6, doi: 10.1088/2041-8205/783/1/L6 Weiss, L. M., Marcy, G. W., Petigura, E. A., et al. 2018, AJ, 155, 48, doi: 10.3847/1538-3881/aa9ff6 Wright, J. T., Upadhyay, S., Marcy, G. W., et al. 2009, The Astrophysical Journal, 693, 1084, doi: 10.1088/0004-637X/693/2/1084 Xie, J.-W., Dong, S., Zhu, Z., et al. 2016, Proceedings of the National Academy of Science, 113, 11431, doi: 10.1073/pnas.1604692113 Zhu, W., Petrovich, C., Wu, Y., Dong, S., & Xie, J. 2018, ArXiv e-prints. https://arxiv.org/abs/1802.09526
1309.0508
1
1309
2013-09-02T20:00:02
Characterising thermal sweeping: a rapid disc dispersal mechanism
[ "astro-ph.EP", "astro-ph.SR" ]
(Abridged) We consider the properties of protoplanetary discs that are undergoing inside-out clearing by photoevaporation. In particular, we aim to characterise the conditions under which a protoplanetary disc may undergo `thermal sweeping', a rapid (< 1e4 years) disc destruction mechanism proposed to occur when a clearing disc reaches sufficiently low surface density at its inner edge and where the disc is unstable to runaway penetration by the X-rays. We use a large suite of 1D radiation-hydrodynamic simulations to probe the observable parameter space, which is unfeasible in higher dimensions. These models allow us to determine the surface density at which thermal sweeping will take over the disc's evolution and to evaluate this critical surface density as a function of X-ray luminosity, stellar mass and inner hole radius. We find that this critical surface density scales linearly with X-ray luminosity, increases with inner hole radius and decreases with stellar mass and we develop an analytic model that reproduces these results. This surface density criterion is then used to determine the evolutionary state of protoplanetary discs at the point that they become unstable to destruction by thermal sweeping. We find that transition discs created by photoevaporation will undergo thermal sweeping when their inner holes reach 20-40 AU, implying that transition discs with large holes and no accretion (which were previously a predicted outcome of the later stages of all flavours of photoevaporation model) will not form. We emphasise that the surface density criteria that we have developed apply to all situations where the disc develops an inner hole that is optically thin to X-rays. It thus applies not only to the case of holes originally created by photoevaporation but also to holes formed, for example, by the tidal influence of planets.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2002) Printed 8 October 2018 (MN LATEX style file v2.2) Characterising thermal sweeping: a rapid disc dispersal mechanism James E. Owen1⋆, Mathias Hudoba de Badyn1,2, Cathie J. Clarke3 & Luke Robins3 1Canadian Institute for Theoretical Astrophysics, 60 St. George Street, Toronto, M5S 3H8, Canada. 2Department of Physics and Astronomy, University of British Columbia, Vancouver, V6T 1Z1 Canada. 3Institute of Astronomy, Madingley Rd, Cambridge, CB3 0HA, UK. 8 October 2018 ABSTRACT We consider the properties of protoplanetary discs that are undergoing inside-out clearing by photoevaporation. In particular, we aim to characterise the conditions under which a protoplanetary disc may undergo 'thermal sweeping', a rapid (. 104 years) disc destruction mechanism proposed to occur when a clearing disc reaches sufficiently low surface density at its inner edge and where the disc is unstable to runaway penetration by the X-rays. We use a large suite of 1D radiation-hydrodynamic simulations to probe the observable parameter space, which is unfeasible in higher dimensions. These models allow us to determine the surface density at which thermal sweeping will take over the disc's evolution and to evaluate this critical surface density as a function of X-ray luminosity, stellar mass and inner hole radius. We find that this critical surface density scales linearly with X-ray luminosity, increases with inner hole radius and decreases with stellar mass and we develop an analytic model that reproduces these results. This surface density criterion is then used to determine the evolutionary state of protoplanetary discs at the point that they become unstable to destruction by thermal sweeping. We find that transition discs created by photoevaporation will undergo ther- mal sweeping when their inner holes reach 20 − 40 AU, implying that transition discs with large holes and no accretion (which were previously a predicted outcome of the later stages of all flavours of photoevaporation model) will not form. Thermal sweep- ing thus avoids the production of large numbers of large, non-accreting holes (which are not observed) and implies that the majority of holes created by photoevaporation should still be accreting. We emphasise that the surface density criteria that we have developed apply to all situations where the disc develops an inner hole that is optically thin to X-rays. It thus applies not only to the case of holes originally created by photoevaporation but also to holes formed, for example, by the tidal influence of planets. Key words: planetary systems: protoplanetary discs - stars: pre-main-sequence. 3 1 0 2 p e S 2 . ] P E h p - o r t s a [ 1 v 8 0 5 0 . 9 0 3 1 : v i X r a 1 INTRODUCTION Understanding the dispersal of protoplanetary discs remains a key problem in star and planet formation. In particular, the time at which the disc disperses sets the time in which (gas) planets must form. Additionally, the method and time- scale on which disc clearing operates strongly influences the dust and gas physics in the disc, as well as the evolution of currently forming planets. Observationally, optically thick primordial discs are known to survive for several Myr before being dispersed ⋆ E-mail: [email protected] c(cid:13) 2002 RAS in both dust (Haisch et al. 2001; Hernandez et al. 2007; Mamajek 2009) and gas (Kennedy & Kenyon 2009), leav- ing behind a pre-main sequence star possibly surrounded by an optically thin debris disk, containing second genera- tion dust particles (Wyatt 2008). A small fraction of young stars show evidence for a cleared gap or hole in their dust discs, indicated by a lack of opacity at NIR wavelengths, but values comparable with an optically thick disc at MIR wave- lengths. These 'transition' discs (Strom et al. 1989; Sturski et al. 1990; Calvet et al. 2002) have been interpreted as discs caught at a stage which is intermediate between disc bearing and disc-less. Furthermore, the lack of objects observed in regions of parameter space that represent discs that are no 2 Owen et al. longer optically thick at IR wavelengths indicates that the time-scale for disc dispersal is short compared to the disc's lifetime (Kenyon & Hartmann 1995). Comparing the ratio of observed transition discs to primordial discs indicates that the clearing time-scale for the inner disc in roughly 10% of the disc's lifetime (Luhman et al. 2010; Ercolano et al. 2011; Koepferl et al. 2013). Furthermore, the lack of objects which show no NIR or MIR excess but show excesses at FIR or sub-mm/mm wavelengths (Duvert et al. 2000; Andrews & Williams 2005;2007; Cieza et al. 2013) indicate that this inner disc clearing is correlated with outer disc clearing. While 'transition' discs have several possible origins such as grain growth (Dullemond & Dominik 2005; Birn- stiel et al. 2012), tidal truncations by planets (Calvet et al. 2005; Rice et al. 2006; Zhu et al. 2012, Clarke & Owen 2013) or photoevaporation (Clarke et al. 2001), it is planet forma- tion (Armitage & Hansen 1999) or photoevaporation that offer the best scenarios for disc dispersal. While no quali- tative or quantitative predictions can be made about how planet formation clears the disc, one would assume it slows as it proceeds through the disc due to the longer planet formation time and large mass-reservoir to remove. Photoe- vaportion driven by X-rays from the central star can explain a large fraction of the observed transition discs (Owen et al. 2011, 2012) but not the full sample; in particular transition discs with large holes and high accretion rates (Espaillat et al. 2010; Andrews et al. 2011) are inconsistent with a pho- toevaporative origin. However, this category of 'transition' disc may have a different origin and not even be associated with the disc clearing process at all (Owen & Clarke 2012; Clarke & Owen 2013; Nayakshin 2013) being instead per- haps a relatively rare but long-lived phenomenon. Photoevaporation can explain the population of ob- served transition discs with small holes (Rhole < 20 AU) and modest accretion rates ( M∗ < 10−8 M⊙ yr−1). However, the same model predicts a comparable sample of non-accreting transition discs with large holes; termed 'relic discs' by Owen et al. (2011), and a similar population may be expected if planet formation is the dominant dispersal mechanism. While the exact population of transition discs with large holes is yet to be quantified, the lack of Weak T Tauri stars (WTTs) with sub-mm detections (Duvert et al. 2000; An- drews & Williams 2005;2007, Mathews et al. 2012) or Her- schel detections (Cieza et al. 2013) is inconsistent with a sizeable (comparable to the population of 'transition' discs) population of relic discs. The 1D viscous calculation of disc dispersal typically employed in disc evolution calculations (Clarke et al. 2001; Alexander et al. 2006; Alexander & Armitage 2009; Owen et al. 2010,2011) neglect the fact that the disc may be unsta- ble to dynamical clearing, where the disc's mid-plane can be entirely penetrated by the X-rays. Owen et al. (2012) pre- sented simulations of a process they termed 'thermal sweep- ing' whereby a disc with an optically thin inner hole is dy- namically cleared (on a time-scale ∼ 103 years) once the surface density at the disc's inner edge falls to a sufficiently low value. In the low mass discs simulated by Owen et al, this process set in at radii ∼ 10− 20 AU. Owen et al. (2012) argued that this process prevents the production of a long- lived population of low mass discs with large cavities ('relic discs') and that it shifts the predicted populations of transi- Figure 1. Schematic diagram of the flow/disc structure resulting from X-ray irradiation of a disc with a cleared inner hole. tion discs created through photoevaporation to those domi- nated by small holes and low accretion rates. In addition, Owen et al. (2012) presented a back of the envelope calculation to estimate the surface density ΣT S at which thermal sweeping begins; however, such a calcu- lation was only compared to two multi-dimension radiation hydrodynamic calculations. A parameter span using multi- dimensional calculations is not currently computationally feasible. Thus, in this paper, we build on the framework for thermal sweeping outlined in Owen et al. (2012) in order to determine the critical surface density for thermal sweep- ing to proceed by performing a large set of 1D radiation- hydrodynamic calculations. In Section 2 we describe the thermal sweeping mechanism and summarise the back of the envelope calculation performed by Owen et al. (2012). Section 3 outlines the numerical method used for suite of 1D radiation hydrodynamic calculations and present our find- ings in Section 4. In Section 5 we discuss the implications for thermal sweeping on disc evolution and discuss an improved model for determining the critical density at which thermal sweeping proceeds and finally in Section 6 we summarise our findings. 2 BASICS OF THERMAL SWEEPING In this section we describe the theoretical basis for how X- ray irradiation of a disc with an inner hole leads to rapid, dynamical disc dispersal. We also present the back of the envelope calculation presented by Owen et al. (2012) as well as pointing out the assumptions that may break down. 2.1 Understanding thermal sweeping In order to explain the origin of thermal sweeping, we show in Figure 1 a schematic depiction of the flow structure that results from X-ray irradiation of a disc with an inner hole, based on the simulations of Owen et al. (2010,2011,2012). The X-rays heat up the inner disc and drive a sub-sonic wind radially inwards. The simulations of Owen et al. (2012) showed that the gas becomes significantly decelerated by centrifugal effects 1 at a radius of roughly Rhole/2 and at this 1 Recall that viscous effects are negligible in the dynamical wind flow and thus the gas in this region conserves its specific angular momentum c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? point is deflected vertically by pressure gradients, attaining a sonic transition at a height z ∼ R. Owen et al (2012) showed that the mass flux in the wind is set at the sonic point and that the structure of the radially inflowing thermal wind region (see Figure 1) can be understood in terms of a one-dimensional, angular mo- mentum conserving flow subject to this boundary condition on the mass flux. The total radial column density in this thermal wind region is given by: Mw 2πµvR(H/R) N ∼ nRhole = = 1021 cm−2(cid:18) Mw 10−8 M⊙ yr−1(cid:19)(cid:16) v 106 cm s−1(cid:17)−1 (1) ×(cid:18) R 10 AU(cid:19)−1(cid:18) H/R 0.1 (cid:19)−1 Mw the where N is the column, n is the number density, mass-loss rate, µ the mean particle mass, v the gas veloc- ity and H the scale height. This is significantly lower than the column required to absorb the X-rays (∼ 1022 cm−2 - Ercolano et al. 2008;2009). Thus, in this flow topology the material that dominates the absorption of the X-rays is not actually in the thermal wind, but is still bound to the disc, essentially in hydrostatic equilibrium with temper- atures < 1000 K. Owen et al. (2012) argued that for this bound X-ray heated region to remain in dynamical equilibrium it must have a radial pressure scale (∆ ≡ dR/d log P ) smaller than its vertical pressure scale (dz/d log P ≡ H = cs/Ω). Since in order to obtain dynamical equilibrium, the disc must be able to adjust itself radially to changes in the vertical structure on a time-scale considerably shorter than the dynamical time- scale of the disc (∼ 1/Ω). In situations where ∆ ∼ H (and where the time-scales for attaining equilibrium in the radial and vertical directions are comparable), this bound X-ray heated region is unstable to progressive penetration by the X-rays as follows. In the case of a small vertical expansion of the flow, the disc cannot re-gain equilibrium in the radial direction on the vertical expansion time-scale and so the ra- dial column density at the mid-plane (and the local volume density) are reduced. Consequently, the temperature of the X-ray heated flow rises and the vertical expansion is accel- erated. The further reduction in mid-plane column density results in further penetration of the X-rays and the process of vertical evacuation of the disc is repeated for a fresh layer of previously unheated disc material. The key difference be- tween this behaviour and 'normal' photoevaporative clearing from a holed disc is that vertical expansion leads to runaway heating of the disc inner rim. 2.2 Column limited estimate for thermal sweeping Here we repeat the simple analytical argument presented by Owen et al. (2012), but emphasise the assumptions made in its derivation. Since the X-ray bound heated region must be in approximate pressure equilibrium with the passively heated (by dust) disc at large radii one can write: kbnX TX = P dust h (2) c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Characterising Thermal Sweeping 3 where kb is Boltzmann's constant, nX & TX is the number density and temperature in the bound X-ray heated region and P dust is the gas pressure in the passively heated disc. Writing ∆ as approximately NX /nX , and using the fact that thermal sweeping begins when ∆ = H, we can recast Equa- tion 2 in terms as a criterion for thermal sweeping as: h P dust h 6 NX kbTX H (3) where NX is the radial column density corresponding to complete absorption of the X-rays (∼ 1022 cm−2). In the case of a disc in vertical hydrostatic equilibrium this can be re-cast in terms of a critical surface density for the inner rim of the dust-heated disc: ΣT S = 0.4 g cm−2(cid:18) TX 400 K(cid:19)1/2(cid:18) Tdust 20 K(cid:19)−1/2 (4) Given Tdust ∝ R−1/2, this resulted in a weak R1/4 scaling for the critical surface density with radius, with no explicit dependence on stellar mass or X-ray luminosity. However, this derivation presented by Owen et al. (2012) makes two strong assumptions: i) that the bound X- ray heated region always represents a column density of 1022 cm−2; ii) the surface density at the transition between the X-ray heated disc and passively heated disc is representative of the peak surface density of the disc's profile. Assumption i) may break down if the gas density is so high that X-ray heating is insufficient to heat the gas above the dust temperature even if it is optically thin to the X- rays. This typically occurs at ionization parameters ξmin . 10−7 erg cm s−1 (where ξ = LX /nr2). In the case that the boundary of the X-ray heated region is set by the criterion ξ = ξmin, we can estimate the column density in the bound X-ray heated region as N = n∆. At the point of onset of thermal sweeping this can be written N = (H/R)nR which (from the definition of the ionization parameter) becomes: N = 1022 cm−2(cid:18) H/R 10 AU(cid:19)−1(cid:18) ×(cid:18) R 0.1 (cid:19)(cid:18) LX 1030 erg s−1(cid:19) 10−7 erg cm s−1(cid:19)−1 ξmin (5) Therefore, since we expect thermal sweeping to take place when the hole radius is large (Rhole & 10 AU), then we need to consider cases where the total column density in the bound X-ray heated region is < 1022 cm−2. This means that Equation 4 will not represent how the critical surface density for thermal sweeping varies in certain regions of the param- eter space. In particular, if the criterion in some regions of parameter space is set by a minimum ionization parameter rather than total column density then it should depend on the X-ray luminosity as well as inner hole radius. In this paper we assume without further discussion that ∆ = H is a necessary and sufficient condition for initiat- ing thermal sweeping (an assumption that we will revisit in future two-dimensional studies motivated by the results of the present paper.) Here we use 1D radiation-hydrodynamic models to perform a parameter space study, calculating how the critical surface density (ΣT S - defined as the peak sur- face density at the inner edge of the dust-heated disc for which ∆ = H) varies with stellar mass, X-ray luminosity and inner hole radius. The advantage of a one-dimensional treatment is that it is possible to perform a large suite of 4 Owen et al. ] 3 − m c g [ y t i s n e D 10−12 10−13 10−14 10−15 10−16 1000 100 ] K [ e r u t a r e p m e T 7 8 9 10 Radius [AU] 20 10 7 8 9 10 Radius [AU] 20 Figure 2. Density structure for a simulation of a disc with an inner hole around a 0.1 M⊙ star with an X-ray luminosity of 2 × 1030 erg s−1. We note the outer boundary is not shown and is at a radius of 38 AU in this calculation. high resolution simulations that are beyond the scope of 2D simulations and that can inform the choice of parameters and resolution requirements of future 2D studies. The jus- tification for a one dimensional treatment is that - prior to the onset of thermal sweeping - the flow is to a good approx- imation one dimensional in the vicinity of the inner edge. 3 NUMERICAL METHOD In order to calculate the inner disc structures we must make use of numerical hydrodynamic calculations. We use the zeus code (Stone et al. 1992; Hayes et al. 2006), which has been modified by Owen et al. (2010;2012) to include the heating of circumstellar material by stellar X-ray photons. We use this modified code to calculate the mid-plane inner disc structure; therefore, we run the code radially in one- dimension (although we also evolve the azimuthal velocity2). In the absence of X-ray heating the disc's temperature is set to the mid-plane temperature expected for a passively irra- diated protoplanetary disc (e.g. Chiang & Goldreich 1997; D'alessio et al. 2001): Tmid = max"T1AU(cid:18) R 1 AU(cid:19)−1/2 , 10 K# (6) where T1AU is the mid-plane temperature at 1 AU, and we adopt typical values of 100 K for a 0.7 M⊙ star and 50K for a 0.1 M⊙ star (D'alessio et al. 2001). As initial conditions we adopt a Σ ∝ R−1 profile, where in order to convert between surface density and mid-plane gas density we assume the disc is vertically isothermal and use: Σ(R) = √2πH(R)ρ(R) (7) This mid-plane density profile is setup in hydrostatic equilib- rium including the necessary modification to the Keplerian rotation due to the radial pressure gradient. In order for the simulations to obtain stable configura- tion the boundary conditions require careful consideration. The outer boundary condition is implemented in order to maintain hydrostatic equilibrium across the outer boundary 2 Often referred to as 1.5D simulations. Figure 3. Temperature structure for the same simulation shown in Figure 2. throughout the simulation, so sub-sonic radial flows are not allowed to interact with the boundary. Explicitly we do this by maintaining a Σ ∝ R−1 power law across the boundary and adjust the azimuthal velocity in the boundary cells in order to maintain this equilibrium. In X-ray heated winds, the mass-flux is determined explicitly by the conditions at the sonic surface (Owen et al. 2012; Owen & Jackson 2012). It is not possible to follow the flow to the sonic point in a one dimensional simulation since this occurs at a height Z ∼ R above the disc. Therefore we adjust the inner bound- ary condition in order to match the flow properties (density, temperature, velocity) onto the results of an on-axis stream- line taking from the multi-dimensional inner hole simula- tions performed by Owen et al. (2012). Such a method is sufficient provided the inner boundary lies between the cen- trifugal radius and inner hole radius (so the problem remains remains essentially 1D, see Figure 1). The radial grid is non-uniform, spanning [Ri, 6Ri] - where Ri in the radius of the inner boundary - with a high resolution of 200 cells; typically this gives ∼ 10 − 15 cells per radial pressure scale height in the inner hole region. In general, the structures are spatially resolved with a lower resolution, typically 100 cells; however, the higher resolu- tion reduces the scatter in properties calculated (particu- larly ∆). Under this set-up the inner disc is then irradiated, with the X-rays driving a flow off the inner disc with a mass- flux specified by the conditions at the inner boundary; this flow slowly erodes material from the disc resulting in an in- ner hole radius that increases slowly with simulation time. The evolution is slow enough that the structure maintains a quasi-dynamical equilibrium at each time-step. We stop the simulation once the inner hole moves to such a radius that the inner boundary is no longer outside the centrifugal radius, and the 1D approximation is no-longer appropriate. In Figure 2 & 3 we show an example of the quasi-steady state flow structure obtained from one of the simulations, for a disc around a 0.1 M⊙ star with an X-ray luminosity of 2 × 1030 erg s−1. Using 1D simulations allows us to probe a large range of the observable parameter space of disc models at high resolution. We perform roughly 5000 simulations covering a range of initial surface densities and radial ranges for two stellar masses. For the 0.1 M⊙ star we consider X-ray lumi- c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Characterising Thermal Sweeping 5 0.2 0.15 0.1 0.05 0 ) / H ∆ ( g o l −0.05 0 1 −0.1 −0.15 −0.2 −0.25 −1 ] 2 − m c g [ ) y t i s n e D e c a f r u S ( 0 1 g o 1.5 1 0.5 0 −0.5 −1 −1.5 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 l −2 0.8 1.8 2 1 log 10 1.2 1.4 1.6 (Inner Hole Radius) [AU] (Peak Surface Density) [g cm−3] log 10 Figure 4. The variation in the ratio of ∆/H with peak surface density for a disc with an inner hole at ∼ 10 AU around a 0.7 M⊙ star and X-ray luminosity of 2×1030 erg s−1. The points represent individual values computed at a given time in the simulation, where the scatter is indicative of the resolution. The dashed line shows a polynomial fit to the points. Points which sit far from the line occur at early time-steps where the disc has not reached quasi-dynamical equilbrium. nosities3 of 2×1029 and 2×1030 erg s−1 and hole sizes in the range ∼ 1− 20. For the 0.7 M⊙ star we consider X-ray lumi- nosities of 2× 1029, 2× 1030 and 2× 1031 erg s−1 with inner hole sizes in the range ∼ 10− 70 AU. For each simulated ra- dial range we vary the initial surface density normalisation at intervals of 0.15dex over the range expected in clearing discs. These simulations are then used to compute how the ratio of ∆/H varies with surface density, thus extracting the critical surface density for thermal sweeping and how it varies varies with stellar mass, X-ray luminosity and inner hole radius. 4 RESULTS Our suite of simulations cover a large range of parameter space and provide the mid-plane structure of discs with large inner holes. For a given inner hole radius and surface density normalisation in the disc we use the simulation to solve for the ratio of ∆ to H and can hence determine how this ratio varies as a function of the maximum surface density in the dust heated disc. An example of how ∆/H varies with the peak surface density in a disc is shown in Figure 4. Figure 4 demonstrates how we use the simulation data to calculate ΣT S . We use the simulations to regularly store the current flow state (density, temperature, velocity); thus for each data dump we compute the ratio of ∆/H, the cur- rent hole radius and the peak surface density of the disc. Then for each radial range we string together all these val- ues in order to produce a plot of how this ratio varies with 3 Although low-mass stars may have lower X-ray luminosities, the simulated range is sufficient to calibrate the model discussed in Section 5, which can be applied to the full range of X-ray luminosities. c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Figure 5. Critical surface density for thermal sweeping (∆ = H) shown as a function of inner hole radius around a 0.7 M⊙ mass star for three X-ray luminosities: 2 × 1031 erg s−1 (triangles), 2 × 1030 erg s−1 (circles) and 2 × 1029 erg s−1 (squares). The error bars show the uncertainity in the calculation of the criti- cal surface density introduced by the finite grid resolution (see text and Figure 4).We also show a Σ ∝ R1/4 fit (perfomed at 10 AU) as the dotted line, to represent the column-limited model (Equation 4), which clearly fails to reproduce the simulations. the peak surface density in the disc. These data points nec- essarily have some scatter due to the finite grid resolution (around 0.05dex for our resolution), and therefore we fit a polynomial to this data series (typically a quadratic). We then use the fitted polynomial to compute the peak surface density in the disc at which ∆/H = 1 and the inner hole radius at which this occurs. Then we can combine the simu- lations at different radial ranges to see how ΣT S varies with inner hole radius, X-ray luminosity and stellar mass. 4.1 Variation with inner hole radius & stellar mass The resulting variation in the critical surface density for thermal sweeping is shown as a function of inner hole radius and X-ray luminosity in Figure 5 & 6 for discs around 0.7 & 0.1 M⊙ solar mass stars respectively. The finite grid res- olution that results in the scatter in Figure 4 gives rise to an uncertainty in the surface density of ∼ ±0.05 dex which we represent as error bars in the Figures. We emphasise this scatter does not mean the simulations are not spatially re- solved, in-fact the simulations are well resolved (we perform several resolution tests and note the simulations are resolved with ∼100 cells, we use 200 cells) and the resolution has been chosen to minimise this uncertainty while allowing a computationally feasible parameter study. The general trends of the critical surface density with inner hole radius and X-ray luminosity are characterised in Figure 5. In general we find the surface density increases with radius as well as X-ray luminosity. At the highest X- ray luminosity of 2×1031 erg s−1, we find the surface density is roughly constant at small radius (Rhole < 10 AU). This range represents the column limited case discussed by Owen et al. (2012) (equation 4), where the size of the X-ray heated region is purely set by the absorption of the X-rays and ∆ is simply given by ∼ NX /nX . However, as we will discuss further in Section 5, over most of parameter space the extent 6 Owen et al. ] 2 − m c g [ ) y t i s n e D e c a f r u S ( 0 1 g o 0.6 0.4 0.2 0 −0.2 −0.4 −0.6 −0.8 l −1 0.4 0.6 0.8 log (Inner Hole Radius) [AU] 10 1 ] 2 − m c [ y t i s n e D n m u o C l ] 1 − s m c 1022 1021 1 10 Inner Hole Radius [AU] g r e [ r e t e m a r a P n o i t i a z n o I 10−6 10−7 100 Figure 6. Same as Figure 5 but shown for a 0.1 M⊙ mass star of the X-ray heated region is set by a minimum ionization parameter, rather than an optical depth cut-off. 5 DISCUSSION AND OBSERVATIONAL IMPLICATIONS In the previous section we have shown how the critical sur- face density for thermal sweeping varies with stellar mass, inner hole radius and X-ray luminosity. As expected we find that thermal sweeping will proceed once the remaining mass in the disc is low . 1 MJup. However, we do not find the shal- low R1/4 dependence expected from the Owen et al. (2012) model, but a steeper variation with inner hole radius. Fur- thermore, we find an explicit, roughly linear, dependence on the X-ray luminosity, not predicted by Owen et al. (2012). In Section 2 we outlined the two assumptions that underpinned the previous calculation: that the total column density in the bound X-ray heated layer is ∼ 1022 cm−2, and that the sur- face density at the inner edge of the passively heated disc may not be representative of the peak surface density in the disc. We find that while the second clause certainly plays a role, it is not the major reason why the simulations do not show the expected dependence. It is the first reason: that the total column density in the bound X-ray heated layer is not always 1022 cm−2 and in the majority of realistic cases is lower than this value and a function of radius. For example we show in Figure 7 how the total column density and the ionization parameter at the transition from X-ray bound to dust heated layer varies with inner hole ra- dius for the case of a disc around a 0.7 M⊙ star irradiated at a high X-ray luminosity of 2 × 1031 erg s−1. This Fig- ure clearly shows that, at about 10 AU the X-ray heated region makes a transition from being column limited to be- ing ionization parameter limited, or more correctly X-ray temperature limited. 5.1 An improved model for thermal sweeping We can now use the above result: that the properties of the warm X-ray heated bound portion of the disc are set by a minimum ionization parameter rather than total column condition. Since the gas velocity in the bound portion of the disc (both dust and X-ray heated) is extremely sub-sonic, we Figure 7. The column density (solid line) and ionization param- eter (dashed line) in the X-ray heated bound region, shown as a function of inner hole radius for a disc around a 0.7 M⊙ star irradiated at an X-ray luminosity of 2 × 1030 erg s−1. The points show the radius of individual simulations. can treat these regions of the disc as being approximately hydrostatic. Inside a radius Rc (i.e. radius in the disc at which the source of material for the wind is in centrifugal balance) the specific angular momentum is constant and the effective gravity in the radial direction is given by: geff = − = Ω2δ(R) GM∗ R2 (cid:18)1 − Rc R (cid:19) (8) where δ(R) = Rc − R and Ω is the local Keplerian velocity. Now using the definition that ∆ is the radial pressure scale length in the X-ray heated gas we see that: ∆ ≡ c2 X /geff = H 2 X /δ (9) where cX is the sound speed at the outer edge of the X-ray heated disc and we have applied the condition for hydro- static equilibrium normal to the disc plane, i.e. cX = HXΩ. So on the point of thermal sweeping (∆ = HX ), Equation 9 tells us that δ = HX so that δ ≪ Rc. Thus we proceed, in the limit δ ≪ Rc and retain terms in geff to first order in δ/R: we solve for radial hydrostatic equilibrium in the dust- heated disc, where the approximation δ ≪ Rc ensures that the dust-heated disc is approximately isothermal at a tem- perature Tc. (We note that while Tc varies with radius on a scale length ∼ Rc, making the isothermal approximation is formally equivalent to only retaining terms first order in δ/R.) This then yields a Gaussian profile for the dust-heated disc such that P (δ) = Pc exp"−(cid:18) 1 2 δ Hc(cid:19)2# (10) where Pc is the pressure at Rc and Hc is the hydrostatic scale height of the disc in the vertical direction at R = Rc. The innermost extent of the dust-heated disc is thus set by the condition of pressure balance with the X-ray heated region (pressure PX ) - which we set as Rhole for convenience - and is given by: δ(Rhole) = 2Hcslog(cid:18) Pc PX(cid:19)2 (11) c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Applying the condition for thermal sweeping (∆ = HX ) allows Equation 10 to be re-written as: Pc PX Hc (cid:19)2# = exp" 1 2 (cid:18) HX = exp(cid:18) TX 2Tc(cid:19) (12) We can qualitatively understand this result by consid- ering a case where TX /Tc is fixed. If Pc/PX is high then the X-ray heated gas can only achieve pressure equilibrium with the cold gas at a point that is relatively far from Rc, so the effective gravity is large (Equation 8) and the ra- dial scale length correspondingly small (Equation 9). How- ever, as Pc/PX is reduced, the X-ray heated region is pushed closer to Rc, where geff is low and the radial scale length be- comes larger. When Pc/PX is reduced sufficiently, the ther- mal sweeping criterion (∆ = HX) is activated. This minimum value of Pc can be readily converted into a minimum surface density criterion since in hydrostatic equilibrium Pc ∝ ΣccΩ, where cc is the sound speed at Rc. Thus we have: ΣT S = √2π PX ccΩ 2Tc(cid:19) exp(cid:18) TX (13) However, Equation 13 only allows us to evaluate ΣT S (for a given radius in the disc and corresponding temperature of the dust heated disc) if one also knows TX and PX . It is now necessary to consider the form of the relationship between ξ and T which we use to assign X-ray temperatures. In the dense conditions at the base of the X-ray heated flow (i.e. at low ionisation parameter), the relationship is nearly flat at a temperature of ∼ 100K and then steepens at ξ = ξmin, with pressure then declining mildly as the temperature is reduced below ∼ 100 K. Where the structure of the temperature- ionization parameter relation is determined by the falling X-ray heating efficiency with ionization fraction (Xu and Mccray 1991), in combination with cooling by lines (Owen et al. 2010). The number density on the cold side of the interface is such that the X-ray temperature at this point is equal to Tc since in this case the X-rays would be unable to heat this region above Tc. This condition sets the pressure at the interface; on the X-ray heated side of the interface the gas shares the same pressure but has a temperature of ∼ 100K, since this places it in the nearly isothermal portion of the T − ξ relationship. We can see from Figure 8 that a line of constant pressure (T ∝ ξ) that originates on the ξ − T curve at a typical dust temperature of 10 − 100 K crosses the curve at a value of ∼ ξmin, as indicated by our simulations (see Figure 7). We can therefore simply set PX and TX by requiring ξ = ξmin at this point, where comparing with Figure 8 and our simulations indicate a sensible choice for ξmin = 3×10−7 erg s−1 cm−1. Substituting this condition c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Characterising Thermal Sweeping 7 104 103 102 ] K [ e r u t a r e p m e T 101 10−8 10−7 10−6 10−5 10−4 10−3 10−2 Ionization Parameter [erg s−1 cm−1] Figure 8. Temperature plotted as a function of X-ray ionization parameter, the thin dashed lines show isobars. into 13 we obtain: ΣT S = √2πµcc(cid:18) LX TX ΩξminR2 LX 2Tc(cid:19) holeTc(cid:19) exp(cid:18) TX 100 K(cid:19)−1/2 1030 erg s−1(cid:19)(cid:18) T1AU 10 AU(cid:19)−1/4 100 K(cid:19)−1/2# = 0.14 g cm−2(cid:18) × (cid:18) M∗ × exp"(cid:18) Rhole 0.7 M⊙(cid:19)−1/2(cid:18) Rhole 10 AU(cid:19)1/2(cid:18) T1AU (14) (15) Note that in constructing 15 from 13 we have also assumed that the temperature in the dust-heated disc follows a R−0.5 profile (cf equation 6) and T1AU is the dust temperature at 1 AU defined in Equation 6. We can compare Equation 15 to our simulations by plot- ting ΣT S/LX for the two sets of simulations around a 0.1 & 0.7 M⊙ mass star. Given the result for ΣT S /LX presented above has no further dependence on X-ray luminosity, but is only a function of mass and radius then the simulation points should line up along the same line. This comparison is shown in Figure 9, where we plot the simulations (0.7 M⊙ - open points; 0.1 M⊙ - filled points) for the various X-ray luminosities to Equation 15 (0.7 M⊙ - dashed line; 0.1 M⊙ - solid line). We see the agreement is exceptionally good, for the full range of parameters explored (apart from the small radius runs at the highest X-ray luminosity that are col- umn limited, and are not covered by the model as discussed above). Figure 9 also clearly shows that thermal sweeping is more efficient around lower-mass stars, possibly indicating that it may become ineffective around higher mass stars, as the X-ray luminosity fails to increase with stellar mass around intermediate mass stars. Furthermore, the general form (Equation 14) could be applied to cases where an optically thin (to the X-rays) gap is created by a planet (e.g. Rice et al. 2006, Zhu et al. 2012) or by a combination of photoevaporation and a planet (Rosotti et al. 2013), to see if planet formation could trigger rapid disc dispersal through thermal sweeping. 5.2 Properties of observed transition discs Confident that our model presented above can accurately predict the surface density of discs that are unstable to rapid 8 Owen et al. −29.5 −30 −30.5 −31 ] 4 − m c 3 s [ ) L / Σ ( g o X S T 0 1 l −31.5 0.4 0.6 1 0.8 log 10 1.2 1.4 1.6 (Inner Hole Radius) [AU] ] U A i [ s u d a R e o H l r e n n I x a M 1.8 2 50 45 40 35 30 25 20 1028 1029 1030 X−ray Luminosity [erg s−1] 1031 Figure 9. Comparision between the thermal sweeping model pre- sented in Equation 15: plotted as ΣT S /LX verus radius, and the simulations presented in Section 4. The open points and dashed line are for the 0.7 M⊙ star and the filled points and solid line are for the 0.1 M⊙ star. The square, circular and triangular points are for X-ray luminosities of 2 × 1029, 2 × 1030 and 2 × 1031 erg s−1 respectively. clearing by thermal sweeping, we can now make observa- tional predictions for the inner hole radius at which thermal sweeping clears the disc. We use Equation 15 to re-analyse the disc population synthesis model performed by Owen et al. (2011), which was designed to match the evolution of disc fraction as a function of time. The model well reproduced a large fraction of transition discs with small holes and low accretion rates, and fully explains the population of transi- tion discs with low mm fluxes that Owen & Clarke (2012) identified as discs most likely to be transitioning from disc bearing to a disc-less state. However, in doing so the model also predicted a large population of 'relic' discs with large holes and no accretion. This population is yet to be observed and Owen et al. (2012) noted that this problem would be alleviated by the thermal sweeping mechanism. Thus we take the synthetic disc population run by Owen et al. (2011), and assume the disc is dispersed instanta- neously by thermal sweeping once the disc's surface density drops below ΣT S . We determine the distribution of max- imum cavity radius for transition discs created by photo- evaporation. The resulting radius distribution is shown in Figure 10, where we plot the maximum inner hole radius reached before thermal sweeping takes over for discs evolv- ing around a 0.7 M⊙ star with different X-ray luminosities. Figure 10 clearly shows that thermal sweeping prevents the formation of relic discs (i.e those with large holes and no accretion), explaining the lack of a significant popula- tion of transition discs observed to have large holes with no accretion. The trend of smaller hole radii reached around higher X-ray luminosity stars is due to the ability of thermal sweeping to begin at higher surface densities with higher X- ray luminosities. At low X-ray luminosities (< 1029 erg s−1) the hole radius is independent of X-ray luminosity: this is because both ΣT S and also the surface density of the disc at a given radius at the point of photoevaporative hole opening scale linearly with X-ray luminosity and so equality is always achieved at a fixed radius. At larger X-ray luminosities, the process of 'photoevaporation starved accretion' (Drake et Figure 10. Maximum inner hole radii reached by clearing discs before thermal sweeping takes over. Each point represent an indi- vidual disc model taken from the population synthesis calculation of Owen et al. (2011), that followed the evolution of a group of discs evolving under photoevaporation around a 0.7 M⊙ star with an X-ray luminosity drawn from the observed X-ray luminosity function (Gudel et al. 2007). al. 2009; Owen et al. 2011) causes a slightly sub-linear de- pendence on X-ray luminosity for the disc surface density at hole opening. Consequently the thermal sweeping radius declines somewhat with X-ray luminosity. However, at the highest X-ray luminosities this is countered by the fact that gap formation happens at earlier times and hence higher surface densities for stars with higher X-ray luminosities. Furthermore, the rapid dispersal of transition discs cre- ated by photoevaporation at radii < 45AU means that the ratio between the time spent in an accreting transition disc phase to a non-accreting transition disc phase is greatly en- hanced, compared to the photoevaporation model analysed in Owen et al. 2011. For example, our new thermal sweeping models predicts that the median model (LX = 1.1×1030 erg s−1 - Figure 9 of Owen et al. 2011), initially opens the hole at ∼ 2.5 AU and the disc is destroyed by thermal sweep- ing when Rhole ∼ 30 AU, with a remaining disc mass of a few Jupiter masses. Furthermore, it spends approximately 2 × 105 years as an accreting transition disc and 6.5 × 104 years as a non-accreting transition disc. Thus, the addition of thermal sweeping suggests that transition discs created by photoevaporation spend the majority of their life as accret- ing transition discs. Due to the variation of thermal sweeping with stellar mass, thermal sweeping can begin earlier around lower mass stars, and may become inefficient around inter- mediate mass stars, where the X-ray luminosity no longer increases with stellar mass above 1-2M⊙ (e.g. Guedel et al. 2007, Albacete Colombo 2007) 5.3 Possible limits for the time-scale of thermal sweeping One possible limit on the time-scale at which thermal sweeping proceeds through the disc is energetic. The multi- dimensional simulations that were initially used to investi- gate thermal sweeping enforced local radiative equilibrium. This condition meant energetic consideration were neglected (see discussion in Owen et al 2010 who demonstrated that c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? for standard X-ray photoevaporation the mechanical lumi- nosity of the wind is a comfortably small fraction (. 8%) of the X-ray luminosity of the source). So P dV work done escaping from the potential was not accounted for. The dy- namical time-scale on which thermal sweeping begins is not going to be limited by this consideration (because the mass at the inner edge is small). However, as most of the disc material is at large radius, it may take longer than the dy- namical time-scale at the inner edge for the disc material at such large radius to be fully heated up to the equilibrium temperature. We can estimate an 'energy-limited' time-scale for clearing to occur by comparing the gravitational poten- tial energy stored in the disc, to the rate at which the X-rays inject energy into the system. We can calculate the gravitational binding energy of a power-law disc of the form Σ = Σhole(R/Rhole)−1 as: Udisc = Z Rout Rhole R GM∗ 2πRdR Rhole(cid:19)−1 Σhole(cid:18) R Rhole(cid:19) = 2πGM∗ΣholeRhole log(cid:18) Rout ∼ 1040 ergs(cid:18) M∗ 0.7 M⊙(cid:19)(cid:18) Σhole 0.3 g cm−2(cid:19)(cid:18) Rhole 10 AU(cid:19)(16) Comparing this binding energy to the received X-ray lumi- nosity, we find an energy-limited clearing time-scale of: τclear = Udisc ǫLX = 2 × 103 years(cid:18) ×(cid:18) Rhole 10 AU(cid:19)(cid:18) M∗ LX 1030 erg s−1(cid:19)−1 0.7 M⊙(cid:19)(cid:18) (cid:16) ǫ 0.25(cid:17)−1 0.14 g cm−2(cid:19) Σhole (17) where ǫ represents the fraction of X-rays intercepted by the disc (the X-ray photosphere to the star occurs at height > H, Owen et al. 2010). This time-scale is still much shorter than the clearing time-scale for the inner disc by stan- dard photoevaporation ∼ 3 × 105 years (Clarke et al. 2001, Alexander 2006; Owen et al. 2011). Thus, we expect ther- mal sweeping to still be efficient in removing the outer disc. Moreover, since ΣT S ∝ LX , this energy limited time-scale is independent of X-ray luminosity. However, future multi- dimensional calculations that relax the assumption of radia- tive equilibrium will be required to fully assess the actual efficiency of the thermal sweeping process at large radius. 6 SUMMARY In this work we have investigated the rapid disc dispersal mechanism for holed disc 'thermal sweeping', introduced by Owen et al. (2012) which takes over from photoevapora- tion and destroys the remaining disc material on a short (. 104 year) time-scale once the surface density of the disc has dropped to sufficiently low values. We use the criterion suggested by Owen et al (2012) (involving equality of the disc's vertical scale height and the radial thickness of the X- ray heated bound gas at the rim of the holed disc) in order to define the point at which a disc will undergo rapid dispersal, using a large suite of 1D radiation-hydrodynamical simula- tions. In this way we identify the surface density at which thermal sweeping proceeds as a function of X-ray luminosity, c(cid:13) 2002 RAS, MNRAS 000, 1 -- ?? Characterising Thermal Sweeping 9 stellar mass and inner hole radius. Our results however do not replicate the analytic estimate presented by Owen et al 2012 since this was based on the assumption that the extent of the X-ray heated region at the disc's inner rim is set by a fixed absorption column. Instead we find (apart from at the largest X-ray luminosities and smallest radii) that the limits of X-ray heating are set by the ionisation parame- ter falling to the value ξmin, below which the temperature- ionisation parameter relation steepens strongly (see Figure 8), at which point the X-rays do not heat the gas above the local disc temperature. We use this result to derive a new criterion (Equation 15) for how the critical surface density for thermal sweeping varies with physical properties of disc systems and find that this provides an excellent match to the simulation results (see Figure 9). Our main findings are summarised below: (i) Thermal sweeping can rapidly clear the disc, once the disc mass drops below a few Jupiter masses and the inner hole is at sufficiently large radii & 10 AU. (ii) The critical surface density for thermal sweeping to proceed scales linearly with X-ray luminosity, and increases with inner hole radius. (iii) Thermal sweeping is considerably more efficient at clearing discs around lower mass stars and could possibly be- come ineffective around intermediate mass stars which have comparatively weaker X-ray emission. (iv) Using the derived condition for thermal sweeping, we show that transition discs created through photoevaporation are completely destroyed once their inner hole radii reach sizes > 20 − 40 AU. (v) The destruction of transition discs by thermal sweep- ing, prevents the formation of 'relic' discs meaning that the majority of transition discs created by photoevaporation should be found to be accreting, albeit modestly ( M∗ < 10−8 M⊙ yr−1). Future multi-dimensional radiation-hydrodynamic sim- ulations are required to understand the details of thermal sweeping. In particular, energy considerations raise ques- tions about how rapid thermal sweeping can be in heating the outer parts of the disc once runaway penetration is in progress. Additionally, penetration and heating of the disc by the FUV radiation field (e.g. Gorti & Hollenbach 2009; Gorti et al. 2009) may aid in thermal sweeping. The inclusion of FUV heating in hydrodynamic simulations still remains challenging, and we cannot speculate on the role played by FUV heating in thermal sweeping at this stage. ACKNOWLEDGEMENTS We are grateful to the referee for comments that improved the paper. We thank Giovanni Rosotti, Barbara Ercolano, Geoff Vasil, Emmanuel Jacquet and Tom Haworth for in- sightful discussions. JEO is grateful to hospitality from the IoA, Cambridge during the initial stages of the work. MHB acknowledges support of an NSERC summer research grant held at CITA during 2012. The calculations were performed on the Sunnyvale cluster at CITA which is funded by the Canada Foundation for Innovation. We would like to ac- knowledge the Nordita program on Photo-Evaporation in 10 Owen et al. Astrophysical Systems (June 2013) where part of the work for this paper was carried out. REFERENCES Albacete Colombo, J. F., Flaccomio, E., Micela, G., Sciortino, S., & Damiani, F. 2007, A&A, 464, 211 Alexander R. D., Clarke C. J., Pringle J. E., 2006, MNRAS, 369, 229 Alexander R. D., Armitage P. J., 2009, ApJ, 704, 989 Andrews S. M., Williams J. P., 2005, ApJ, 631, 1134 Andrews S. M., Williams J. P., 2007, ApJ, 671, 1800 Andrews S. M., Wilner D. J., Espaillat C., Hughes A. M., Dullemond C. P., McClure M. K., Qi C., Brown J. M., 2011, ApJ, 732, 42 Armitage P. J., Hansen B. M. S., 1999, Natur, 402, 633 Birnstiel T., Andrews S. M., Ercolano B., 2012, A&A, 544, A79 Calvet N., D'Alessio P., Hartmann L., Wilner D., Walsh A., Sitko M., 2002, ApJ, 568, 1008 Chiang E. I., Goldreich P., 1997, ApJ, 490, 368 Clarke C. J., Gendrin A., Sotomayor M., 2001, MNRAS, 328, 485 Clarke C. J., Owen J. E., 2013, arXiv, arXiv:1305.2483 Cieza L. A., et al., 2013, ApJ, 762, 100 D'Alessio P., Calvet N., Hartmann L., 2001, ApJ, 553, 321 Drake, J. J., Ercolano, B., Flaccomio, E., & Micela, G. 2009, ApJL, 699, L35 Dullemond C. P., Dominik C., 2005, A&A, 434, 971 Duvert G., Guilloteau S., M´enard F., Simon M., Dutrey A., 2000, A&A, 355, 165 Espaillat C., et al., 2010, ApJ, 717, 441 Ercolano B., Drake J. J., Raymond J. C., Clarke C. C., 2008, ApJ, 688, 398 Hayes J. C., Norman M. L., Fiedler R. A., Bordner J. O., Li P. S., Clark S. E., ud-Doula A., Mac Low M.-M., 2006, ApJS, 165, 188 Hern´andez J., et al., 2007, ApJ, 671, 1784 Kennedy G. M., Kenyon S. J., 2009, ApJ, 695, 1210 Kenyon S. J., Hartmann L., 1995, ApJS, 101, 117 Koepferl C. M., Ercolano B., Dale J., Teixeira P. S., Ratzka T., Spezzi L., 2013, MNRAS, 428, 3327 Luhman K. L., Allen P. R., Espaillat C., Hartmann L., Calvet N., 2010, ApJS, 186, 111 Mamajek E. E., 2009, AIPC, 1158, 3 Mathews G. S., Williams J. P., M´enard F., 2012, ApJ, 753, 59 Nayakshin S., 2013, MNRAS, 431, 1432 Owen J. E., Ercolano B., Clarke C. J., Alexander R. D., 2010, MNRAS, 401, 1415 Owen J. E., Ercolano B., Clarke C. J., 2011, MNRAS, 412, 13 Owen J. E., Clarke C. J., Ercolano B., 2012, MNRAS, 422, 1880 Owen J. E., Jackson A. P., 2012, MNRAS, 425, 2931 Owen J. E., Clarke C. J., 2012, MNRAS, 426, L96 Rosotti, G. P., Ercolano, B., Owen, J. E., & Armitage, P. J. 2013, MNRAS, 430, 1392 Skrutskie M. F., Dutkevitch D., Strom S. E., Edwards S., Strom K. M., Shure M. A., 1990, AJ, 99, 1187 Stone J. M., Norman M. L., 1992, ApJS, 80, 753 Strom K. M., Strom S. E., Edwards S., Cabrit S., Skrutskie M. F., 1989, AJ, 97, 1451 Wyatt M. C., 2008, ARA&A, 46, 339 Zhu Z., Nelson R. P., Dong R., Espaillat C., Hartmann L., 2012, ApJ, 755, 6 c(cid:13) 2002 RAS, MNRAS 000, 1 -- ??